You are on page 1of 40

Rings and ArithmeticNotes for the Reader

The notes I am posting here are due to Dr Brian Stewart from last year with minor revisions.
I intend to revise them further during Michaelmas term. These notes are a rough guide to
the contend of the course, you will probably nd it helpful to read also from books. After
each lecture I have added some true/false questions that will help you check that you have
assimilated the basic notions. I advice you to attempt these questions before working on the
problem sheets.
Please tell me about any errors by sending an email to papazoglou@maths.ox.ac.uk
Panos Papazoglou
1 Commutative Rings with Identity
You have already encountered rings in the rst-year course Groups, Rings and Fields. In
this course we will concentrate on an important subclass of rings: commutative rings with
identity. In this setting we will be able to generalize in a more abstract framework results
and notions from the familiar case of the integers, Z. For example we will generalize Euclids
algorithm, the notion of prime numbers and we will prove a more general version of the
fundamental theorem of arithmetic.
Whenever you are in doubt about what a theorem means, then the example to keep
returning to is the ring of integers, Z.
1.1 The Denition
Denition 1. A commutative ring with identity is a non-empty set R, equipped with
two operations + : R R R, : R R R satisfying the following axioms:
R1 (R, +) is a commutative group (and we denote by 0 its zero element).
R2 a b = b a for all a, b R. [ is commutative]
R3 a (b c) = (a b) c for all a, b, c R. [ is associative]
R4 There exists an element 1 = 0 such that a 1 = a for all a R. [identity for ]
R5 a (b + c) = a b + a c for all a, b, c R. [ distributes over +]
1.1.1 Notation
We denote by a the inverse element of a, so a + (a) = 0.
We write

ab for a b
a b for a + (b).
1.1.2 Comments
Note that the axioms are satised in Z and capture (we hope) the algebraic essence of the
integers. (What we have not attempted to build in is the order enjoyed by the integers.)
In the rst year rings were dened as sets with +, satisfying (R1), (R3) and (R5) where
we postulate also that (b + c) a = b a + c a.
Recall the axioms dening abelian groups: (A, +) is an abelian group if the following
hold:
(A1) a + b = b + a for all a, b A
(A2) a + (b + c) = (a + b) + c for all a, b, c A
(A3) There is an element 0 A such that a + 0 = a for all a A
(A4) For any a A there is an element a such that a + (a) = 0.
We dont propose to repeat work done in the rst year Groups course; for example we
will use without fuss such facts as the zero element is unique.
In a similar vein we will not repeat work done in the rst year Analysis I course; much
of what we did when we investigated the real numbers from an axiomatic point of view
can be used here too. For example, we dont mean to fuss at all when we use facts like
(b + c) a = b a + c a.
We will also follow the practice we have learned in Groups and Vector Spaces: all zero
elements will be denoted by 0, and all identity elements by 1.
Some authors do not require that 1 = 0. Note that there is only one ring that does not
satisfy this: the trivial ring 0.
2
We state in the following lemma some familiar computational rules that hold also for
rings.
Lemma 1.1.1. Let R be a commutative ring with identity. Then for any a, b R the
following hold
1. a0 = 0a = 0.
2. (a) = a.
3. a(b) = (a)b = (ab).
4. (a)(b) = ab.
5.(1)a = a
Proof. The proofs are quite straightforward. We give some hints below. To show 1 note that
a0 = a(0 + 0) = a0 + a0 a0 = 0
Assertion 2 was proven last year in the groups course. For 3 we have
a(b) + ab = a(b + (b)) = a0 = 0
hence a(b) is the additive inverse of ab, ie a(b) = (ab). We leave 4,5 as exercises.
1.1.3 An example
We will deal with examples later, but here is an example rather dierent in avour from the
integers Z. For R take the set of diagonal nn matrices with real entries; for the operations
take the usual matrix operations. Then we have a commutative ring with identity.
1.2 Two important classes
We begin with two denitions.
Denition 2. A non-zero element z of a commutative ring with identity R is called a zero-
divisor if there exists a non-zero element w R such that zw = 0.
For example, in the commutative ring with identity consisting of the 2 2 diagonal
matrices with real entries the element

1 0
0 0

and its friend

0 0
0 1

are both zero-divisors.


More generally, the zero-divisors are precisely the

a 0
0 d

with a = 0 or d = 0.
Denition 3. An element u of a commutative ring with identity R is called a unit if there
exists an element v R such that uv = 1. In this case we say that v is the inverse of u.
We denote the set of units by R

, which we call the (multiplicative) group of units of R.


We see easily that (R

, ) is indeed an abelian group. We remark that if u is a unit then


u is not a zero divisor. Indeed if u is a unit then there is some v such that uv = 1. If au = 0
then (au)v = 0v a(uv) = 0 a = 0. So u is not a zero divisor.
For a trivial example, in any R the identity is a unit. For a more elaborate example take
again for R the 2 2 diagonal matrices with real entries. Then the units are precisely the

a 0
0 d

with ad = 0.
3
1.2.1 Integral Domains
We can now dene this important class of rings.
Denition 4. We say that a commutative ring with identity R is an integral domain if
there are no zero-divisors.
For example, Z is an integral domain; other examples appear below.
1.2.2 Fields
Even more specialized are the elds.
Denition 5. We say that the commutative ring with an identity is a eld if every non-zero
element is a unit.
Since units are not zero divisors elds are integral domains.
For example, R is a eld.
Note that this denition of eld (setting elds in a more general picture) is completely
consistent with the denition used in the Linear Algebra course.
We note that
elds integral domains commutative rings with identity
If R is an integral domain one can dene the eld of fractions K of R. This is a
generalization of the construction of Q from Z. We outline this in the appendix.
1.3 Examples, Non-examples and Nearly Examples
1.3.1 The integers
We repeat: the integers, with the usual operations form a commutative ring with identity.
1.3.2 The integers mod n
The set of integers mod n, Z
n
is a commutative ring with identity.
1.3.3 Polynomials over a eld
Let K be any eld. Then the set of polynomials K[X], with the usual polynomial denitions
of addition and multiplication forms a commutative ring with identity.
Next to Z these are our most important examples of commutative rings with identity.
1
1.3.4 Some elds
Here are examples of elds that we have met already: the rational eld, Q, the real eld R,
the complex eld C.
There are other, more exotic elds: many we will construct later as examples of theorems
we prove. For the moment you may like to check that Q[

2] := a + b

2 [ a, b Q is a
eld; and that C(X) :=
f(X)
g(X)
[ f, g polynomials with complex coecients, g = 0 is a eld.
1
Z is what number theorists study, the polynomial rings are what the geometers study; the similarity
between the structures goes very deep.
4
1.3.5 Not quite examples
The set of nn matrices with entries from a eld K is not a commutative ring with identity;
it fails the commutative requirement. But suitably adapted much of what we say and prove
can be adapted to this situation. Various subsets, however, yield useful examples.
The set of even integers is not a commutative ring with identity; it fails the identity
requirement. Again some of what we say can be suitably adapted to this sort of situation.
1.4 Subrings
Denition 6. Let R be a commutative ring with identity. A subset A R is said to be a
subring [more properly, a sub-(commutative ring with identity)] if it contains the identity
and is a commutative ring with identity under the same operations.
For example, Z is a subring of Q.
Just as for subgroups we have a
Proposition 1.4.1 (Test for Subringhood). Let R be a commutative ring with identity. Then
A R is a sub-(commutative ring with identity) if and only if
(i) 1 A;
(ii) if a, b A then (a b) A;
(iii) if a, b A then ab A.
Proof. The proof is just as for groups or vector spaces; these criteria guarantee that the
operations restrict to operations on A and then the fact that the axioms which hold for all
elements of R certainly hold in A.
As an application: the only sub-(commutative ring with identity) of Z is Z.
Note 1. Note that if by ring we mean, as most authors do, a system satisfying our axioms
(R1), (R3) and distributivity for , + , then there are many subrings of Z: for each d Z the
set dZ := dr [ r Z is a sub-ring, but has no identity. It is therefore sometimes important
to adopt the tedious sub-(commutative ring with identity) language.
1.5 Direct Products
This is an easy recipe to make new rings from old.
Proposition 1.5.1. Let R
1
and R
2
be commutative rings with identity. Then the set
R
1
R
2
:= (x
1
, x
2
) [ x
i
R
i
, i = 1, 2 is a commutative ring with identity under the
coordinatewise operations:
(i) the zero element is (0, 0);
(ii) (a
1
, a
2
) := (a
1
, a
2
);
(iii) (a
1
, a
2
) + (b
1
, b
2
) := (a
1
+ b
1
, a
2
+ b
2
);
(iv) the identity element is (1, 1);
(v) (a
1
, a
2
) (b
1
, b
2
) := (a
1
b
1
, a
2
, b
2
).
5
Proof. Trivial.
We usually denote this ring by R
1
R
2
(or sometimes just R
1
R
2
).
For an example, take R R. This is a commutative ring with identity. Considered just
as an additive group it is isomorphic to C; but as rings they are very dierent, C has no
zero-divisors, but every R
1
R
2
has.
1.6 Polynomial Rings
This is another recipe to make new rings from old. Let R be a commutative ring with
identity. A polynomial over R in the indeterminate X is a formal expression of the form
a
0
+ a
1
X + ... + a
n
X
n
where a
0
, a
1
, ..., a
n
R. The elements a
0
, a
1
, ..., a
n
are called coecients of the polynomial.
If p(X) = a
0
+ a
1
X + ... + a
n
X
n
and a
n
= 0 we say that the degree of p(X) is n (if p(X)
is the zero polynomial then the degree is not dened). We add and multiply polynomials in
the familiar way; if
p(X) =
n

i=0
a
i
X
i
, q(X) =
k

i=0
b
i
X
i
then we dene their sum by
p(X) + q(X) =

i=0
(a
i
+ b
i
)X
i
where by convention a
i
= 0 if i > n and b
i
= 0 if i > k. We dene the product p(X)q(X) to
be the polynomial
r(X) =
n+k

t=0
c
t
X
t
where
c
t
=
t

i=0
a
i
b
ti
It is easy to see that with this operations the set of polynomial with coecients in R
becomes a commutative ring with identity denoted by R[X]. We give in an appendix to this
section a more formal denition of R[X], which has the advantage that it gets rid of the
mysterious indeterminate X.
We can repeat the process, and manufacture for example R[X][Y ], which we usually
abbreviate to R[X, Y ]. The study of real plane curves is essentially the study of this ring.
Remark. One shouldnt confuse polynomials in R[X] with functions f : R R. For ex-
ample there are nitely many functions f : Z
2
Z
2
but innitely many distinct polynomials
in Z
2
[X].
6
1.6.1 Power Series Rings
If R is a commutative ring with identity we can consider also innite formal expressions of
the form

i=0
a
i
X
i
, a
i
R
Such an expression is called a power series over R. Note that here convergence is irrelevant,
these are just formal expressions. Addition and multiplication are dened again in the
obvious way. The set of all power series over R is a commutative ring with identity, denoted
by R[[X]].
1.7 Important: Notation
All the rings in this course are commutative rings with identity. We will from
now on usually just say ring. We will say subring and mean sub-(commutative
ring with identity) and later we will speak of ring homomorphism and mean
homomorphism of commutative rings with identity And so on.
7
1.8 Appendix
Integral domains and elds
Let R be an integral domain. We consider the set of pairs:
S = (a, b) : a, b R, b = 0
We want to see these pairs as fractions in R. However we know from the example of Q that
dierent fractions may represent the same number. So we dene an equivalence relation:
(a, b) (c, d) if ad = bc
It is easy to see that is indeed reexive, symmetric and transitive. We denote the equiva-
lence class of (a, b) by
a
b
and we consider the set
K =
a
b
: (a, b) S
We dene now addition and multiplication on K.
a
b
+
c
d
=
ad + bc
bd
and
a
b

c
d
=
ac
bd
It is easy to check that these operations are well dened on equivalence classes. One
veries easily that K is a commutative ring with identity, for example we dene
0 :=
0
1
, 1 :=
1
1
and we check that axioms R1-R5 hold. One sees further that K is a eld as
a
b

b
a
= 1
One can see R as a subring of K via the identication
a =
a
1
Polynomials over rings
We give here a dierent denition of the ring of polynomials over a ring in order to
demystify the unknown.
Proposition 1.8.1. Let R be a commutative ring with identity. Then the set of sequences

(a
k
)

k=0
: a
k
R, only a nite number of the entries a
k
non-zero

is a commutative ring with identity under the operations:


(i) the zero element is (0, 0, . . . );
(ii) (a
k
)

k=0
:= (a
k
)

k=0
;
8
(iii) (a
k
)

k=0
+ (b
k
)

k=0
:= (a
k
+ b
k
)

k=0
;
(iv) the identity element is (1, 0, 0, . . . );
(v) (a
k
)

k=0
(b
k
)

k=0
:= (

r+s=k
a
r
b
s
)

k=0
.
What has this got to do with polynomials in X? Well, write X := (0, 1, 0, . . . 0),
and note that X
2
= (0, 0, 1, 0, . . . ) and so on. With that in place we then recover all
polynomials: for example, (a
0
, a
1
, a
2
, a
3
, 0, 0, . . . ) = a
0
+ a
1
X + a
2
X
2
+ a
3
X
3
.
With this choice of name X for (0, 1, 0, 0, . . . ) we call this new ring R[X]. If we called
(0, 1, 0, . . . ) by the name Y wed call the new ring R[Y ].
We can perform the same construction on the set of all sequences. In that case we get
the power series ring denoted by R[[X]]. Note that this is algebra, theres no question
of convergence.
Proof. It is easy to check that the set is closed under the operations. The addition axioms
are trivial. For this convolution multiplication the axioms are slightly tedious to check, but
not dicult.
9
Which of the following are true?
1. If a, b, c are non zero elements of a ring R then ab = ac implies that b = c.
2. If a, b, c are non zero elements of an integral domain R then ab = ac implies that b = c.
3. If R is a nite integral domain and a R, a = 0 then for some n 1, a
n
= 1.
4. If a, b are non zero elements of an integral domain R then the equation ax = b has
exactly one solution.
5. Any ring has nitely many units.
6. Every subring of a eld is a eld.
7. If R is an integral domain then R[X] is also an integral domain.
8. Z[X] has innitely many subrings.
9. If p(X), q(X) Z
6
[X] and deg p(X) = 2, deg q(X) = 3 then deg(p(X)q(X)) = 5.
10. If A, B are commutative rings with identity and A B then A is a subring of B.
10
2 Ideals, Quotients, Homomorphisms
This section is a brief review of some of the key ideas of any algebraic system: the manufacture
of quotient structures and the analysis of the homomorphisms.
2.1 Ideals
We begin with two denitions.
Denition 7. We say that a subset I of a ring R is an ideal and write I R if
(i) I is a subgroup under +;
(ii) for all a R and all i I we have ai I.
Examples. 0 and R are always ideals. If K is a eld then the only ideals of K are 0
and K. If n Z then the set of all multiples of n, nZ is an ideal of Z. Generally if R is a
ring and a R then aR = ax : x R is an ideal of R. The ideal aR is sometimes denoted
by < a > or (a) and is known as the principal ideal generated by a.
Proposition 2.1.1 (Test for ideals). Let R be a ring. Then I R is an ideal of R if
(i) 0 I;
(ii) if a, b I then (a b) I;
(iii) if a I, r R then ar I.
Proof. Easy.
Denition 8. Suppose that R is a ring and I R. For each a R we call
a := a + I := a + i [ i I
the coset of a.
The notation a is neat, but needs care if there are dierent ideals around as it doesnt
identify I in the way that the a + I notation does.
2.2 Quotients
Suppose that R is a ring and that I R, I = R. Then the set of cosets, which we denote by
R or by R/I can be made into a ring. We recap briey from the rst year work.
2.2.1 Operations
(i) as zero element, the class 0;
(ii) a := a;
(iii) a + b := (a + b);
(iv) as identity element, the class 1;
(v) a b := (a b).
At once we have a doubt: are these good denitions? Let us deal only with the last one.
Suppose that a = a

and b = b

; can we be sure that a b = a

? Of course we can, but it


takes a moment to check.
11
2.2.2 Axioms
Now that we have the operations we need to see whether the axioms are satised. Again let
us do only one, (R5) say. We need to prove that for all a, b, c R the following holds:
a (b + c) = a b + a c.
Well
LHS = a (b + c) assumption
= a ((b + c) denition of + on R
= (a (b + c) denition of on R
= (a b + a c) Axiom (R5) in R
= (a b + a c) denition of + on R
= (a b + a c) denition of on R (twice)
= RHS assumption
and we are done.
2.3 Applications
2.3.1 Modular Arithmetic
Clearly dZ := dn [ n Z is an ideal. We can therefore carry out the construction, and
manufacture the quotient ring; in this case the coset notation Z = Z/dZ helps us keep track
of the modulus d. We call this the ring of integers modulo d and we denote it by Z
d
.
We can use these rings to illustrate other things we have done: for example, when is Z/dZ
an integral domain? Let 0 = a Z/dZ, and suppose that for some 0 = b Z/dZ we have
that ab = 0. Then ab is divisible by d. If d = d
1
d
2
is composite this is always possible, just
take a := d
1
and b := d
2
. So for an integral domain d must be prime (or 0). In both these
cases (for detailed proof see later, but most of us think it is obvious) we do get an integral
domain.
Perhaps more interestingly, what are the units of Z/dZ? Here we are seeking those u for
which we can nd a v such that uv 1 is divisible by d. That is, given d and u we ask when
we can nd v and m such that vu+md = 1. If this is possible then the only common factors
of u and d are 1; conversely if the only common factors are 1 then by Euclids algorithm
we can nd v and m. That is,
(Z/dZ)

= u [ (u, d) = 1.
We denote the order of this group by (d); for example (12) = 4, since there are four units,
1 and 5.
2.3.2 The Complex Numbers
Suppose we start with the real numbers. We can construct the polynomial ring R[X]. In this
ring the multiples of (X
2
+1) form an ideal, which we write as 'X
2
+1`trivial calculation.
So we can form the quotient ring R[X]/'X
2
+ 1`.
12
What is it? It is, once again, the complex numbers C. Write i := X = X + 'X
2
+ 1`,
then every element can be expressed as a +bi (use the Division Algorithm to see that every
polynomial can be written as a + bX + g(X)(X
2
+ 1)). We have that
i
2
=

2
= X
2
= X
2
+ 1 1 = X
2
+ 1 1 = 0 1 = 1.
Weve now provided a theoretical underpinning for ideas like adjoin a new number whose
square is 1.
2.3.3 The square root of 2
For a moment suppose we are Ancient Greeks. With much hard work we have constructed
(in our own way) the rational eld Q. Then we start drawing right-angled triangles and try
to nd the length of the hypotenuse of the isoceles right-angled triangle of side 1. To our
horror we nd we need a number whose square is 2, and of course we have good proofs that
no such rational number exists. Our construction above would save the day: just look at the
ring Q[X] and consider the ideal 'X
2
2` consisting of the multiples of X
2
2. The coset
:= X +'X
2
2` in the quotient Q[X]/'X
2
2` is the number we are looking for.
(There is much more to be said here: see later in the course.)
2.4 More about ideals
Suppose that R is a ring, and that I, J R. The following are easy to check.
Proposition 2.4.1. The set I J is an ideal of R, and whenever K R with K I and
K J we have that K I J.
Proposition 2.4.2. The set I + J := i + j [ i I, j J is an ideal of R, and whenever
K R with I K and J K we have that I + J K.
Proposition 2.4.3. The set I J :=

r
i
r
j
r
[ i
r
I, j
r
J is an ideal of R, and
I J I J.
Denition 9. An ideal I of a ring R is said to be maximal if I = R and I J R implies
that I = J or J = R.
Theorem 2.4.4. Let R be a ring and I R. Then R/I is a eld if and only if I is maximal.
Proof. Suppose that I is maximal, and that x I. Then J := I+ < x >= I. Clearly J R,
so J = R. As 1 R we can write 1 = i + a x for some i I and a R. Then 1 = i + ax,
or ax = 1, and we have found an inverse for x.
Suppose that R/I is a eld, and that I J R with I = J. We can therefore choose
x J ` I. Then x = 0, and so has an inverse a. That is, ax 1 = 0, so that ax 1 I. As
ax J and I J we get that 1 J. For any t R then, we get t = t 1 J: hence R = J.
Since R/I has at least two elements I = R.
Denition 10. An ideal I of a ring R is said to be prime if I = R and xy I implies that
either x I or y I.
13
Theorem 2.4.5. Let R be a ring and I R. Then R/I is an integral domain if and only if
I is prime.
Proof. Suppose that I is prime, and that xy = 0. Then xy = 0, so xy I. It follows that
either x = 0 or y = 0. So R/I is an integral domain.
Suppose that R/I is an integral domain. Then R/I has at least two elements so I = R.
Suppose now that xy I. Then xy = 0 so xy = 0. As R/I is an integral domain x = 0 or
y = 0. So either x I or y I. Hence I is prime.
Corollary 2.4.6. Let R be a ring and I R. If I is maximal then I is prime.
Proof. If I is maximal then R/I is a eld. So R/I is an integral domain, hence I is prime.
2.5 Constructing new elds
One can use theorem 2.4.4 to construct examples of elds.
Let K be a eld. Recall that a polynomial of positive degree p(X) K[X] is called
irreducible if there are no polynomials of positive degree f(X), g(X) K[X] such that
p(X) = f(X)g(X).
In the rst year course you dened the highest common factor of two polynomials f, g
and you saw that there exist polynomials m, n such that
hcf(f, g) = mf + ng
The proofs were done only in the case of R[X] but the same proofs apply for any eld K.
Proposition 2.5.1. Let K be a eld, p(X) K[X] and I =< p(X) >. Then I is maximal
if and only if p(X) is irreducible. It follows that the quotient ring K[X]/I is a eld if and
only if p(X) is irreducible.
Proof. Assume that I is a maximal ideal. Let
p(X) = a(X)b(X)
be a factorisation of p(X). Clearly I < a(X) >. Since I is maximal either < a(X) >= I
or < a(X) >= R.
If < a(X) >= I then a(X) = p(X)q(X) for some q(X) K[X] so
p(X) = p(X)q(X)b(X)
and q(X)b(X) = 1. It follows that deg b(X) = 0.
If < a(X) >= R then 1 = a(X)q(X) for some q(X) K[X], so deg a(X) = 0. We
conclude that either deg a(X) = 0 or deg b(X) = 0. So p(X) is irreducible.
Assume now that p(X) is irreducible. Lets say that I J where J is an ideal of R,
J = I. Let f(X) J ` I. Since p(X) is irreducible hcf(f(X), p(X)) = 1. So there are
a(X), b(X) K[X] such that
a(X)p(X) + b(X)f(X) = 1
Since p(X), f(X) J we have that 1 J. So J = R and I is maximal.
Finally we remark that by theorem 2.4.4, K[X]/I is a eld if and only if I is maximal.
So K[X]/I is a eld if and only if p(X) is irreducible.
14
Remark 2.5.2. A polynomial of degree 2 or 3 in K[X] (K a eld) is irreducible if and only
if it has no roots in K.
Example 2.5.3. The quotient ring
Z
3
[X]/ < X
2
+ 1 >
is a eld.
Proof. Indeed using the previous proposition it suces to show that X
2
+ 1 is irreducible.
However X
2
+ 1 has no roots in Z
3
since 0
2
+ 1 = 1, 1
2
+ 1 = 2, 2
2
+ 1 = 2 in Z
3
.
2.6 Homomorphisms
When we study algebraic objects the appropriate maps to consider are the maps that pre-
serve the structure of these objects. So for K-vector spaces the appropriate maps are linear
transformations, for groups it is the group homomorphisms and so on. So we make the
following denition.
Denition 11. Let R and S be commutative rings with identity
2
. We will say that a map
f : R S is a homomorphism if, for all x, y R,
(i)f(1) = 1; (ii)f(x + y) = f(x) + f(y); (iii)f(x y) = f(x) f(y).
For an example, take R := R[x] and S := C and let for any a C,
a
be the evaluation
map,
a
:

N
0
c
k
x
k

N
0
c
k
a
k
. This is a homomorphism.
For another example, take R := Z and for any d Z let S := Z/dZ. Then the map
: R S dened by : n n (which maps each n to its equivalence class modulo d) is a
homomorphism. Our construction of the quotient Z/dZ achieved precisely this.
For a non-example, consider the map p : Z Z Z Z given by p(n
1
, n
2
) = (n
1
, 0).
This satises the conditions (ii) and (iii), for being a homomorphism: but fails to map the
identity to the identity.
We also make the following denition.
Denition 12. Let f : R S be a homomorphism of rings. We say that f is an isomor-
phism if f is 11 and onto. In this case we write R

= S.
2.6.1 The Kernel
Denition 13. Let f : R S be a homomorphism of rings. We say that f
1
(0) :=
z [ f(z) = 0 is the kernel of the homomorphism. Sometimes we denote it by ker f.
Suppose that we have a homomorphism f : R S of rings. Which elements get mapped
to the same place? Well, it is clear that mapping to the same place is an equivalence relation
on R. f(x) = f(a) if and only if f(x a) = f(x) f(a) = 0, and so using x = a + (x a)
we have
x [ f(x) = f(a) = a +z [ f(z) = 0 = a + ker f.
Note that we have at once a good test for f to be one-to-one: this is equivalent to
ker f = 0.
2
It matters here, so we emphasise the with identity.
15
Lemma 2.6.1. Let f : R S be a homomorphism of rings, with kernel K. Then K is an
ideal of R.
Proof. Clearly f(0) = 0 so 0 K. If a, b K then
f(a b) = f(a) f(b) = 0
so a b K. We conclude that K is a subgroup under +. Also if a K and r R then
f(ra) = f(r)f(a) = f(r)0 = 0
so ra K. We conclude that K is an ideal.
2.7 The Image
Suppose that we have a homomorphism f : R S of rings.
For example, we might be looking at
0
: Q[x] C. In this case there is a lot of S = C
which is quite irrelevant to the homomorphism. All that really matters is the part of S
consisting of elements mapped from R.
We make the denition.
Denition 14. Let f : R S be a homomorphism of rings. We say that
f(R) := y S [ for some x R, f(x) = y
is the image of the homomorphism. Sometimes we denote it by imf.
Note that at once we have a silly test for f to be onto: this is equivalent to imf = S.
Lemma 2.7.1. Let f : R S be a homomorphism of rings, with image f(R). Then f(R)
is a subring of S.
Proof. We use the subring test. 1 = f(1) imf. If x, y imf then there are x
1
, y
1
R
such that x = f(x
1
), y = f(y
1
). So
x y = f(x
1
) f(y
1
) = f(x
1
y
1
) imf, xy = f(x
1
)f(y
1
) = f(x
1
y
1
) imf
It follows that imf is a subring of S.
2.8 The Isomorphism Theorem
We can now give a complete description of any homomorphism.
Theorem 2.8.1 (The Isomorphism Theorem for Commutative Rings with Identity).
Let R and S be commutative rings with identity, and let f : R S be a homomorphism.
Then ker f is an ideal of R and imf is a subring of S.
Moreover, f : R/ ker f imf where f : x f(x) is a well-dened isomorphism.
Proof. We have already shown the rst part in lemmas 2.6.1, 2.7.1. Clearly if x = y then
f(x) = f(y). So f is well dened. It is obvious that f is onto. Finally f(x) = 0 if and only
if f(x) = 0 if and only if x ker f if and only if x = 0. So f is also one-to-one. Hence an
isomorphism.
Just as for groups, we usually write the Isomorphism Theorem much more briey: some-
thing like Let f : R S be a homomorphism; then R/ ker f

= imf.
We remark that there is an Isomorphism Theorem for vector spaces too, but at least in
the nite dimensional case it doesnt say more (or less) than the RankNullity Theorem.
16
2.8.1 A Key Example Evaluation
Let R := R[X] and let S := C; consider the evaluation homomorphism
a
: R S given by

a
:

N
0
c
k
X
k

N
0
c
k
a
k
.
What is the kernel of
a
? By denition, ker
a
:= (X) [ (a) = 0. So we must
decide which polynomials vanish at a. If a R, then the Remainder Theorem tells us the
answer: it is those polynomials which are exactly divisible by (X a); that is, ker
a
=
(X)(X a) [ R[X]. If a R then things are a bit more complicated. If (a) = 0
and has real coecients, then we also have that (a) = 0the complex conjugate is also a
root. Now, by the Remainder Theorem in C[X] we see that both (Xa) and (Xa) divide
; whence the product (X a)(X a) = (X
2
2aX + [a[
2
) divides . These conditions
are also clearly sucient, so that we have
ker
a
= '(X a)` := (X)(X a) [ R[X] if a is real,
and
ker
a
= '(X
2
2aX +[a[
2
)` := (X)(X
2
2aX +[a[
2
) [ R[X] if a is not real.
What is the image of
a
? If a R then surely (a) R. Moreover, given c R the
constant polynomial c evaluates to c. So we get that im
a
= R in this case. When a R
then we get more than R, for instance X evaluates to a, and so (X a) evaluates to
(a)i = 0. Now its clear we get every complex number = +i; it is the evaluation of the
real polynomial +

a
(X a).
What does the Isomorphism Theorem tell us? Well,
(i) if a R then R[X]/'(X a)`

= R;
(ii)if a R then R[X]/'(X
2
2aX +[a[
2
)`

= C.
(Note that any real monic quadratic polynomial with non-real roots can be expressed,
for some a, as (X
2
2aX +[a[
2
).)
2.8.2 AnotherModular Arithmetic
Let R := Z and S := Z
d
. The mapping n n of n to its coset modulo d is a homomorphism;
the image is all of S and the kernel is (unsurprisingly) dZ. The Isomorphism Theorem tells
us that Z
d

= Z/dZ which is not really very surprising given its denition!
17
Which of the following are true?
1. If I, J are ideals then I J is an ideal.
2. If F, K are elds and : F K is an onto homomorphism then is an isomorphism.
3. If R is an integral domain and I is an ideal of R then R/I is an integral domain.
4. There is a ring homomorphism f : Z
n
Z.
5. There is a ring homomorphism f : Z Z
n
.
6. There is a ring homomorphism f : Q Z.
7. There is a ring homomorphism f : Q Z
p
,where p is a prime.
8. The rings Z[i] and Z[X] are isomorphic.
9. C and R are isomorphic.
10. If an ideal I R contains a unit of the ring R then I = R.
11. If p Z is a prime number and xy (p) then either x (p) or y (p).
12. Q is an ideal of R.
13. If R is a ring and a
1
, ..., a
n
R then the set a
1
r
1
+ ... + a
n
r
n
: r
1
, ..., r
n
R is an
ideal of R.
14. < X > is a prime ideal of Z[X].
15. Every prime ideal of Z[X] is a maximal ideal.
16. If f : R S is an isomorphism of the rings R, S then I R is an ideal of R if and
only if f(I) is an ideal of S.
17. If I R is an ideal and f : R S is a ring homomorphism then f(I) is an ideal of S.
18. If I S is an ideal of S and f : R S is a ring homomorphism then f
1
(I) is an
ideal of R.
19. a + I R/I is a unit of R/I if and only if < a > +I = R.
18
3 The Chinese Remainder Theorem
3.1 Introduction
In the rst year course we learned about the Division Algorithm and Euclids Algorithm for
both the ring of integers Z and polynomial rings over elds, K. We are going to slightly
extend these results, proving one of the most versatile theorems of algebra, the Chinese
Remainder Theorem.
3.2 Abstract version
Although one of the most important things about the CRT is its eciency and practical use,
we start quite abstractly. This deals with rather dull technicalities once and for all, so that
when we come to concrete versions we can concentrate on what is interesting.
Lemma 3.2.1. Let R, S
1
, S
2
be rings and let f
1
: R S
1
and f
2
: R S
2
be homomor-
phisms. Then f : R S
1
S
2
by f : a (f
1
(a), f
2
(b)) is a homomorphism whose kernel is
ker f
1
ker f
2
.
Proof. As the operations on S
1
S
2
are dened coordinatewise it is trivial to see that f is
a homomorphism: for example,
f(a + a

) =

f
1
(a + a

), f
2
(a + a

f
1
(a) + f
1
(a

), f
2
(a) + f
2
(a

= (f
1
(a), f
2
(a)) +

f
1
(a

), f
2
(a

= f(a) + f(a

).
The kernel of f is
ker f = a [ f(a) = 0 = a [ (f
1
(a), f
2
(a)) = 0 = (0, 0) = a [ f
1
(a) = 0, f
2
(a) = 0 = ker f
1
ker f
2
.
Lemma 3.2.2. Let R be a ring, and let I
1
, I
2
R be ideals such that 1 I
1
+I
2
. Then the
map : R R/I
1
R/I
2
given by : a (a + I
1
, a + I
2
) is an onto homomorphism.
Proof. By the previous lemma applied to the natural homomorphisms R R/I
1
and R
R/I
2
we have that is a homomorphism. We must prove it is onto, so let (a
1
+ I
1
, a
2
+ I
2
)
be an arbitrary member of the codomain. By hypotheses we know that for some i
1
I
1
and
i
2
I
2
we have that 1 = i
1
+ i
2
. Consider (key step) the element x = a
2
i
1
+ a
1
i
2
. Then we
have that
x + I
1
= a
2
i
1
+ a
1
i
2
+ I
1
= a
1
i
2
+ I
1
= a
1
(1 i
1
) + I
1
= a
1
+ I
1
as a
2
i
1
and a
1
i
1
lie in the ideal I
1
. A similar argument deals with the coset modulo I
2
, and
we get that x has the required image.
We now have:
19
Theorem 3.2.3 (Abstract CRT). Let R be a ring, and let I
1
, I
2
R be ideals such that
I
1
+ I
2
= R. Then
R/(I
1
I
2
)

= R/I
1
R/I
2
where the isomorphism is the natural x + (I
1
I
2
) (x + I
1
, x + I
2
).
Proof. We consider the map : R R/I
1
R/I
2
given by a (a + I
1
, a + I
2
). By the rst
lemma this is a homomorphism with kernel I
1
I
2
. As R = I
1
+I
2
we have that 1 I
1
+I
2
so we can use the second lemma to get that is onto. Now apply the Isomorphism Theorem
to and get the result.
3.3 The CRT for Z
Originally the Chinese Remainder Theorem was the proposition that one could, if a, b are
coprime integers, and r, s are any integers, nd a solution to the simultaneous congruences
x r (mod a), x s (mod b).
That is only part of what we are now able to prove.
Theorem 3.3.1 (CRT for integers). Let a, b Z have highest common factor 1. Then
Z/abZ

= Z/aZ Z/bZ,
the isomorphism being the natural x + abZ (x + aZ, x + bZ).
Proof. By Euclids Algorithm we can nd R, S such that 1 = Ra +Sb, so that Z = aZ+bZ.
Also, we have that lcm(a, b) = ab/ hcf(a, b) = ab, so that aZ bZ = abZ. The theorem now
follows from the abstract version.
3.4 Applications
3.4.1 Simultaneous congruences
The result about simultaneous congruences, that if a, b are coprime integers, and r, s are any
integers, then we can nd a solution to the simultaneous congruences
x r (mod a), x s (mod b).
is an easy corollary of our CRT. For consider the pair of cosets (r + aZ, s + bZ); by the CRT
there exists a unique coset x + abZ such that (x + aZ, x + bZ) = (r + aZ, s + bZ).
Although ne in theory, this is not yet of practical use. How do we nd the solution
x? We have answered this implicitly. When we proved the ontoness of the map the key
consideration was that we could express 1 = i
1
+ i
2
with i
1
I
1
and i
2
I
2
; then we got x
as si
1
+ ri
2
.
So in practice, we use the (extended) Euclid Algorithma very ecient processto
calculate the integers R and S such that 1 = aR + bS; and the solution we then seek is, as
per the proof of our abstract theorem, x = aRs + bSr.
3.4.2 Eulers function
See problems sheet for an application.
20
3.4.3 Speeding up Arithmetic
Suppose again that a, b Z have highest common factor 1. By the CRT we have that
Z/abZ

= Z/aZ Z/bZ.
We can then carry out arithmetic calculations in Z/abZ in the following way:
Step 1 For each input x := x + abZ calculate x := x + aZ and x := x + bZ.
Step 2 Carry out the required calculations on the x in Z
a
to get the answer y; and on the
x in Z
b
to get y.
Step 3 Using the inverse of the isomorphism, calculate the value of y which maps to the
pair ( x, x).
Can this possibly be a good idea? The answer is a resounding Yes! Essentially we have
some setup costs we must pay once and for all: the calculation of integers R and S such that
Ra+Sb = 1 which allow us to compute the inverse isomorphism. This is not very expensive:
Euclids algorithm is very fast, requiring O(log a) steps. The reductions of step 1 are not
expensive, although an extra overhead. But the savings come in step 2: although we have
to carry out the calculations twice we do so in much smaller systems. If we make a careful
analysis well see that this really works.
3.5 The CRT for K[X]
If we look at our proof of the CRT for Z we will see that all we needed to use about the ring
Z and the elements a, b with hcf(a, b) = 1 were these facts, both consequences of Euclids
Algorithm:
(i) 1 = Ra + Sb for some R, S;
(ii) aZ bZ = abZ.
We have the Division Algorithm, highest common factors, and Euclids Algorithm in the
ring K[X], where K is any eld; we saw this in the rst-year course. Therefore we have, with
no more work to be done:
Theorem 3.5.1 (CRT for K[X] ). Let K be a eld and let f(X), g(X) K[X] have highest
common factor 1. Then
K[X]/f(X)g(X)K[X]

= K[X]/f(X)K[X] K[X]/g(X)K[X],
the isomorphism being the natural t + f(X)g(X)K[X] (t + f(X)K[X], t + f(X)K[X]).
3.6 An application
There would be no point in this, of course, if there were not important applications.
21
3.6.1 Interpolation
Suppose we have k + 1 distinct members a
i
of a eld K; and k + 1 arbitrary members b
i
of
K.
We can apply the CRT rst to the polynomials (X a
1
) and

i>1
(X a
i
), and then
inductively and get
K[X]/

(X a
i
)K[X]

K[X]/(X a
i
)K[X].
Let us look for a moment at K[X]/(X a)K[X]. What does a coset b + (X a)K[X]
represent? It is, as the Remainder Theorem tells us, the set of all polynomials which take
the value b at the point a.
So what the CRT tells us in part is this: there is a unique coset of polynomials
t(X) +

(X a
i
)K[X]
consisting of those polynomials which take the values b
i
at the points a
i
. By the Division
Algorithm we can then nd in the coset a unique t(X) of degree at most k with this property.
Once again, note that the CRT is actually constructive: Euclids Algorithm lets us
compute the inverse of the isomorphism eciently, and nd t(X) from the data (a
i
, b
i
),
i = 1, . . . , k + 1.
22
4 Divisibility in integral domains
We have until this moment used Z and K[X] both as rich sources of examples, and also
as prototypes: we have tried to develop ring theory in a way that captures the essential
properties of these structures. In this section we will try to generalize the divisibility and
factorization properties of these rings in a more general settings. For example we would like
to generalize the Fundamental Theorem of Arithmetic and the Euclidean Algorithm from
the case of Z to more general rings. We begin with introducing the appropriate terminology
in the abstract setting of integral domains.
4.1 Divisibility
Denition 15. Let R be a ring, and let a, b R. We say that b divides a (and write
b[a) if for some c R we have that a = bc. We will also call b a divisor of a in these
circumstances.
So for example the units are the divisors of 1, and everything divides 0.
Denition 16. Let R be a ring, and let a R. An element a

of R is called an associate
of a if for some unit u R we have that a

= ua.
We remark that this is an equivalence relation.
For example the associates of n Z are n. Note that unique factorization in Z is not
really unique, for example 6 = 2(3) = (2)3. Of course this non uniqueness up to a sign
is quite harmless. In the unique factorization theorem that we will prove later uniqueness
will fail up to units rather than signs, which is still not bad. As we want to think of 2 and
2 as the same factor of 6 associate elements are the same as far as divisibility goes for
general rings.
Remark 4.1.1. If R is an integral domain and a, b R are such that a[b and b[a then a, b
are associates.
Indeed if both a, b are 0 they are clearly associates. Otherwise lets say that b = 0. Since
a[b and b[a we have b = au and a = bv for some u, v R. Substituting a in the rst equality
we get
b = (bv)u b b(vu) = 0 b(1 vu) = 0 1 vu = 0
So u, v are units and a, b associates.
Denition 17. Let R be a ring, and let a R, a = 0. We say that a is irreducible in R
if a / R

and
a = xy =x is a unit, or y is a unit.
Denition 18. Let R be a ring, and let a R ` 0 but a R

. We say that a is prime


in R if
a[xy =a[x or a[y.
Essentially, irreducible elements cant be factorised further; prime elements are those
which only divide products of which they already divide one factor.
In general we have this:
Proposition 4.1.2 (Prime Irreducible). Let R be an integral domain, and let 0 = x R
be prime. Then x is irreducible.
23
Proof. Suppose that x = yz. Then x = x1 so that x[x = yz. By the denition of prime
we will get x[y or x[z; suppose the former, that y = xt say. Then x = yz = xtz, and so
x(1tz) = 0. As x = 0 we get tz = 1, and z is a unit. Hence x is by denition irreducible.
Denition 19. Let R be a ring, and let a, b R. We say that d is a highest common
factor of a, b if
(i) d[a and d[b;
(ii) e[a and e[b implies that e[d.
So for example in Z a highest common factor of 9 and 6 is (3); another is 3. Note that
in general there is no guarantee that highest common factors exist.
Proposition 4.1.3. Let R be an integral domain and let a, b R`0. Suppose that d
1
and
d
2
are highest common factors of a, b. Then d
1
and d
2
are associates.
Proof. It is clear that neither d
1
nor d
2
is zero, as a and b are multiples of each.
By condition (i) applied to d
1
we see that d
1
[a and d
1
[b. So apply condition (ii) to d
2
using e = d
1
to get that d
1
[d
2
. Similarly we get d
2
[d
1
. So by remark 4.1.1 we have that d
1
, d
2
are associates.
Proposition 4.1.4. Let R be an integral domain and let a, b R ` 0. Suppose that d
1
is a highest common factor of a, b and that d
2
is an associate of d
1
. Then d
2
is a highest
common factor too.
Proof. Suppose that d
2
= d
1
u for some u such that uv = 1.
For condition (i) note that a = d
1
x
1
and b = d
1
y
1
, so that a = d
2
(vx
1
) and b = d
2
(vy
1
).
For condition (ii) suppose that e[a and e[b. Then we have that e[d
1
, or d
1
= ez. Then
d
2
= d
1
u = ezu and we are done.
In some rings there is a sensible way to choose a particular highest common factorin Z
we usually choose the non-negative associate, in K[X] the monic associateand call it the
highest common factor. But often we can work quite comfortably with the uncertainty of
up to a unit multiple.
4.2 Euclidean Rings
In order to generalize the theorems of Arithmetic to rings we have to restrict to a special
class of rings: Euclidean rings.
4.3 Denition
What we must do is express abstractly what is going on in the Division Algorithm. We
divide a by b = 0 and get a quotient q, leaving a remainder r which is smaller in some
way (size, degree, . . . ) than the divisor b.
Denition 20. We say that R is a Euclidean Ring with Euclidean function d if
(a) R is an integral domain;
(b) the function d : R ` 0 N ` 0 satises:
24
(i) for all x, y R ` 0 we have d(xy) d(y);
(ii) for all a, b R with b = 0 there exist q, r R such that a = bq + r and r = 0 or
d(r) < d(b).
The element q = q(a, b) is called the quotient of a by b, and the element r = r(a, b) is called
the remainder.
In our customary rather slovenly way we will say R is a Euclidean ring when the function
d is so obvious as to be understood without mention. However, see the last example below
for a warning.
4.4 Examples
4.4.1 The Integers
The ring of integers, equipped with the function n [n[ on the non-zero elements, is a
Euclidean Ring: we have known this since we learned about division, and we proved that it
is true in the rst year course
3
.
4.4.2 Rational Polynomials
The ring Q[X], equipped with the degree function f deg f, is a Euclidean Ring: weve
known this since we learned about long division, and we proved that it is true in the rst
year course
4
.
Of course the same thing is true for polynomials over any eld.
4.4.3 The Gaussian Integers
Let Z[i] := a+bi [ a, b Z. We call these the Gaussian integers. This subset of C is clearly
a subring of C and as C has no zero-divisors it is actually an integral domain.
How can we measure size, and nd a Euclidean function d? The obvious choice is
d(a + bi) := [a + bi[
2
,
but does it satisfy the requirements?
Condition (i) is easy:
d() = [[
2
= [[
2
[[
2
= [[
2
d()
and as [[ = a
2
+ b
2
for integral a, b we get [[
2
1.
Condition (ii) is more complicated; the argument is important as it can be used for certain
other Z[

n] and not just Z[

1].
So let := a + bi and := c + di = 0 be in the ring. Then in Q[i] := x + yi [ x, y Q
we can rationalise the denominator and get that

=
a + bi
c + di
=
ac + bd
c
2
+ d
2
+
bc ad
c
2
+ d
2
i =: x + yi
say.
3
On the course website there will be a proof.
4
On the course website there will be a proof.
25
This is the exact quotient in Q[i], but what is the best we can do in Z[i]? The nearest
we can get is the number := m + ni, where m is the integer nearest x, and n the integer
nearest y; note that [x m[
1
2
and [y n[
1
2
.
Now

= x + yi = m + ni + (x m) + (y n)i
and multiplying by we get
= (m + ni) + ((x m) + (y n)i) .
Put as quotient q(, ) := m + ni Z[i]. As remainder we then would have
r(, ) := ((x m) + (y n)i) = q(, ) Z[i].
We now compute
d(((x m) + (y n)i) ) = [ ((x m) + (y n)i) [
2
= [ ((x m) + (y n)i) [
2
[[
2
= ((x m)
2
+ (y n)
2
) [[
2

1
2
[[
2
< d()
and see that condition (ii) is satised.
4.4.4 Fields
Let K be a eld, and dene d : K` 0 N ` 0 by d(x) = 1. It is then trivial to see that
we have a (very dull) Euclidean Ring where all the remainders are 0.
4.4.5 The Integers, but not as we know them
The ring of integers, equipped with the function
d(n) = the number of digits when [n[ is expressed in base 2,
is a Euclidean Ring.
For condition (i), note that

2
M
+ lower powers of 2

2
N
+ lower powers of 2

2
M+N
+ lower powers of 2

and so d(xy) = d(x) + d(y) d(y).


For condition (ii), all is clear if a = 0, or indeed if d(a) < d(b); just take q = 0 and
r = a. So argue by induction on d(a). Suppose that a =

2
M
+ lower powers of 2

, and
b =

2
N
+ lower powers of 2

with = 1 and = 1. We are assuming M N, so


consider a =

a 2
MN
b

+ 2
MN
b. The number

a 2
MN
b

requires fewer than


M = d(a) binary digits, so we can nd q and r such that

a 2
MN
b

= b q + r
and r = 0 or d(r) < d(b).
Taking q = q + 2
MN
(= q 2
MN
) gives what we need.
This example shows that we need to take care when we say R is a Euclidean Ring.
26
4.5 Units and Associates
If we are interested in factorisations the rst thing we need to deal with are the factorisations
of the identity: we must nd the units of the ring. So let R equipped with d be a Euclidean
ring.
Lemma 4.5.1. For all a R, d(a) d(1).
Proof. Condition (i) applied to a = a 1 gives this at once.
Lemma 4.5.2. For all units u R

, d(u) d(1).
Proof. Condition (i) applied to 1 = v u gives this at once.
Lemma 4.5.3. For all x R such that d(x) = d(1), we have that x is a unit.
Proof. Use condition (ii) to get q and r such that
1 = xq + r with r = 0, or d(r) < d(x).
If r = 0 then we have by hypothesis d(r) < d(x) = d(1); this contradicts Lemma 1. So we
get exact division, 1 = xq and x is a unit.
To summarise:
Proposition 4.5.4 (Units of a Euclidean ring). Let R equipped with d be a Euclidean ring;
then the group of units is given by
R

= x R [ d(x) = d(1) .
In fact we may arrange things so that we have d(1) = 1. For suppose that d(1) = k + 1
for k N. Weve just seen that for all a R we have that d(a) d(1), so the function

d : R ` 0 N ` 0 by

d : a d(a) k is well-dened. It is clear that it also satises
condition (i) and condition (ii) with the same quotient and remainder.
The following is also useful about a Euclidean ring R:
Lemma 4.5.5. Let u R

and a R. Then d(ua) = d(a).


Proof. Let v a be such that vu = 1. We then have by condition (i) that d(a) = d(vua)
d(ua) d(a); equalities rule, and d(a) = d(ua).
27
The following subsections, establishing that the integers and polynomial rings over elds
are Euclidean are included purely for completeness; they were covered in the rst year course.
Revision: integers
Let d : Z ` 0 N ` 0 be given by d(n) = [n[.
For condition (i) we can use the properties of and [n[ we developed in Analysis I and
see that d(xy) = [xy[ = [x[[y[ = [x[d(y) d(y).
For condition (ii) note that
a = bq + r (a) = b(q) + (r)
and [r[ = [ r[; so it is enough to deal with the case a 0.
We can argue by (strong) induction on a; the result is true for a = 0 if we take q = 0 and
r = 0. Indeed the result is true for a < [b[, just take q = 0 and r = a. For a [b[ note that
a = (a [b[) +[b[, so we can use the inductive hypothesis to get (a [b[) = b q +r with r = 0
or [r[ < [b[. Taking q = q 1 as case may be completes the proof.
Revision: polynomials
Let d : K[X] ` 0 N ` 0 be given by d(f) = deg(f), the degree of f.
For condition (i) note that if f =

m
k=0
a
k
X
k
and g =

n
k=0
b
k
X
k
, with a
m
= 0 and
b
n
= 0 then f g =

m+n
k=0
c
k
X
k
where c
m+n
= a
m
b
n
. Hence deg f g = deg f +deg g deg g.
For condition (ii), again it is clearly true if f = 0; just take q = r = 0. Otherwise note
that if the top coecient of f is = 0 and the top coecient of g is = 0, then
f = gq + r
1

f =

;
clearly deg

= deg f, deg

= deg g and deg

= deg r. Hence we may assume


that f, g are monic, that is have top coecients 1.
Now we may argue by induction on deg f; if deg f = 0 or more generally deg f < deg g we
put q = 0 and r = f and are done. Otherwise, note that f(X) =

f(X) X
deg fdeg g
g(x)

+
X
deg fdeg g
g(x), and that deg

f(X) X
deg fdeg g
g(x)

< deg f. By the inductive hypoth-


esis we can then nd q and r such that

f(X) X
deg fdeg g
g(X)

= g(X) q(X) + r(X)


with deg r < deg g or r = 0. Now put q(X) = q(X) + X
deg fdeg g
and we are done.
28
Which of the following are true?
1. If K is a eld then any non zero element of K divides any other non zero element of K.
2. If a, b are associates then a[b and b[a.
3. If a is irreducible and a, b are associates then b is also irreducible.
4. If a is prime and a, b are associates then b is also prime.
5. a[b if and only if (a) (b).
6. The ideal < X > + < Y > of R[X, Y ] is principal.
7. R[X, Y ] is a Euclidean domain.
8. If R is Euclidean ring with Euclidean function d then d(a) > d(1) for all a = 0, a R.
29
5 Factorisation in Euclidean Rings
5.1 Ideals are principal
We recall a trivial fact and a denition:
Lemma 5.1.1. Let R be a ring and a R; then the set aR := ar [ r R is an ideal of R.
Denition 21. Let R be a ring. The ideal I R is called principal if it is of the form aR
for some a R.
Denition 22. A ring R is called a principal ideal ring if every ideal is principal; it is
called a principal ideal domain if it is an integral domain and a principal ideal ring.
Here is the crucial property of Euclidean rings.
Theorem 5.1.2 (E.R. P.I.D.). Let R be a Euclidean ring; then R is a principal ideal
domain.
Proof. Let I R. If I = 0 then I = 0R, so suppose I = 0. Choose some a I with d(a)
as small as possible.
As I R we clearly have that aR I.
Suppose that x I; then x = qa + r for some q, r with r = 0 or d(r) < d(a). Now
r = x qa I + aR = I, so d(r) < d(a) is impossible. Hence x = qa aR.
5.2 Highest Common Factors
Theorem 5.2.1 (EDs have HCFs). Let R be a Euclidean domain, and let a, b R, with
b = 0. Then there exists a highest common factor d of a, b, expressible as d = ar + bs for
r, s R.
Proof. We mimic in a suitably abstract way the Euclid algorithm proof for Z.
So consider the ideals aR and bR of R. We can then form the ideal aR + bR. As every
ideal in R is principal (EDPID) we have that aR +bR = dR for some d. We claim that d
is a highest common factor; clearly d = ar + bs.
For condition (i) note that aR aR + bR = dR, so that a = a1 dR, yielding a = da

for some a

. Hence d[a; similarly d[b.


For condition (ii) suppose that e[a and e[b. Then a = ex so that a eR. Similarly b eR.
Then for all r, s we will get ar + bs eR since eR is an ideal. That is dR = aR + bR eR.
Hence d = d1 eR, and d = ed

for some d

. That is, e[d.


As in the proof we used only the fact that R is a PID we have this more general result:
Corollary 5.2.2 (PIDs have HCFs). Let R be a Principal Ideal Domain, and let a, b R,
with b = 0. Then there exists a highest common factor of a, b.
30
5.3 Factorisation
Proposition 5.3.1 (Irreducible Prime in a PID). Let R be a Principal Ideal Domain,
and let x R be irreducible. Then x is prime.
We rst prove an important lemma.
Lemma 5.3.2 (Eulers Lemma). Suppose that h[ab and h, a have 1 as highest common
factor. Then h[b
Proof. We have that 1 = hr + as, and that ab = hk for some r, s, k. Then
b = b(hr + as) = bhr + bas = bhr + hks = (br + ks)h
and we are done.
Now we prove the proposition.
Proof. Now suppose that x is irreducible and that x[ab. Suppose that x [a. Let d be a
highest common factor of x and a. Then d[x by denition, and so x = de for some e. As x is
irreducible we have that d or e is a unit. Suppose that e is a unit, with ef = 1. Then d = xf,
and so xf[a; then at once x[a, a contradiction. So we have that d is a unit, an associate of
1. hence 1 is also a highest common factor, and then we are done by the lemma.
The following is an abstract version of the concrete fact that in Z we cannot go on
factorising a number into smaller and smaller pieces.
Proposition 5.3.3 (PID ascending ideal chains terminate). Let R be a Principal Ideal
Domain, and I
k
R, and I
k
I
k+1
for k = 0, 1, . . . . Then for some N, I
k
= I
N
for all
k N.
Proof. Consider the subset I :=

k
I
k
; we prove that it is an ideal. Clearly 0 I
0
I.
Suppose that x, y I and a R; then x I
r
and y I
s
for some r, s 0. Hence with
n := max(r, s) we have that x, y I
n
R. Then x, x + y, r x I
n
I and we are done.
Now R is a PID, so we have that I = aRfor some a R. Then a = a1 aR = I =

k
I
k
,
so that a I
N
for some N, and indeed a I
k
R for all k N. Now we have that
aR I
k
I = aR and are done.
5.4 Unique Factorisation
This is an area where our intuition developed in Z and K[X] notoriously leads us astray in
more general domains.
Denition 23. We will say that the integral domain R is a Unique Factorisation Do-
main (UFD), or that it has unique factorisation, if
(i) every element 0 = a R is either a unit or can be written a = f
1
f
2
. . . f
M
as a product
of (a nite number of ) irreducible elements f
1
, f
2
, . . . , f
M
R;
(ii) if f
1
f
2
. . . f
M
= g
1
g
2
. . . g
N
for irreducibles f
1
, f
2
, . . . , f
M
and g
1
, g
2
, . . . , g
N
then M = N,
and for each k = 1, . . . , M there exists a unit u
k
such that after relabelling the g
k
we
have g
k
= u
k
f
k
.
31
As a trivial example, every eld is a UFD. From our experience we know that Z is a
UFD.
Theorem 5.4.1 (PID UFD). Let R be a Principal Ideal Domain. Then R is a Unique
Factorisation Domain.
Proof. There are two dierent things to be proved: the existence of factorisation into irre-
ducibles and the uniqueness.
For the existence, consider 0 = x R, and suppose that we cannot express x in the
required form. Clearly x cant be a unit, so x = yz where neither y nor z is a unit. If both
y and z are expressible as products of irreducibles so is x; so one of these, without loss y
cant be so expressed. Now put I
0
:= xR and I
1
:= yR. As x = yz we have that I
0
I
1
.
Moreover, if I
1
= I
0
we would have that y = y1 yR = xR, so that y = xa for some a.
Then x = yz = xaz, and as x = 0 we have 1 = az and z is a unitwhich it is not. So
we have found unequal ideals I
0
I
1
. But we can repeat this process, starting now with y,
and so construct eventually a properly ascending chain of idealswhich we have seen to be
impossible.
As for uniqueness, we prove something slightly stronger. We prove that if uf
1
f
2
. . . f
M
=
vg
1
g
2
. . . g
N
for irreducibles f
1
, f
2
, . . . , f
M
and g
1
, g
2
, . . . , g
N
and units u, v then M = N and
after relabelling f
k
and g
k
are associates. We do this by induction on M + N, the result
being vacuously true when M + N = 0.
As f
1
is irreducible it is prime, and so f
1
divides some g
k
; relabel so that f
1
[g
1
and g
1
=
f
1
u
1
say. But g
1
is irreducible, and so u
1
is a unit. We therefore have that uu
1
f
1
f
2
. . . f
M
=
vu
1
g
1
g
2
. . . g
N
= vf
1
g
2
. . . g
M
, and so (as R is an integral domain) (uu
1
)f
2
. . . f
M
= vg
2
. . . g
M
.
We can now apply the inductive hypothesis and get M1 = N 1 and after relabelling the
remaining f
k
and g
k
are associates.
Remark 5.4.2. We have shown the inclusions
Euclidean rings Principal ideal domains Unique factorisation domains
All these are proper inclusions. We will see in the exercises that Z[X] is a UFD but not a
PID. It can be shown that Z[
1+

19
2
] is a PID but it is not a Euclidean ring.
32
Which of the following are true?
1. Z[X] is a UFD.
2. Let f(X), g(X) Q(X). Then there is some h(x) Q(X) such that
< f(X) > + < g(X) >=< h(X) >.
3. Let f(X), g(X) Q(X). Then there is some h(x) Z(X) such that
< f(X) > + < g(X) >=< h(X) >.
4. Let a be a prime element in a PID. Then the ideal < a > is maximal.
5. If F is a eld and f : F[X] C is a homomorphism then ker f is a maximal ideal of
F[X].
33
6 The Euclidean Algorithm
In this section we revisit the Euclidean Algorithm we learned in the rst year, and discuss
how far it applies in general Euclidean Domains. We will see that the main part of the
algorithm is still in place.
6.1 Algorithms
Sometimes we dont want simple assertions about the existence of a mathematical object, not
matter how carefully theyve been proved: we want a way to nd it in practice. In the rst
year courses we met several algorithms which allowed us to nd or construct things: explicit
processes which can be carried out in a nite number of steps, and which are guaranteed to
deliver the object we seek.
For examplebut there are many others if you look closelywe have: the Gaussian
Elimination process provides algorithms for nding all solutions of a set of linear equations
(and hence inverting matrices where possible); the Division Algorithm for integers, and
for polynomials with rational coecients, allows us to calculate a particular q, r such that
a = bq + r; and the Extended Euclidean Algorithm lets us calculate hcf(x, y) and express it
as Xx + Y y.
Some interesting things, though, are beyond our reach: theres no eective way to calcu-
late the decimal expansion of a/b where a, b R are given by their decimal expansions.
6.2 The Division Algorithm for K[X]
Given f, g K[X] with g = 0 we can nd algorithmically q, r K[X] such that f = gq + r
where r = 0 or deg r < deg g: provided we can somehow carry out the arithmetic
operations in K. There is no way to do this in general, so in this subsection we assume
that we have an Oracle which will on demand deliver to us 0, x, x+y, 1, xy, and (if y = 0)
x/y.
Provided with this Oracle all we need to do is mimic the steps of the Division Algorithm
we learned for Q[X].
It is clear what to do if deg f < deg g; just take q = 0 and r = f.
Otherwise suppose that f = aX
M
+lower degree terms and g = bX
N
+lower degree terms;
here a, b are nonzero and M N 0. Consider f(X) (a/b)X
NM
g(X); the oracle helps
us with the coecients of this polynomial. This is a polynomial of degree less than deg f we
use recursion, and nd q, r such that
f(X) (a/b)X
NM
g(X) = g(X) q(X) + r(X) where r(X) = 0 or deg r < deg g.
Now put r(X) := r(X) and q(X) := g(X) +(a/b)X
NM
g(X), where again the Oracle helps
us with the latter calculation.
It is easy to prove by induction that this process works.
6.3 Euclids Algorithm in Euclidean Domains
Let R equipped with d : R ` 0 N ` 0 be a Euclidean Domain.
Note that for a, b R this merely asserts the existence of a q, r such that a = bq + r
with r = 0 or d(r) < d(b. Weve seen above that sometimes there are algorithms to calculate
these, although we may need an Oracle or Giant Look-up Table to help. Sometimes there
34
may not be. So throughout this subsection assume that there is an Oracle which will deliver
on demand a suitable q and r.
In that case, the Extended Euclidean Algorithm we learned in the rst-year course will
calculate hcf(x, y) and express it as Xx + Y y.
Proposition 6.3.1 (Extended Euclidean Algorithm). Let R equipped with d be a Euclidean
Domain, and suppose that for any x, y = 0 we are given a denite q(x, y) and r(x, y) such
that x = yq(x, y) + r(x, y) where r(x, y) = 0 or d(r(x, y)) < d(y).
For every pair of elements a, b R, not both 0, proceed as follows:
1. (a) let a
0
:= a, m
0
:= 1, and n
0
:= 0;
(b) let a
1
:= b, m
1
:= 0, and n
1
:= 1;
(c) let i := 1;
2. while a
i
= 0 repeat the following steps:
(a) let q
i+1
:= q(a
i1
, a
i
);
(b) i. let a
i+1
:= a
i1
q
i+1
a
i
;
ii. let m
i+1
:= m
i1
q
i+1
m
i
;
iii. let n
i+1
:= n
i1
q
i+1
n
i
;
(c) increase i to i + 1;
3. (a) let d = a
i1
;
(b) let m = m
i1
;
(c) let n = n
i1
.
Then the following are true:
1. the process stops after a nite number of steps;
2. every divisor of both a and b also divides d;
3. d divides both a and b;
4. d = ma + nb.
Proof. The proof is identical to the one for integers.
6.4 Chinese Remainder Theorem
A nal remark is in order. The Extended Euclidean Algorithm is exactly what we need to
calculate the inverse of the isomorphism R/xyR

= R/xRR/yR of the Chinese Remainder


Theorem.
35
7 Factorisation in Q[X]
In this section we deal with some very concrete matters concerning the factorisation of poly-
nomials with rational coecients. Any such polynomial is a rational multiple of a polynomial
with integer coecients, and so we may concentrate on the integral associate. For such poly-
nomials we have various techniques to help us, in particular we can work modulo p for
various primes; or we can use other arithmetic tricks. We describe some of these, and we
prove the important theorem which reconciles factorisation of polynomials over Q and over
Z.
7.1 Factorisation in Z[X] and Z
p
[X]
We begin with a general lemma.
Lemma 7.1.1. Let R and S be rings, and let f : R S be a homomorphism, f : a a.
Then : R[X] S[X] given by :

N
k=0
a
k
X
k

N
k=0
a
k
X
k
is a homomorphism with
image f(R)[X] and kernel (ker f)[X].
Proof. In view of how we add and multiply polynomials all we need to check is that
a
k
+ b + k = a
k
+ b
k
and

r+s=n
a
r
b
s
=

r+s=n
a
r
b
s
.
The remarks on kernel and image are clear.
We can use what we have learned about rings and homomorphisms to give a useful
practical test for irreducibility
5
in the ring Z[X].
Proposition 7.1.2 (Eisensteins criterion). Let f(X) =

N
k=0
c
k
X
k
be a polynomial in Z[X],
and let p Z be a prime. Suppose that p[c
k
for all k = 0, 1, . . . , N 1, p [c
N
and p
2
[c
0
.
Then f(X) has no factor of smaller degree in Z[X].
Proof. Suppose that we had that f(X) = g(X)h(X), with g(X) =

N
1
k=0
a
k
X
k
and h(X) =

N
2
k=0
b
k
X
k
and N
1
, N
2
> 1. Extending the natural homomorphism Z Z
p
to the poly-
nomial rings we would then have that (g(X)(h(X)) = (f(X)) = c
N
X
N
= 0, so that
(g(X)) = aX
N
1
and h(X) = bX
N
2
by the unique factorisation in the Euclidean domain
Z
p
[X]. Hence we have that a
0
= b
0
= 0; that is p[a
0
and p[b
0
, yielding p
2
[a
0
b
0
= c
0
contrary
to hypothesis.
There are many examples of this. For instance, (X
2
+125)
2
(X
3
+25)
4
+5 has only factors
of degree 16 in Z[X] and so, being monic, is irreducible.
More famous are the cyclotomic polynomials or prime order. Let
p
(X) :=
X
p
1
X1
; then

p
has no factor of degree k with 1 < k < (p 1). We cant use Eisenstein directly; but
instead we apply the result to
p
(X + 1); the details are an exercise.
5
Well, almost irreducibility: see later.
36
7.2 Factorization in Z[X] and Q[X]
Although we often work with polynomials with integral coecients, we are really interested
in whether these factorise in the ring Q[X]. It is clear that in a trivial way a polynomial
may be a product of two irreducibles in Z[X] but of only one irreducible in Q[X]take for
example f(X) = 2X. We can easily check that in the large ring Q[X] this is irreducible; but
in the small ring it is the product of two irreducibles 2 and X. What we want to see is that
essentially nothing more complicated than this is possible.
Theorem 7.2.1. Let 0 = h Z[X]. Suppose that h = f g with f, g Q[X]. Then there
exist

f, g Z[X] with deg f = deg

f and deg g = deg g, such that h =

f g.
Proof. Write f =

M
k=0
a
k
X
k
, g =

N
k=0
b
k
X
k
and let c
k
:=

r+s=k
a
r
b
s
be the coecients
of h.
We may multiply f by the least common multiple of the denominators of its coecients,
and then divide the resulting polynomial by the highest common factor of its (integral!)
coecients. In that way we get f =

f, where Q and

f is a polynomial of the same
degree as f, and has integral coecients whose highest common factor is 1. We can nd
corresponding Q and g Z[X] with g = g.
We can nd corresponding Zthere are no denominators to clear so is integral
and

h Z[X] with h =

h.
Then we have, after re-arranging the rationals,
n

h(X) = m

f(X) g(X), for some m, n Z.
(The Xs are inserted into the polynomials for clarity.)
How are m and n related? Well n is the highest common factor of the coecients of the
polynomial on the left hand side. But the highest common factor of the coecients of the
polynomial on the right hand side is m times the highest common factor of the coecients
of

f(X) g(X). In a moment we will prove that the latter highest common factor is 1. So
m = n and we have that

h(X) =

f(X) g(X),
which we multiply by Z to get
h(X) =

h(X) =

f(X) g(X).
Dene

f :=

f and g := g and we are done.
The missing lemma is so important it deserves a separate subsection.
7.3 Gausss Lemma
We begin with some pieces of notation.
Denition 24. Let 0 = f Z[X]. Then by the content of f, denoted by c(f), we mean
the highest common factor of the coecients of f. If the content of f is 1 then we say f is
primitive.
Proposition 7.3.1 (Gausss Lemma). Let f, g Z[X] and suppose that c(f) = c(g) = 1.
Then c(f g) = 1.
37
Proof. If false let the prime p divide all the coecients of the product. Extend the natural
homomorphism Z Z
p
to a homomorphism of the polynomial rings in the standard way.
Then
0 = (f g) = (f) (g)
and so, since Z
p
[X] is an integral domain, either (f) = 0 or (g) = 0. But that implies p
divides every coecient of f or every coecient of g, a contradiction.
Corollary 7.3.2. Let f, g Z[X]; then c(f g) = c(f)c(g).
Proof. Trivial.
7.4 Z[X] is a UFD
We can use Gausss Lemma and the techniques of the previous sections to prove that in the
ring Z[X] every polynomial factorises uniquely into a product of primes in Z times a product
of primitive polynomials each irreducible in Q[X].
7.5 An algorithm for factorisation in Q[X]
How can we factorise a polynomial f Q[X]? As weve seen we may assume that f Z[X]
and carry out our factorisations over Z.
Here are a couple of trivial starting points:
Lemma 7.5.1. Suppose that n Z. If f(n) = 0 then X n[f.
Proof. Remainder Theorem and Gausss Lemma.
Lemma 7.5.2. Suppose that g[f in Z[X]; and that n Z. Then g(n)[f(n).
Proof. Trivial.
With these in mind we search for factors of f in Z[X] of degree d = 1, 2, . . . , deg f 1.
Of course whenever we nd a factor we divide it out and then look recursively at two factors
we have found.
The following lemma controls the possible values of a factor of degree d may take.
Lemma 7.5.3. Let a
0
, a
1
, a
2
, . . . , a
d
Z be distinct integers, and suppose that f(a
k
) = 0 for
each k. Then there are only nitely many (d+1)-tuples (b
0
, b
1
, . . . , b
d
) such that b
k
[f(a
k
) for
each k.
Proof. This is essentially the fact that Z is a UFD.
The following lemma allows us to reconstruct a polynomial of degree d from its values at
(d + 1) places.
Lemma 7.5.4. Let a
0
, a
1
, a
2
, . . . , a
d
Z be distinct integers, and let b
0
, b
1
, . . . , b
d
Q. There
is exactly one polynomial g Q[X] of degree at most d such that g(a
k
) = b
k
for all k.
38
Proof. This is just the Chinese Remainder Theorem. To be absolutely specic we could, as
Lagrange did, take
g(X) =
d

k=0
b
k

j=k
X a
j
a
k
a
j
.
By the previous two lemmas we can now construct a nite list of all those polynomials g
of degree d in Q[X] which can possibly be divisors of f.
be more precise if m = deg f for any d m 1 we calculate, f(0), f(1), ..., f(d + m).
Since f has at most m roots, at least d + 1, among those are = 0. Lets say that for
a
0
, a
1
, ..., a
d
0, 1, ..., d +m we have f(a
i
) = 0. Using lemmas 7.5.3, 7.5.4 we produce a
nite list of polynomials of degree d which are possible divisors of f. If g is such a polynomial
we check whether in fact g Z[X], and then, by long division, whether g[f. If g[f for some
g then we write f = gh and we repeat the procedure for g, h. This gives us an algorithm to
decompose any polynomial f Q[X] as a product of irreducibles. Of course if no g from our
nite list divides f we conclude that f is irreducible.
7.6 Factorization in R[X] and C[X]
According to the Fundamental Theorem of Algebra any f C[X] has a root in C. In
particular a polynomial f C[X] is irreducible if and only if deg f = 1.
Let f R[X] be a polynomial with deg f > 2. If f has areal root clearly it is not
irreducible. If a non-real a C is a root of f then a is also a root. So (Xa)(X a) divides
f. But (X a)(X a) = X
2
2aX + [a[
2
R[X]. So again f is not irreducible. Hence
the irreducible polynomials in R[X] are the polynomials of degree 1 and the polynomials of
degree 2 with no real roots.
39
Which of the following are true?
1. If the polynomial f(x) Z[X] is irreducible in Z[X] then f(X) is also irreducible in
Z
3
[X].
2. If the polynomial f(x) Z
7
[X] is irreducible in Z
7
[X] then f(X) is also irreducible in
Z[X].
3. If K is a eld, a polynomial of degree 3, f(X) K[X] is irreducible if and only if it has
no roots in K.
4. There is a monic polynomial f(X) Z[X] such that f(
2008
2009
) = 0.
40

You might also like