M3P8
M3P8
Modules
Rings
Integral domains
Euclidean domains
Fields
Dedekind domains
Noetherian domains
Noetherian rings
Noetherian modules
Syllabus
Rings. Homomorphisms, ideals, and quotients. Factorisation. The Chinese remainder theorem. Fields
and field extensions. Finite fields. R-modules. Noetherian rings and modules. Polynomial rings in
several variables. Integral extensions and algebraic integers. Dedekind domains. Integers in number
fields. Introduction to algebraic geometry.
1
M3P8 Algebra III Contents
Contents
0 Introduction 4
3 Factorisation 14
3.1 Divisibility, units, associates, and irreducibles . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2 Unique factorisation domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.3 Principal ideal domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.4 Euclidean domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6 Finite fields 25
6.1 Finite fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6.2 The Frobenius automorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6.3 Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
6.4 The multiplicative group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
6.5 Uniqueness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
7 R-modules 29
7.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
7.2 Submodules, quotients, and direct sums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
7.3 Module homomorphisms, kernels, and images . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
7.4 Free modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
7.5 Generators and relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2
M3P8 Algebra III Contents
11 Dedekind domains 49
11.1 Dedekind domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
11.2 Ideal class groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3
M3P8 Algebra III 0 Introduction
0 Introduction
Lecture 1
This course is an introduction to ring theory. The topics covered will include ideals, factorisation, the theory Friday
of field extensions, finite fields, polynomial rings in several variables, and the theory of modules. In addition 05/10/18
to the lecture notes, the following will cover much of the material we will be studying.
• M Artin, Algebra, 1991
Rings are contexts in which it makes sense to add and multiply. For example, Z, Q, R, C, polynomials,
functions {0, 1} → R, and Z/nZ are rings. The goals of this course include
• to unify arguments that apply in all of the above contexts, and
• to study relationships between different rings.
4
M3P8 Algebra III 1 Basic definitions and examples
Definition 1.1.1. A commutative ring with identity R is a set together with two binary operations
+R : R × R → R, ·R : R × R → R,
0R , 1R ,
– for all r ∈ R,
0R +R r = r +R 0R = 0R ,
– for all r, s, t ∈ R,
(r +R s) +R t = r +R (s +R t) ,
– for all r, s ∈ R,
r +R s = s +R r,
– for all r ∈ R, there exists −r ∈ R such that
r +R (−r) = (−r) +R r = 0R .
5
M3P8 Algebra III 1 Basic definitions and examples
rr−1 = r−1 r = 1R .
We do not consider the zero ring {0R } to be a field. We have seen many examples of rings at this point.
Example.
• The sets Z, Q, R, C are all rings with their usual notion of addition and multiplication. All of them but
Z are in fact fields.
• We have the ring Z/nZ of integers mod n. Let n ∈ Z>0 , and recall that a and b are said to be
congruent mod n if a − b is divisible by n. It is easy to check that this is an equivalence relation on
Z. Moreover, since any a ∈ Z can uniquely be written as qn + r with q, r ∈ Z and 0 ≤ r < n, the set
{[0]n , . . . , [n − 1]n }
is a complete list of the equivalence classes under this relation, where [a]n denotes the set of all integers
congruent to a mod n. We denote this n-element set by Z/nZ, and we can define addition and
multiplication in Z/nZ by setting
This defines a ring structure on Z/nZ, once one checks that it is well-defined. This is the first example
of a general construction we will see more of later, the quotient of a ring by an ideal.
6
M3P8 Algebra III 1 Basic definitions and examples
r0 + · · · + rn X n , n ∈ Z≥0 , r0 , . . . , rn ∈ R.
in which all but finitely many ri are zero. This makes it easier to define addition and multiplication. The
degree of such an expression is the largest i such that ri is nonzero. We add and multiply in R [X] just as
we would any other polynomials.
∞ ∞ ∞
! !
X X X
i i
ri X +R[X] si X = (ri +R si ) X i ,
i=0 i=0 i=0
∞ ∞ ∞
! ! i
X X X X
ri X i ·R[X] si X i = (rj ·R si−j ) X i .
i=0 i=0 i=0 j=0
What about polynomial rings in more than one variable? Since the construction of polynomial rings takes
an arbitrary ring as input, one can iterate it. Start with a ring R, and consider first the ring R [X] and then
the ring (R [X]) [Y ]. A polynomial of this has the form
X∞ X∞
rij X j Y i , rij ∈ R.
i=0 j=0
On the other hand, we can consider the ring (R [Y ]) [X], whose polynomials have the form
X∞ ∞
X
rij Y j X i , rij ∈ R.
i=0 j=0
Alternatively, we could consider the ring R [X, Y ] whose polynomials are formal expressions of the form
∞
X
rij X i Y j , rij ∈ R,
i,j=0
with only finitely many nonzero coefficients rij and define addition and multiplication in the usual way. It
is not hard to see that all three approaches yield the same ring. If we identify the elements
X∞ ∞
X X∞ X∞ X∞
rij X j Y i ∈ (R [X]) [Y ] , rij Y j X i ∈ (R [Y ]) [X] , rij X i Y j ∈ R [X, Y ] ,
i=0 j=0 i=0 j=0 i,j=0
we see that addition and multiplication in any of these three rings gives the same answer. We will therefore
primarily use notation like R [X, Y ] for polynomial rings in multiple variables, but we will occasionally need
to know that this is the same as (R [X]) [Y ] or (R [Y ]) [X]. The identifications we have made here are an
example of isomorphisms of rings, a notion we will make precise later.
7
M3P8 Algebra III 1 Basic definitions and examples
8
M3P8 Algebra III 1 Basic definitions and examples
a a0 ab0 + ba0
+ 0
= ,
b b bb0
a a0 aa0
· = .
b b0 bb0
Then K (R) is a field, and it contains R in a natural way as a subring if we identify r with r/1R ∈ K (R).
0R 1R
0K(R) = , 1K(R) = .
1R 1R
If a 6= 0 in R, then b/a ∈ K (R), so
a b ab 1
· = ∼ .
b a ba 1
The field K (R) is in some sense the smallest field containing R as a subring. When we talk about homomorph-
isms and isomorphisms, we will be able to state this more precisely. More generally, let the multiplicative
system S be a subset of R that contains 1R , does not contain 0R and is closed under multiplication. That
is, if a, b are in S then so is ab. For any integral domain R and any multiplicative system S, we can define
S −1 R to be the subring of K (R) consisting of all fractions of the form a/b with b ∈ S. It is easy to see that
this is closed under addition and multiplication, and defines a ring in between R and K (R).
Example. If R = Z and S is the set of powers of 2, then S −1 R = Z [1/2]. On the other hand, if S is the
set of odd integers, then S −1 R is the set of all rational numbers of the form a/b with b odd.
In general S −1 R is the smallest subring of K (R) containing R in which every element of S has a multiplicative
inverse, that is 1/b ∈ S for all b ∈ S. The process of obtaining S −1 R from R is called localisation and is
an extremely powerful tool. One can even make sense of it when R is not an integral domain, but one has
to be more careful. The equivalence relation on fractions is trickier, for example. We will not discuss this in
this course but it will be quite useful in future courses.
9
M3P8 Algebra III 2 Homomorphisms, ideals, and quotients
3. for all r, r0 ∈ R,
f (r ·R r0 ) = f (r) ·S f (r0 ) .
Adding the additive inverse, in S, of f (0R ) to both sides gives 0S = f (0R ). Thus we do not need to require
this as an axiom. On the other hand we do need to require f (1R ) = 1S . For certain R, S one can construct
examples of maps f : R → S that satisfy properties 2 and 3 of the definition without satisfying property 1.
Definition 2.1.2. A bijective homomorphism f : R → S is called an isomorphism. Write S ∼ = R for S
is isomorphic to R. In this case one verifies easily that the inverse map f −1 : S → R is also a bijective
homomorphism.
Example.
• If R is a subring of S, then the inclusion of R into S is a homomorphism. This is just a fancy way of
saying that the addition and multiplication on R are induced from the corresponding operations on S.
In particular the inclusions Z ⊂ Q ⊂ R ⊂ C are all homomorphisms.
• The composition of two homomorphisms is a homomorphism, as is easily checked from the definitions.
• The map Z → Z/nZ that takes m ∈ Z into its congruence class mod n is a ring homomorphism.
In fact, this is a special case of the following construction.
Proposition 2.1.3. Let R be any ring. Then there is a unique ring homomorphism f : Z → R such that
1R + · · · + 1R
n>0
f (n) = 0R n=0.
− (1R + · · · + 1R ) n<0
Proof. Let f : Z → R be a homomorphism. Then, directly from the definition, we have f (0) = 0R and
f (1) = 1R . In particular for all n > 0,
f (n) = f (1 + · · · + 1) = 1R + · · · + 1R ,
we find that f (−n) is the additive inverse of − (1R + · · · + 1R ). Thus f (n) is determined, for all n, completely
by the fact that f is a homomorphism. In the converse direction, it is not hard to check that the map defined
above is in fact a homomorphism.
Thus, for any ring R, we can regard an integer as an element of R via this homomorphism.
10
M3P8 Algebra III 2 Homomorphisms, ideals, and quotients
P (s) = r0 + · · · + rn sn ∈ R.
φs,f : R [X] −→ S
.
r0 + · · · + rn X n 7−→ f (r0 ) + · · · + f (rn ) sn
That is, by applying f to the coefficients and substituting s for X. Again, this is clearly a homomorphism.
The evaluation homomorphisms φs,f are a fundamental property of polynomial rings. In some sense, they
are the reason polynomial rings are worth studying. In fact, the ring R [X] is uniquely characterised by
the fact that homomorphisms from R [X] to S are in bijection with pairs (s, f ), where f : R → S is a
homomorphism and s is an element of S.
Im (f ) = {f (r) | r ∈ R} ⊆ S.
The kernel of f is
Ker (f ) = {r ∈ R | f (r) = 0} ⊆ R.
11
M3P8 Algebra III 2 Homomorphisms, ideals, and quotients
so s ∈ I for all s ∈ R. Conversely, if R has only the zero ideal and the unit ideal, let r ∈ R 6= 0, and let
I = {sr | s ∈ R}. This is an ideal that is not the zero ideal, so it is all of R. In particular, 1 ∈ I, so there
exists s ∈ R such that sr = 1.
More generally is the following.
Definition 2.4.2. If S is a subset of elements of R, then any ideal containing S consists of all elements of
R of the form
r0 s0 + · · · + rn sn , n ∈ Z≥0 , ri ∈ R, si ∈ S.
The set of all elements of this form is an ideal of R, known as the ideal of R generated by S, and denoted
hSi. It is the intersection of all the ideals of R containing S. It is also the smallest ideal of R containing S.
If S has one element, hSi is a principal ideal. We will show soon that any ideal of Z is a principal ideal,
as is any ideal of the ring K [X] for any field K. You may well have seen this in last year’s algebra course.
On the other hand, there are rings in which not every ideal is principal. For example, the ideal hX, Y i of
K [X, Y ] is not a principal ideal. Given ideals I and J there are several ways to create new ideals.
Example.
• If I, J are ideals, then the intersection I ∩ J is an ideal. Note that if I and J are given by generators,
it might be hard to find generators for the intersection. Certainly it is not enough to intersect the
generating sets.
• The union of ideals is not usually an ideal. Taking R = Z, h3i ∪ h5i contains 3, 5 but not 3 + 5.
• If I, J are ideals, then the sum I + J is an ideal, which are all expressions of the form i + j for i ∈ I and
j ∈ J. It is the smallest ideal containing both I and J, or equivalently the ideal generated by I ∪ J.
• If I, J are ideals, the product I · J or IJ is the ideal generated by elements of the form ij with i ∈ I
and j ∈ J. This may be strictly larger than the set of such products. For example, consider the
product of the ideals I = hX, Y i and J = hZ, W i in R = K [X, Y, Z, W ] for K a field. The product
IJ = hXZ, XW, Y Z, Y W i contains XZ + Y W , but the latter is not a product of an element in I with
an element in J.
• If I, J are general ideals, the product of ideals I and J is always contained in the intersection of I and
J, but the two need not be equal, even in simple rings like Z. h3i · h3i = h9i ⊆ Z and h3i ∩ h3i = h3i.
2.5 Quotients
Let R be a ring and let I be an ideal of R. If x, y are elements of R, we say that x is congruent to y mod I
if x − y is in I. This is an equivalence relation on R. We denote the equivalence class of r by r + I, or as the
alternative notations [r]I , r. It is the set {r + s | s ∈ I}. Let R/I denote the set of equivalence classes on R
mod I. This set has the natural structure of a ring. The additive and multiplicative identities are 0R + I
and 1R + I, respectively, and addition and multiplication are defined by
(r + I) + (s + I) = (r + s) + I,
(r + I) · (s + I) = (rs) + I.
respectively. One has to check that these are well-defined, but this is not difficult. The ring R/I is called
the quotient of R by the ideal I.
12
M3P8 Algebra III 2 Homomorphisms, ideals, and quotients
Example. If R = Z and I is the ideal generated by n, then R/I is the ring Z/nZ that we have already seen.
There is a natural quotient homomorphism, reduction mod I,
R
R −→
I .
r 7−→ r+I
13
M3P8 Algebra III 3 Factorisation
3 Factorisation
In these notes R always denotes an integral domain.
14
M3P8 Algebra III 3 Factorisation
We will show later that a very mild finiteness condition on a domain R, the condition that R is Noetherian,
actually guarantees that 1 holds. Another way to interpret condition 2 is as follows.
Definition 3.2.2. We say an element r of R is prime if the principal ideal hri of R is a prime ideal. In
other words, for any s, s0 in R, if r divides ss0 , then r | s or r | s0 .
Proposition 3.2.4. Let R be a domain in which condition 1 holds. Then condition 2 above holds for R if
and only if every irreducible element of R is prime.
Proof. First suppose condition 2 holds, and let r be an irreducible element of R. If r divides ab, we can
write rs = ab for some s ∈ R. Expanding out s, a, b as products of irreducibles we see that r is an associate
of some irreducible dividing a or b, so r is prime. Conversely, if every irreducible element of R is prime, and
we have products of irreducibles
p1 . . . pn = q1 . . . qm ,
then, since p1 is prime, it divides the product q1 . . . qm and is thus an associate of some qi . We can thus
cancel p1 from the left and qi from the right, after introducing a unit on one side. This is possible because
R is an integral domain. Repeating the process we find that, up to reordering the terms and multiplying by
units, the two expressions coincide.
hr0 i ( hr1 i ( . . . .
Let I be the union of all these ideals generated by r0 , r1 , . . . . Then I is an ideal, so it is generated by some
element s ∈ I. Thus s divides ri for all i. On the other hand, s lives in some hrj i, so rj divides s. Thus
s is an associate of rj , and therefore an associate of ri for all i > j, that is I ⊆ hrj i. This contradicts our
construction, because hrj+1 i ⊆ I and hrj+1 i =6 hrj i.
Thus r has an irreducible divisor s0 .
15
M3P8 Algebra III 3 Factorisation
Lemma 3.3.4. Let R be a PID. Every nonzero nonunit r ∈ R is a finite product of irreducibles.
Proof. Consider rs−1 −1
0 . If this is a unit we are done. If not let s1 be an irreducible divisor of rs0 . If
−1
r (s0 s1 ) is a unit we are done. Otherwise repeat. We obtain a sequence of irreducibles s0 , s1 , . . . such that
s0 . . . si divides r for all i, so
r = r0 s0 = r0 r1 s1 = . . . ,
with r0 , r1 , . . . irreducible. If this process ever terminates we are done. Suppose it does not. Then we have
a strictly increasing tower of ideals
hri ( hs0 i ( hs1 i ( . . . .
This cannot continue forever. Arguing as above we arrive at a contradiction.
Now we show 2.
Proof of Theorem 3.3.2. It suffices to show that in a PID every irreducible is prime. Let r ∈ R be irreducible,
and suppose that r divides st. Want r | s or r | t. Let q be a generator of the ideal hr, si of R, so hr, si = hqi.
Then q divides r, so either q is a unit or q is an associate of r. If q is an associate of r, then since q divides
s, r divides s. On the other hand, if q is a unit, then the ideal generated by r and s is the unit ideal and
1 ∈ hr, si, so we can write 1 = xr + ys for x, y elements of R. We then have t = xrt + yst, and since r divides
both yst and xrt, r divides t.
Lecture 6
Wednesday
3.4 Euclidean domains 16/10/18
One technique for proving that rings are PIDs is Euclid’s algorithm. We formalise this in an abstract setting
as follows.
Definition 3.4.1. Let R be an integral domain.
• A Euclidean norm on R is a function N : R → Z≥0 such that for all a, b ∈ R, with b 6= 0, there exists
q, r ∈ R such that a = qb + r, and either r = 0 or N (r) < N (b).
16
M3P8 Algebra III 3 Factorisation
3.5 Examples
Example.
• The classic example of a Euclidean domain is Z, with
N (x) = |x| , qx ∈ Z.
2
• The ring Z [i] is a Euclidean domain, with N (z) = zz = |z| , so
2
N (x + yi) = |x + yi| = x2 + y 2 .
To see this, note that given a and b in Z [i] for b 6= 0, set q 0 = a/b ∈ Q [i]. Write
q 0 = x0 + iy 0 , x0 , y 0 ∈ Q.
• Similar arguments can be used to prove that Z [α] is a Euclidean domain for
√ √ √
α = −2, α = −1+2 −3 , α = −1+2 −7 .
Beyond this one needs other tricks, and for most α unique factorisation fails.
• A critical example is the polynomial ring K [X] for K a field. Here we can take N (P (X)) to be the
degree of P (X). Then, given polynomials P (X) , T (X) ∈ K [X] and T (X) 6= 0, we can use polynomial
long division to write
P (X) = Q (X) T (X) + R (X) ,
for some Q (X) with the degree of R (X) strictly less than that of T (X), unless T (X) is constant, in
which case we can make R (X) = 0. To prove this, fix T (X). If deg (T (X)) = 0, T (X) is constant, so
T (X) = c 6= 0 ∈ K. Take
1
Q (X) = P (X) , R (X) = 0.
c
Otherwise induct on deg (P (X)). If deg (P (X)) < deg (T (X)), set
Suppose the claim is true for polynomials of degree n and P (X) has degree n + 1, so
n+1
X d
X
P (X) = ai X i , T (X) = bi X i , d < n + 1.
i=0 i=0
Then
an+1 n+1−d
S (X) = P (X) − X T (X)
bd
has degree n. By inductive hypothesis there exist Q (X) , R (X) with deg (R (X)) < deg (T (X)) such
that S (X) = Q (X) T (X) + R (X), so
an+1 n+1−d
P (X) = X + Q (X) T (X) + R (X) .
bd
17
M3P8 Algebra III 4 The Chinese remainder theorem
a ∈ a1 + I1 , ..., a ∈ ar + Ir ?
4.1 Products
Definition 4.1.1. Let R1 , . . . , Rn be rings. The direct product R1 × · · · × Rn is a ring whose elements
are n-tuples (r1 , . . . , rn ) with ri ∈ Ri for all i. The addition and multiplication are given componentwise.
Note. The product comes with natural homomorphisms πi for all i, the projection onto the i-th factor,
defined by
πi : R1 × · · · × Rn −→ Ri
.
(r1 , . . . , rn ) 7−→ ri
The product also comes with the following universal property.
Theorem 4.1.2 (Universal property of the product). Let S, R1 , . . . , Rn be any rings. For any homomorph-
isms
f1 : S → R1 , ..., f n : S → Rn ,
there exists a unique homomorphism
f : S → R1 × · · · × Rn ,
such that πi ◦ f = f for all i.
Proof. Given fi , the homomorphism f is defined by
Then (πi ◦ f ) (s) = fi (s). For uniqueness, if (π ◦ g) (s) = fi (s) for all i, then
Q
More generally, if I is any index set, and for each i ∈ I we have a ring Ri , we can define the product i Ri .
An element r of this product is a choice, for each i ∈ I, of an element of Ri . We write such an element as
(ri )i∈I . For each j ∈ I we have a map
Q
πj : i Ri −→ Rj
.
(ri )i∈I 7−→ rj
Such a product satisfies a very similar universal property. For any collection
fi : S → Ri
such that πj ◦ f = fj .
18
M3P8 Algebra III 4 The Chinese remainder theorem
and ej will be congruent to 1 mod Ij and zero mod Ii for all j 6= i, so ej has one only in the j-th place. So
R R
R→ × ··· ×
I1 Ir
is surjective. The result follows.
19
M3P8 Algebra III 4 The Chinese remainder theorem
4.3 Examples
When R = Z, then every ideal is principal, so we can write Ij = hnj i for all j. The condition that Ii + Ij
is the unit ideal becomes the condition that ni ∈ Z are pairwise relatively prime. In this case the ideal J is
generated by the product n of the ni . Specialising, we find the version of the Chinese remainder theorem
from elementary number theory.
Theorem 4.3.1. If {nj ∈ Z} is a finite collection of pairwise relatively prime integers, and n is their product,
then for any c1 , . . . , cr ∈ Z, there exists c ∈ Z unique up to congruence mod n such that c is congruent to ci
mod ni for all i.
Now let K be a field and take R = K [X]. If c1 , . . . , cr ∈ K are distinct elements of K, the ideals
Ii = hX − ci i ⊆ R
fi : R −→ K
.
P (X) 7−→ P (ci )
Let
f : R −→ K × ··· × K
.
P (X) 7−→ (P (c1 ) , . . . , P (cr ))
Then the following diagram commutes.
f
R K × ··· × K
∼
.
R R R
∼ × ··· ×
J I1 Ir
Theorem 4.3.2. For any c1 , . . . , cn ∈ K, there is a polynomial P (X) in K [X], unique up to congruence
mod
(X − a1 ) . . . (X − an ) ,
such that P (ai ) = ci for all i.
20
M3P8 Algebra III 5 Fields and field extensions
ι : Z −→ K
, n ≥ 0.
n 7−→ nK = 1K + · · · + 1K
Let I be the kernel. Then Z/I ,→ K so Z/I is an integral domain, so I is a prime ideal. Thus I is either
the zero ideal {0}, if K has characteristic zero, or the ideal hpi for some prime p of Z. In the former case
I = {0}, the injection Z ,→ K extends to an inclusion
Q −→ K
a aK .
7−→ (ιa) ιb−1 =
b bK
In the latter case I = hpi, we get an injection of the field Fp ,
Z/pZ ,→ K,
which we often denote Fp when we think of it as a field. Upshot is that every field K contains exactly one
of Q or Fp for p prime, in exactly one way depending on its characteristic. This field is called the prime
field of K, and it is contained in K in a unique way.
Note. Such an inclusion of fields L/K makes L into a K-vector space, that is a vector space over K.
Definition 5.2.2. We say that a field extension L/K is finite if L is finite dimensional as a K-vector space.
If this is the case, the degree of such an extension is the dimension of L as a K-vector space dimK (L), and
is denoted [L : K].
Proposition 5.2.3. Let K ⊆ L ⊆ M be fields. Then M/K is finite if and only if M/L and L/K are both
finite. If this is the case then
[M : K] = [M : L] [L : K] .
Proof. First suppose that M/K is finite. Then L is a K-subspace of M , so finite dimensional as a K-vector
space. Moreover, there exists a K-basis m1 , . . . , mr , and this basis spans M over K and thus also over L.
Thus M is finite dimensional as an L-vector space, so M/L is finite. Conversely, suppose L/K and M/L are
finite. Let e1 , . . . , en be a K-basis for L, and let f1 , . . . , fn be an L-basis for M . Then claim that
e1 f1 , . . . , e1 fm , . . . , en f1 , . . . , en fm
c1 f1 + · · · + cm fm , ci ∈ L.
21
M3P8 Algebra III 5 Fields and field extensions
is an L-linear combination of the fj that is zero. Since the fj are linearly independent over L we must have
X
di,j ei = 0,
i
for all j. Since the ei are linearly independent over K we must have di,j = 0 for all i, j.
Lecture 8
Monday
5.3 Extensions generated by one element 22/10/18
Let L/K be a field extension, and let α be an element of L.
Definition 5.3.1. We let K (α) denote the subfield of L consisting of all elements of L that can be expressed
in the form
P (α)
,
Q (α)
where P and Q are polynomials with coefficients in K and Q (α) is not zero. This is the smallest subfield of
L containing K and α.
Recall that if R, S are rings, f : R → S is a homomorphism, and α ∈ S, then have
φf,a : Pn R [X]i −→ S
n .
f (ri ) αi
P
i=1 ri X 7−→ i=1
22
M3P8 Algebra III 5 Fields and field extensions
So in this case there is no nonzero polynomial P ∈ K [X] with P (α) = 0. In this case the map taking P (X)
to P (α) is an injection of K [X] into K (α) ⊆ L. In particular every nonzero element of K [X] gets sent to
a nonzero, hence invertible, element of L. Thus the map from K [X] to L extends to an injective map from
the field of fractions of K [X],
K (X) −→ L
P (X) P (α) .
7−→
Q (X) Q (α)
By definition of K (α), this map is surjective so the image of this map is K (α). In particular K (X) and
K (α) are isomorphic.
Note. In this case K (α) is infinite dimensional as a K-vector space. It contains a subspace isomorphic to
K [X], for instance.
If α is algebraic over K, then I is a nonzero maximal ideal of the PID K [X], so it is generated by a
single irreducible polynomial Q (X) in K [X]. As a consequence, since the units in K [X] are just the
constant polynomials, the polynomial Q (X) is well-defined up to a constant factor. It is called the minimal
polynomial of α. By definition, it divides every polynomial P (X) such that P (α) = 0. Since hQ (X)i is
maximal, the ring K [X] / hQ (X)i is a field. Recall that for any P ∈ K [X], can write P (X) uniquely as
So
1, . . . , X deg(Q)−1
are a K-basis of K [X] / hQ (X)i. So its dimension as a K-vector space is equal to the degree of Q (X). The
map K [X] → K (α) ⊆ L descends to an injection of K [X] / hQ (X)i into L. Since its image is a subfield of
K (α) containing K and α, this map is an isomorphism
K [X]
K (α) ∼
= .
hQ (X)i
Thus in this case the extension K (α) /K is a finite extension, of degree equal to the degree of Q (X). To
summarise, extend K by a single element by
• building K [X], and
• either passing to field of fractions K (X) to form a transcendental extension, or choosing an irreducible
polynomial Q to form an algebraic extension K [X] / hQ (X)i.
Slightly informally, instead of K [X] / hQ (X)i, we sometimes write K (α), where α is a root of Q (X).
23
M3P8 Algebra III 5 Fields and field extensions
Corollary 5.4.3. Let L/K be a field extension, and suppose α, β are elements of L algebraic over K. Then
α + β and αβ are algebraic over K. Moreover, if α is nonzero then α−1 is algebraic over K.
Proof. Consider the chain of extensions
K ⊆ K (α) ⊆ K (α, β) ,
where we write K (α, β) for (K (α)) (β). Since α is algebraic over K, K (α) is finite over K, of degree deg (α).
Since β is algebraic over K, it is also algebraic over K (α), so K (α, β) is finite over K (α), of degree at most
deg (β). Thus K (α, β) is algebraic over K, of degree at most (deg (α)) (deg (β)). On the other hand, we also
have a chain of extensions
K ⊆ K (α + β) ⊆ K (α, β) ,
so K (α + β) is finite over K, of degree at most (deg (α)) (deg (β)). Hence α + β is algebraic over K. The
proofs for αβ and α−1 are similar.
Corollary 5.4.4. For any extension L/K, let Lalg be the subset of L consisting of all elements of L that
are algebraic over K. Then Lalg is a field.
Proof. We have seen that Lalg is closed
n under addition, multiplication, and taking inverses. For example, if
a0 + · · · + an αn = 0, then a0 α−1 + · · · + an = 0.
Example. In particular, the subfield Q ⊆ C of complex numbers that are algebraic over Q is a field, called
the field of algebraic numbers.
5.5 Example
Example. Consider the polynomial X 2 + X + 1 in F2 [X]. It has no roots in F2 , so it is irreducible, as a
polynomial of degree two any nontrivial factor would be linear. The other polynomials of degree two are
2
X 2, X 2 + X = X (X + 1) , X 2 + 1 = (X + 1) ,
so X 2 + X + 1 is the unique irreducible polynomial of degree two. Thus the quotient F2 [X] / X 2 + X + 1
is a field extension of degree two of F2 , which is denoted F4 . Its four elements are 0, 1, X, X + 1, or more
precisely, their classes mod X 2 + X + 1 .
· 0 1 X X +1
0 0 0 0 0
1 0 1 X X +1 .
X 0 X X +1 1
X +1 0 X +1 1 X
Note that
2
X 2 = −X − 1 = X + 1, X 2 + X + 1 = 0, (X + 1) = X, X 3 = X (X + 1) = 1
in F4 . In particular the multiplicative group of F4 is cyclic of order three. This is not particularly surprising,
as all groups of order three are cyclic. We will see later, though, that the multiplicative group of any finite
field is cyclic.
Proposition 5.5.1. Let K be a field with four elements. Then K ∼
= F4 .
Proof. Let α ∈ K with α 6= 0 and α 6= 1. Consider 1, α, α2 . Since K has dimension two over F2 , there is a
linear dependence. So there exists a polynomial P in F2 [X] of degree at most two such that P (α) = 0. In
fact P must be irreducible of degree two. If it is divisible by something of degree one, then a polynomial of
degree one vanishes on α, so α = 0 or α = 1. So α2 + α + 1 = 0. The map
F2 [X] −→ K
.
X 7−→ α
24
M3P8 Algebra III 6 Finite fields
6 Finite fields
Lecture 9
6.1 Finite fields Wednesday
24/10/18
Let K be a finite field. That is, a field with only finitely many elements. Then K has characteristic p for
some prime p, and is in particular a finite dimensional Fp -vector space. Thus its order is a power pr of p
for r ∈ Z>0 . If we fix a particular prime power pr , then two questions naturally arise. Does there exist a
field of order pr ? If so, can we classify fields of order pr up to isomorphism? We will see that in fact, up to
isomorphism, there is a unique field Fpr of order pr .
0 7→ 0, 1 7→ 1, X 7→ X + 1, X + 1 7→ X.
r
Let K be a field of pr elements. Then αp = α for all α ∈ K. If α = 0, clear. Otherwise α ∈ K ∗ , so K ∗
r r
is an abelian group of order pr − 1. Lagrange’s theorem implies that αp −1 = 1, so αp = α. We have the
following.
r
Proposition 6.2.1. Let K be a field of characteristic p, such that αp = α for all α ∈ K. Let P (X) ∈ K [X]
r r
be an irreducible factor of X p − X over K [X]. Then every element β of K [X] / hP (X)i satisfies β p = β.
Proof. Let d = deg (P ). Can write
β = c0 + · · · + cd−1 X d−1 .
r
Moreover, since P (X) = 0 in K [X] / hP (X)i and P (X) divides X p − X, we have
r
Xp = X
25
M3P8 Algebra III 6 Finite fields
K0 = Fp ( K1 ( K2 ( . . .
r
all satisfying 1 as follows. Suppose we have constructed Ki satisfying 1. If X p − X factors into linear factors
r
over Ki [X], we are done. Otherwise, choose a nonlinear irreducible factor Pi (X) of X p − X in Ki [X] of
degree at least two, and set
Ki [X]
Ki+1 = .
hPi (X)i
Then Ki+1 is strictly larger than Ki and still satisfies 1. On the other hand, in any field Ki satisfying 1,
r
every element is a root of X p − X, so #Ki ≤ pr for all i. Since this polynomial can have at most pr roots,
this process must eventually terminate.
r
Since X p − X has degree pr , we expect the field K constructed above to have pr elements. So it suffices to
r
show that over any field K of characteristic p, X p − X has no repeated roots. To prove this we need an
additional tool.
6.3 Derivatives
Definition 6.3.1. Let R be a ring, and let
P (X) = r0 + · · · + rd X d
r1 + · · · + drd X d−1 .
Note. Just as for differentiation in calculus, we have a Leibniz rule. For P, Q ∈ R [X],
0
(P Q) (X) = P (X) Q0 (X) + P 0 (X) Q (X) ,
by reducing to P, Q monomials.
From this we deduce the following.
Lemma 6.3.2. Let K be a field, and let P (X) be a polynomial in K [X] with a multiple root in K. Then
P (X) and P 0 (X) have a common factor of degree greater than zero.
Proof. Let α ∈ K be the multiple root. Then we can write
2
P (X) = (X − α) Q (X) .
26
M3P8 Algebra III 6 Finite fields
Proof. Since every element of Z/nZ has order d for some d dividing n, the sum over all possible d dividing
n of the number of elements of order d is just the number of elements of Z/nZ, which is n.
Proof of Proposition 6.4.1. Let A be as in Proposition 6.4.1. We must show that A contains an element of
order n. In fact, we will show, by induction on d, that A contains exactly Φ (d) elements of order d for all
d | n. In particular, A has Φ (n) > 0 elements of order n, so it is cyclic. If d = 1, the only element of order
one is the identity of A. Since Φ (1) = 1 the base case holds. Assume the claim is true for all d0 < d. A has d
elements of order dividing d, and Φ (d) elements of order d0 for d0 | d and d0 < d, so the number of elements
of exact order d is X
d− Φ (d0 ) .
d0 |d, d0 <d
27
M3P8 Algebra III 6 Finite fields
6.5 Uniqueness
We now turn to the question of showing that any two fields of pr elements are isomorphic. Let K be such a
field. The cyclicity of K ∗ immediately shows the following.
Proposition 6.5.1. Any finite field K of characteristic p is generated over Fp by a single element α ∈ K.
Proof. Let α be an element of K, that generates K ∗ as an abelian group. Then Fp (α) is contained in K,
but contains αn for all n, so contains K ∗ , hence K = Fp (α).
As a corollary, we deduce the following.
Proposition 6.5.2. For any prime p and any r ∈ Z>0 , there exists an irreducible polynomial P (X) ∈ Fp [X]
of degree r in Fp [X].
Proof. Let K be a finite field of pr elements, let α be an element of K that generates K over Fp , and let P
the minimal polynomial of α over Fp . We then have a surjective map
Fp [X] −→ K
.
X 7−→ α
Its kernel is generated by the irreducible polynomial P (X) of degree deg (P ) = [Fp (α) : Fp ] = r.
We also have the following trick.
r
−1
Lemma 6.5.3. Every irreducible polynomial P (X) of degree r in Fp [X] is a divisor of X p − 1.
r r
Proof. Let K = Fp (α) where α is a root of P . #K = pr so αp − α is zero in K. So P (X) | X p − X.
Fp [X]
K = Fp (α) ∼
= ,
hP (X)i
where P (X) is the minimal polynomial of α over Fp . In particular P (X) is irreducible of degree r. Since
r r
P (X) divides X p −1 − 1 in Fp [X], it also divides X p −1 − 1 in K 0 [X]. Since in K 0 [X], the latter factors
into linear factors, P (X) also factors into linear factors over K 0 . In particular there exists a root α0 ∈ K 0 of
P (X) in K 0 [X] such that P (α0 ) = 0. Then the map
Fp [X] −→ K0
X 7−→ α0
Fp [X]
K → → K0
hP (X)i .
Q (α) 7→ Q (X) 7→ Q (α0 )
Since this is map of fields from K to K 0 that takes α to α0 it is injective. Since both fields K, K 0 have the
same cardinality pr , it is also surjective and an isomorphism.
If K = Q, Q [X] / X 2 − p are pairwise nonisomorphic extensions of degree α for every prime p. Lecture 11
Lecture 11 is a problem class. Monday
29/10/18
28
M3P8 Algebra III 7 R-modules
7 R-modules
Lecture 12
7.1 Definitions Wednesday
31/10/18
Definition 7.1.1. An R-module M is a set, together with two operations
+M : M × M → M, ·M : R × M → M,
such that
1. (M, +) makes M into an abelian group with identity 0M ,
5. for all m ∈ M ,
1R · m = m.
Note. For an abelian group M , let End (M ) denote the set of homomorphisms M → M of abelian groups.
End (M ) is a noncommutative ring. 2 if and only if for all r ∈ R, ·r : M → M lives in End (M ). 3, 4, 5 if
and only if the map R → End (M ) given by 2 is a homomorphism of rings.
Example.
• The usual addition and multiplication on R naturally makes R into an R-module. More generally, any
ideal I of R is an R-module with the usual addition and multiplication.
• If f : R → S, then f makes S into an R-module, where the addition + is the usual addition in S, and
the multiplication law is defined by
r · s = f (r) ·S s, r ∈ S, s ∈ S.
(r + I) · m = rm.
If r + I = r0 + I, then r − r0 ∈ I, so
rm − r0 m = (r − r0 ) m = 0,
by assumption.
29
M3P8 Algebra III 7 R-modules
• Let R = Z, and let M be an abelian group. Then M has the unique natural structure of Z-module, as
follows. Property 3 from the module axioms shows that
m + · · · + m
n>0
n·m= 0 n=0.
(−m) + · · · + (−m) n < 0
Thus the multiplication law Z × M → M is forced on us, and one checks that it does satisfy properties
2 to 5 above. Informally, we say that abelian groups are Z-modules.
• If R is a field, then R-modules are just R-vector spaces.
• Let S be a set, and let MS be the set of R-valued functions f : S → R. We add and multiply pointwise.
For f, g ∈ MS , we can define
f + g = s 7→ f (s) + g (s) ,
rf = s 7→ r · f (s) .
MS is clearly an R-module.
• Also of interest is the R-submodule FS of MS that consists of functions f : S → R such that f (s) = 0R
for all but finitely many s. The R-module FS is called the free R-module on the set S and will be very
important for us.
30
M3P8 Algebra III 7 R-modules
Example. Let M be an R-module and I an ideal of R. Then we can form the R-submodule IM of M
consisting of all elements of M of the form
i1 m1 + · · · + ir mr ,
where the ij are in I and the mj are in M . This is an R-submodule of M , so we can form the quotient
M/IM . Then M/IM is certainly an R-module, but it is also an R/I-module. One can define multiplication
R M M
× −→
I IM IM .
(r + I, m + IM ) 7−→ rm + IM
As always one has to check that this is well-defined, but this is straightforward. We need that if r − r0 lies
in I, and m − m0 lies in IM , then rm − r0 m0 lies in IM . But
rm − r0 m0 = (r − r0 ) m + r0 (m − m0 ) ,
which is clearly in IM .
31
M3P8 Algebra III 7 R-modules
Then f can be written as r1 es1 + · · · + rn esn , so the es span FS . On the other hand, for all s1 , . . . , sn ∈ S
distinct with
X n
ri esi = 0,
i=1
Pn
i=1 ri esi takes the value ri by evaluating at si for all i, and thus is only the zero function when all ri are
zero for all i, so we do have R-linear independence. Thus FS is free, justifying its name.
Proposition 7.4.4. Let F1 , F2 be free R-modules with basis S1 , S2 . Then F1 ⊕ F2 is free with basis
{(s, 0) | s ∈ S1 } ∪ {(0, s0 ) | s0 ∈ S2 } .
Moreover, if F1 and F2 are free of finite ranks n1 and n2 respectively, then F1 ⊕ F2 is free of rank n1 + n2 .
Proof. For linear independence, let s1 , . . . , sm ∈ S1 and s01 , . . . , s0l ∈ S2 be distinct. Suppose we have
r1 , . . . , rm , r10 , . . . , rl0 ∈ R such that
r1 (s1 , 0) + · · · + rm (sm , 0) + r10 (0, s01 ) + · · · + rl0 (0, s0l ) = 0.
Then
r1 s1 + · · · + rm sm = 0 ∈ M1 , r10 s01 + · · · + rl0 s0l = 0 ∈ M2 ,
so ri , ri0 = 0. For spanning set, let (m, m0 ) ∈ M1 ⊕ M2 . Write
m = r1 s1 + · · · + rm sm , si ∈ S1 , m0 = r10 s01 + · · · + rl0 s0l , s0i ∈ S2 ,
then
(m, m0 ) = r1 (s1 , 0) + · · · + rm (sm , 0) + r10 (0, s01 ) + · · · + rl0 (0, s0l ) .
Thus S1 ∪ S2 is a basis for F1 ⊕ F2 , which immediately proves the claim.
32
M3P8 Algebra III 7 R-modules
φf : FS → M,
Note that this is a finite sum since all but finitely many s have g (s) = 0. Then it is clear that this is a
homomorphism of R-modules. On the other hand suppose φ is any other map FS → N with φ (es ) = f (s)
for all s. Then we can write X
g= g (s) es ,
s∈S, g(s)6=0
so uniqueness is clear.
The image of φf is the R-submodule of N generated by the elements f (s), for s ∈ S.
Corollary 7.4.6. Let M be a free R-module with a basis T for M . Let S be any set of the same cardinality
as T , and let g : T → S be any bijection. Then the map φf : FS → M is an isomorphism. In particular, any
two free R-modules of the same rank are isomorphic.
Proof. The map φf : FS → M is such that φf (es ) = f (s). Since elements of T are linearly independent,
this map is injective. Suppose φf (g) = 0. Can write
X
g= ri esi ,
i
P
for si distinct, then φf (g) = i ri f (si ). Since si are distinct, f (si ) are distinct elements of T , so
X
ri f (si ) = 0.
i
Thus M is isomorphic to FS . Since M was arbitrary, any R-module of rank equal to the cardinality of S is
isomorphic to FS and the result follows.
Note. It is also true, but harder to prove, that if M, N are free of different ranks, then M N .
33
M3P8 Algebra III 7 R-modules
and let A be the s by t matrix whose i, j entry is rij . Then A gives a map from Rt to Rs , and the quotient
of Rs by the R-submodule ARt of Rs is isomorphic to M .
Example.
• Let R = Z and M = Z/nZ generated by [1]n . The map Z → M is the quotient map with kernel hni.
So presentation matrix is just (n).
√ √
• Let R = Z −5 and I = 2, 1 + −5 . Then
R2 −→ I √ .
(e1 , e2 ) 7−→ 2, 1 + −5
√ √
Since (2) (3) = 1 + −5 1 − −5 ,
√ √
2e2 − 1 + −5 e1 7→ 0, 3e1 − 1 − −5 e2 7→ 0.
√ √
Claim that the two relations 1 + −5 e1 − 2e2 and 3e1 − 1 − −5 e2 generate K. Let ae1 + be2
√ √
be a relation, so a, b ∈ R and 2a + 1 − −5 b = 0, that is a = 1 + −5 /2 b. A question is for
√
which b does 1 + −5 /2 b lie in R? Claim that the set of such b is an ideal J of R. 2 ∈ J and
√ √ √ √
1 − −5 ∈ J so J contains 2, 1 − −5 . 1 ∈ / J, since 2, 1 − −5 is maximal, J = 2, 1 − −5 .
So we have a presentation matrix
√
1 + −5 3√
: R2 → R2 .
−2 −1 + −5
General idea is if we have a presentation matrix A : Rt → Rs for M , with s rows and t columns, then BAC
is also a presentation matrix for M , where B is s × s and C is t × t, and B and C are invertible matrices
with inverse matrix entries in R.
34
M3P8 Algebra III 8 Noetherian rings and modules
I1 ⊆ I2 ⊆ . . .
Proof. Suppose first that M is Noetherian, and let N be an R-submodule of M . Choose an element n0 of
N , and let N0 be the R-submodule of N generated by n0 . If N0 is all of N , then N is finitely generated.
Otherwise, choose n1 in N \ N0 , and let N1 be the R-submodule of N generated by n0 and n1 . If N is not
finitely generated, we may continue this process indefinitely, choosing for each i an ni in N \ Ni−1 , which is
nonempty since N is not finitely generated, and letting Ni be generated by n0 , . . . , ni . In this way we obtain
a strictly increasing infinite chain
N0 ( N1 ( . . .
of R-submodules of M , contradicting the fact that M is Noetherian. Conversely, suppose that every R-
submodule of M is finitely generated, and let
M0 ⊆ M1 ⊆ . . .
be an increasing chain. We must show that this chain is eventually constant. Let N be the union of the
R-submodules Mi . Note that N is an R-submodule of M . Thus N is finitely generated, say by n1 , . . . , ns .
If n1 , n2 ∈ N , then there exist i, j with n1 ∈ Mi and n2 ∈ Mj . If d ≥ i, j, n1 , n2 ∈ Md , so n1 + n2 ∈ Md , so
n1 + n2 ∈ N . Since N is the union of the Mj , there exist i1 , . . . , is such that nj is in Mij for all j. Let d be
the largest of the ij . Then Md contains n1 , . . . , ns so it contains N . In particular for any d0 ≥ d we have
N ⊆ Md ⊆ Md0 ⊆ N,
35
M3P8 Algebra III 8 Noetherian rings and modules
• to use these properties to show R is Noetherian implies that R [X] is Noetherian and other consequences.
The goal of this section is to prove the following theorem.
Theorem 8.2.1. Any finitely generated R-module M over a Noetherian ring R is Noetherian.
We proceed in several steps. First note the following.
Proof.
1. Since M is Noetherian, any R-submodule of M is finitely generated, and thus any R-submodule of N
is finitely generated.
2. Given a R-submodule N 0 of M/N , let N f0 be its preimage in N under the canonical quotient map
f : M → M/N . We have a surjection from N f0 to N 0 induced by f . N
f0 ⊆ M , so Nf0 is finitely
generated, say by n1 , . . . , ns . Claim that
f (n1 ) , . . . , f (ns )
e = r1 n1 + · · · + rs ns ,
n ri ∈ R.
Then
n) = r1 f (n1 ) + · · · + rs f (ns ) .
n = f (e
36
M3P8 Algebra III 8 Noetherian rings and modules
Proposition 8.2.3. Let M be an R-module, let N be a Noetherian R-submodule of M , and suppose that
M/N is Noetherian. Then M is Noetherian.
Proof. Let M 0 be a R-submodule of M . Then M 0 ∩ N is a R-submodule of M , hence finitely generated. Let
a1 , . . . , as ∈ M 0 ∩ N generate M 0 ∩ N . Let M 0 denote the image of M 0 in M/N . This is a R-submodule of
M/N and thus finitely generated. Let b1 , . . . , bt ∈ M 0 ⊆ M/N generate M 0 , and choose elements b1 , . . . , bt
of M 0 mapping to b1 , . . . , bt in M/N , respectively. We now show that
a1 , . . . , as , b1 , . . . , bt
is a generating set for M , proving the claim. Given any m ∈ M 0 , let m be its image in M/N under
0
37
M3P8 Algebra III 9 Polynomial rings in several variables
then
bm m−n
deg P (X) − X Q (X) < deg (P (X)) .
an
Over a ring, this only goes so far. Over Z, cannot use a multiple of 3X + 4 to reduce the degree of an X n + . . .
unless 3 | an . Let
P (X) = b0 + · · · + bn X n , bn ∈ R∗ .
We say that bn is the leading coefficient of P (X). In general, if I have Q1 (X) , . . . , Qr (X) with degrees
d1 , . . . , dr and leading coefficients a1 , . . . , ar and P (X) of degree d ≥ d1 , . . . , dr then there exist n1 , . . . , nr ∈
R such that
deg P (X) − n1 X d−d1 Q1 (X) − · · · − nr X d−dr Qr (X) < d,
Lemma 9.1.2. Let R be Noetherian and I ⊆ R [X] be an ideal. Let J ⊆ R be the set of leading coefficients
of polynomials in I. That is, the set of a ∈ R such that there exists a polynomial P (X) in I with leading
coefficient a. Then J is an ideal of R.
Proof. Certainly if a ∈ J is the leading coefficient of P (X) ∈ I such that
P (X) = aX n + . . . ,
rP (X) = raX n + . . . ,
so ra ∈ J, so J is closed under multiplication. On the other hand, if a, b ∈ J are the leading coefficients of
P (X) and Q (X) in I, then let n, m be the degrees of
P (X) = aX n + . . . , Q (X) = bX m + . . .
respectively. Without loss of generality we may assume n ≥ m. Then a + b is the leading coefficient of
and the latter polynomial is in I so a + b ∈ J. Thus J is closed under addition, and is therefore an ideal.
38
M3P8 Algebra III 9 Polynomial rings in several variables
Now since R is Noetherian, J is finitely generated, say by a1 , . . . , as ∈ R. By definition of J, there are thus
polynomials P1 , . . . , Ps in I, of degrees d1 , . . . , dn , such that
Pi = ai X di + . . .
has leading coefficient ai for all i. Let N be the largest of the di .
Lemma 9.1.3. Given Q (X) ∈ I of degree d ≥ N . Then there exist R1 (X) , . . . , Rs (X) ∈ R [X] such that
Q (X) − R1 (X) P1 (X) − · · · − Rs (X) Ps (X)
has degree less than N .
Proof. The proof is by induction on d and the base case d < N is clear by setting Ri = 0 for all i. Suppose
the claim is true for polynomials of degree less than or equal to d − 1, with d ≥ N . Let a ∈ J be the leading
coefficient of
Q (X) = aX d + . . . ,
so that Q (X) − aX d has degree at most d − 1. Since a lies in J we can write
a = r1 a1 + · · · + rs as .
Then the leading term of the polynomial
r1 X d−d1 P1 (X) + · · · + rs X d−ds Ps (X)
is aX d , so the difference
Q (X) − r1 X d−d1 P1 (X) − · · · − rs X d−ds Ps (X)
has degree at most d − 1 and lies in I. By the inductive hypothesis this difference is an R [X]-linear
combination of the Pi (X),
R1 (X) P1 (X) + · · · + Rs (X) Ps (X) .
So
Q (X) = R1 (X) + r1 X d−d1 P1 (X) + · · · + Rs (X) + rs X d−ds Ps (X)
is as well.
The following proof is due to Emmy Noether, and is a vast simplification of Hilbert’s original proof.
Proof of Theorem 9.1.1. Let I be an ideal of R [X]. We want to show that I is finitely generated. Let
I≤N = I ∩ R [X]≤N
be the subset of I consisting of all polynomials of degree at most N . Then I≤N is an R-submodule of the
R-module R [X]≤N of all polynomials of degree at most N . The latter is free of rank N + 1 and generated by
1, . . . , X N as an R-module, so it is finitely generated, hence Noetherian. In particular since R is Noetherian
I≤N is also a finitely generated R-module. Let T1 (X) , . . . , Tk (X) generate I≤N as an R-module. We will
show that
P1 (X) , . . . , Ps (X) , T1 (X) , . . . , Tk (X)
generate I as an R [X]-module. More precisely, we will show that Q (X) is an R [X]-linear combination of
the Pi (X) and Tj (X). Given Q (X) ∈ I, there exist R1 (X) , . . . , Rs (X) ∈ R [X] such that
Q (X) = R1 (X) P1 (X) + · · · + Rs (X) Ps (X) + T (X) ,
with T (X) ∈ I≤N . There exist r1 , . . . , rk ∈ R such that
T (X) = r1 T1 (X) + · · · + rk Tk (X) ,
so
Q (X) = R1 (X) P1 (X) + · · · + Rs (X) Ps (X) + r1 T1 (X) + · · · + rk Tk (X) .
39
M3P8 Algebra III 9 Polynomial rings in several variables
R [X1 , . . . , Xn ] −→ S
,
Xi 7−→ si
where f : R → S, is surjective.
Note. Any finitely generated R-algebra S is isomorphic to a quotient R [X1 , . . . , Xn ] /I for some n and some
ideal I. Thus we can rephrase the Hilbert basis theorem as saying that if R is Noetherian, then any finitely
generated R-algebra is Noetherian.
Lecture 17
Lecture 17 is a problem class. Monday
12/11/18
9.2 Polynomial rings over UFDs are UFDs Lecture 18
Wednesday
Our next goal is to study factorisation in polynomial rings of the form R [X]. Z [X] is not a PID nor a UFD.
14/11/18
The idea is to relate factorisations in Z [X] to factorisations in Q [X]. Warning that irreducibility in Q [X]
does not imply irreducibility in Z [X].
Example. 3x + 15 is irreducible in Q [X]. In Z [X]
3x + 15 = 3 (x + 15) .
Certainly if R is not a UFD then we cannot expect to have unique factorisation in R [X], since we do not even
have it in R. Assume R is a UFD. Then the ring R [X] might be quite complicated, but R [X] is contained in
a much simpler ring where we do understand factorisation, the ring K [X], where K is the field of fractions
of R. Our goal will thus be to compare factorisations in K [X] with factorisations in R [X]. Fundamental
question is can we turn factorisations in K [X] of P (X) ∈ R [X] into factorisations in R [X]? The key to
doing this is the following result, often called Gauss’ lemma.
Theorem 9.2.1 (Gauss’ lemma). Let R be a UFD and let K be its field of fractions. Let P (X) ∈ R [X],
and let Q (X) be a polynomial in K [X] that divides P (X) in K [X]. Then there is an element α ∈ K ∗ such
that αQ [X] lies in R [X], and divides P (X) in R [X]. In particular, if P (X) is reducible in K [X], then
P (X) is also reducible in R [X].
Proof. Write
P (X) = Q (X) T (X) ∈ K [X] ,
and choose nonzero elements e1 , e2 ∈ R such that e1 Q (X) and e2 T (X) have coefficients in R, and so that
the greatest common divisor of the coefficients of Q (X) is one, as is the greatest common divisor of the
coefficients of T (X). Letting d = e1 e2 , we have
40
M3P8 Algebra III 9 Polynomial rings in several variables
so R/ hqi [X] is as well. Moreover, if Q0 (X) and T 0 (X) are the images of Q0 (X) and T 0 (X) mod hqi in
R/ hqi [X], then we have
dP (X) = Q0 (X) T 0 (X) ,
so 0 = Q0 (X) T 0 (X) in R/ hqi [X]. Since R/ hqi [X] is an integral domain we must have either Q0 (X) = 0 or
T 0 (X) = 0 in R/ hqi [X]. Without loss of generality assume Q0 (X) = 0. Then all the coefficients of Q0 (X)
are divisible by q. Thus
and Q1 (X) is a multiple of Q (X) in K [X]. If d1 is a unit, done. Otherwise write d1 = d2 q1 for q1 irreducible.
Same trick implies that
and Q2 (X) is a multiple of Q1 (X) in K [X]. Repeat, contradicting our construction of Q0 (X). Thus α = d
is a unit in P (X), and we have
1 1
P (X) = e1 Q (X) e2 T (X) , e1 Q (X) , e2 T (X) ∈ R [X] .
d d
Note. The converse to the last claim of Theorem 9.2.1 is not true. If P (X) is reducible in R [X], it might
be irreducible in K [X].
Example. The polynomial 7x factors into irreducibles as 7 · x in Z [X], but since 7 is a unit in Q [X], 7x is
irreducible in Q [X].
The following lemma shows that this kind of thing is all that can happen, however.
Proposition 9.2.2. Let P (X) in R [X] be a polynomial and suppose that the greatest common divisor of
all of its coefficients is one. Then P (X) is irreducible in K [X] if and only if it is also irreducible in R [X].
Proof. Suppose P (X) is irreducible in R [X], and write
where Q (X) and T (X) are nonunits. If Q (X) or T (X) were constant with degree zero then it would divide
every coefficient of P (X) and thus divide the GCD of those coefficients, making it a unit. Thus Q (X) and
T (X) are nonconstant with positive degree and the factorisation
is also a nontrivial factorisation in K [X], so P (X) is reducible in K [X]. Conversely suppose P is reducible
in K [X]. Then there exist Q (X) ∈ K [X] with
41
M3P8 Algebra III 9 Polynomial rings in several variables
P (X) = dQ (X) ,
where the greatest common divisor of the coefficients of Q (X) is one. Since R is a UFD, d factors into
irreducibles q1 , . . . , qs in R, and these remain irreducible in R [X], so it suffices to show that Q (X) factors
into irreducibles. Factor Q (X) into irreducibles in K [X],
α1 . . . αr = 1, αi Qi (X) ∈ R [X] .
Let Q0i (X) = αi Qi (X). GCD of coefficients of Q01 (X) , . . . , Q0r (X) is one, so Q01 (X) , . . . , Q0r (X) are irredu-
cible in R [X] since they are irreducible in K [X]. For uniqueness of factorisations, it remains to show that
if P (X) ∈ R [X] is irreducible in R [X] and divides A (X) B (X) in R [X] for A (X) , B (X) ∈ R [X], then
P (X) divides either A (X) or B (X) in R [X].
• If P (X) is constant, then P (X) = c is irreducible in R. In R/ hci [X] a domain,
0 = A (X) B (X) ,
αP (X) ∈ R [X] , αP (X) | A (X) ∈ R [X] , A (X) = αP (X) α−1 Q (X) ∈ R [X] ,
by Gauss’ lemma. On the other hand, since P (X) is irreducible in R [X] the GCD of its coefficients
is one, so the only way αP (X) lies in R [X] is if s is a unit and α lies in R. Thus α−1 Q (X) ∈ R [X],
α ∈ R, and P (X) ∈ R [X], so P (X) also divides A (X).
Corollary 9.2.4. If K is a UFD, a field, or a PID, then K [X1 , . . . , Xn ] is a UFD for any n.
Warning that quotients of UFDs are only rarely UFDs themselves.
√
Example. Z [X] is a UFD, but Z [X] / X 2 + 5 = Z −5 is not a UFD.
42
M3P8 Algebra III 9 Polynomial rings in several variables
43
M3P8 Algebra III 9 Polynomial rings in several variables
Proposition 9.3.4. Let Q (X) be a monic polynomial in R [X], and let p be a prime ideal of R. Suppose
that the mod p reduction Q (X) is irreducible in R/p [X]. Then Q (X) is irreducible in R [X].
Proof. Suppose Q (X) were reducible in R [X]. Since Q (X) is monic, Q (X) must factor as
A (X) B (X) ,
where both A (X) and B (X) are not units. Can assume A (X) , B (X) are monic of degree deg (A) , deg (B) >
0, since leading coefficients of A, B multiply to one. Then Q (X) factors in R/p [X] as
A (X) B (X) ,
where both are monic of positive degree between 1 and deg Q (X) − 1, so Q (X) is also reducible.
This means, for instance, that we can show that a monic polynomial in Z [X] is irreducible if we can find
even one prime p for which it is irreducible mod p.
Example.
X 2 + aX + b ∈ Z [X] ,
with a, b odd is irreducible in Q [X]. Its reduction mod 2 is
X 2 + X + 1,
If p = 2k + 1,
p2 = 4k 2 + 4k + 1 = 4 k 2 + k + 1 = 8m + 1,
since k 2 + k is even. 2
−1
Xp − 1 = X 8m − 1 = X 4 + 1 X 8m−4 − · · · − 1 .
There is another sufficient criterion for irreducibility by reducing mod p, known as Eisenstein’s criterion.
Proposition 9.3.5 (Eisenstein’s criterion). Let
44
M3P8 Algebra III 9 Polynomial rings in several variables
45
M3P8 Algebra III 10 Integral extensions and algebraic integers
Definition 10.1.2. Let R be a subring of a ring S, and α an element of S. We say α is integral over R if
there exists a monic polynomial P (X) ∈ R [X] with coefficients in R such that P (α) = 0 in S.
We can characterise integral elements in the following way.
Proposition 10.1.3. An element α ∈ S is integral over R if and only if the subring R [α] of S is a finitely
generated R-module.
Proof. Suppose α is integral over R, so that there exists a monic polynomial P (X) in R [X] with P (α) = 0.
Then R [α] is a quotient of R [X] / hP (X)i so 1, . . . , αd−1 , where d is the degree of P (X), span R [α] over
R. Given x ∈ R [α], can write
x = Q (α) = rn αn + · · · + r0 .
Write
Q (X) = P (X) T (X) + A (X) ∈ R [X] , deg (A (X)) < deg (P (X)) ,
so Q (α) = P (α) T (α) + A (α).
so
x = Q (α) = A (α) = a0 + · · · + ad−1 αd−1 .
Conversely, if R [α] is finitely generated as an R-module, say by x1 , . . . , xr ∈ R [α], we can write
Let n be larger than the degree di of all the Qi (X). We can write αn ∈ R [α] as
r
X r
X
si xi = si Qi (α) , si ∈ R.
i=1 i=1
So let
r
X
P (X) = X n − si Qi (X) ∈ R [X] .
i=1
46
M3P8 Algebra III 10 Integral extensions and algebraic integers
Lemma 10.1.6. Let R ⊆ S ⊆ T be rings, such that S is finitely generated as an R-module and T is finitely
generated as an S-module. Then T is finitely generated as an R-module.
Proof. Let t1 , . . . , tl generate T over S, and let s1 , . . . , sm generate S over R. Then for any element t of T ,
we can write
Xl
t= ai ti , ai ∈ S.
i=1
so that
l X
X m
t= bji sj ti ,
i=1 j=1
S 0 = R [s0 , . . . , sn−1 ] ⊆ S.
Since each si is integral over R, s0 is in particular integral over R and si is integral over R [s0 , . . . , si−1 ].
Thus R [s0 ] is a finitely generated R-module and R [s0 , . . . , si ] is a finitely generated R [s0 , . . . , si−1 ]-module
for each i by induction. By Lemma 10.1.6 above, S 0 is a finitely generated R-module. Since t is integral
over S 0 , S 0 [t] is a finitely generated S 0 -module, and hence a finitely generated R-module by Lemma 10.1.6.
Since R [t] is contained in S 0 [t] and R is a Noetherian ring, R [t] is a finitely generated R-module and thus t
is integral over R.
Corollary 10.1.8. Let R be a Noetherian subring of S and suppose α, β ∈ S are integral over R. Then αβ
and α + β are integral over R.
Proof. The ring R [α] is a finitely generated R-module and thus integral over R. Since β is integral over R
it is integral over R [α]. Thus R [α, β] = R [α] [β] is integral over R [α] and hence over R by Lemma 10.1.6.
Since α + β and αβ lie in R [α, β] they are integral over R.
Lecture 21
Definition 10.1.9. Let R be a Noetherian subring of S. The integral closure of R in S is the subset Wednesday
of S consisting of all elements s ∈ S that are integral over R. This is a subring of R. We say that R is 21/11/18
integrally closed in S if every element in S that is integral over R is contained in R, so R is equal to its
integral closure in S. If R is an integral domain, we say that R is integrally closed if R is integrally closed
in its field of fractions K.
Lemma 10.1.10. Let R be a Noetherian subring of S, and let R0 be the integral closure of R in S. Then
R0 is integrally closed in S.
Proof. Let t be an element of R integral over R0 . Then R0 [t] is a finitely generated R0 -module, so is integral
over R0 , and R0 is integral over R. Thus R0 [t] is integral over R, so t is integral over R and thus lies in
R0 .
√ √ √
Example. Z −3 is integral over Z. As a Z-module, Z −3 is generated by 1, −3. It is not integrally
√
closed. 1+ 2 −3 is in the field of fractions of Z −3 and is a root of X 2 − X + 1.
√
47
M3P8 Algebra III 10 Integral extensions and algebraic integers
so m2 ≡ 1 mod 4, this can only happen if d is odd and b = n/2 with n odd. We then have m2 − n2 d
is a multiple of four. Since m2 , n2 ∈ Z are odd they are congruent to 1 mod 4, so this is only possible
if d is congruent to 1 mod 4.
√ √
Thus if α is an algebraic integer, either α = a + b d with a, b ∈ Z, or α = m+n 2
d
with m, n ∈ Z odd and
d congruent to 1 mod 4.
√ Conversely, it is easy to check that all such elements are algebraic integers. To
summarise, if K = Q d for d squarefree, we thus have
h√ i n √ o
Z d = a + b d | a, b ∈ Z d ≡ 2, 3 mod 4
OK = h √ i n √ o .
Z 1+ d = m+n d | n, m ∈ Z, n ≡ m mod 2 d≡1 mod 4
2 2
Note. The results in this section are also true without any Noetherian hypotheses, but the proofs are more
difficult, and require machinery we have not covered.
48
M3P8 Algebra III 11 Dedekind domains
11 Dedekind domains
Lecture 22
11.1 Dedekind domains Friday
23/11/18
For number theorists, it is often convenient to work in a ring of the form Z [α], where α ∈ C is an algebraic
integer, or more generally in some subring O of C that is integral over Z. Unfortunately, unique factorisation
only rarely holds in such rings. If O is integrally closed, however, there is a substitute for unique factorisation
that is often good enough, unique factorisation of ideals. In this section we develop the ideas behind this
result, in the more general context of what are called Dedekind domains.
Definition 11.1.1. An integral domain R is called a Dedekind domain if
• R is Noetherian,
• R is integrally closed, and
• every nonzero prime ideal of R is maximal, so R has dimension one.
Example.
• In particular, any PID is a Dedekind domain. We have seen that every nonzero prime ideal is maximal
in a PID, and PIDs are certainly Noetherian. They are integrally closed because any UFD is integrally
closed.
• The rings OK , the integral closure of Z in K, with K a quadratic extension of Q are
h√ i
Z d d ≡ 2, 3 mod 4
OK = h √ i .
1+ d
Z
2 d ≡ 1 mod 4
49
M3P8 Algebra III 11 Dedekind domains
Note the resemblance of this to the property, p | ab implies p | a or p | b for p irreducible, which holds in UFDs
and implies unique factorisation. We might hope that the above result thus implies unique factorisation into
primes for arbitrary rings, but this is too much to ask for. The problem is that ideal multiplication is usually
badly behaved compared to multiplication of elements in integral domains. Recall that for I, J ⊆ R ideals,
IJ is the ideal generated by all elements of the form rs for r ∈ I and s ∈ J. In particular if r1 , . . . , rn
generate I and s1 , . . . , sm generate J then
r1 s1 , . . . , r1 sm , . . . , rn s1 , . . . , rn sm
generate IJ.
√
Example. R = Z −3 is not a Dedekind domain, since it fails to be integrally closed. Then the ideal
√
I = 2, 1 + −3 is prime, and we have
√ 2 √ √ √
2, 1 + −3 = 4, 2 + 2 −3, −2 + 2 −3 = 4, 2 + 2 −3 .
There is thus a chain of inclusions
√ √
2, 1 + −3 = I ) h2i ) I 2 = 4, 2 + 2 −3 ,
so the ideal h2i is not a product of prime ideals. Worse is I 2 = h2i I but I 6= h2i, so cannot cancel ideals.
Dedekind domains give precisely the context where this does not happen. In order to make this precise, we
first define the following.
Definition 11.1.4. Let R be a Noetherian integral domain. A fractional ideal of R is a finitely generated
nonzero R-submodule of the field of fractions K of R. A principal fractional ideal is an R-submodule
R · x ⊆ K of K finitely generated by a single nonzero element x for x ∈ K ∗ .
Example. The subgroup of Q generated by 3/5 is a principal fractional ideal of Z. Indeed, every fractional
ideal of Z, or any PID, is principal.
More generally, let R be a Noetherian integral domain, and let I be the R-submodule of K generated by
r1 , . . . , rn ∈ K. Then by definition I is a fractional ideal of R. On the other hand, we can clear denominators.
There exists an r ∈ R, nonzero, such that rri lies in R for all i. Then rI is generated by elements of R, so
is an ideal J of R, and I = 1r J. Thus the fractional ideals of R are precisely the subsets of K of the form
1
r J, where r is a nonzero element of R and J is an ideal of R. Let I and J be fractional ideals of R. The
product IJ is the R-submodule of K generated by all products of the form rs for r ∈ I and s ∈ J. It is a
fractional ideal of R. The multiplication I, J 7→ IJ is an associative and commutative operation.
Note. R is a fractional ideal of R, and RJ = J for any fractional ideal J, so R is an identity element for this
operation.
For a nonzero ideal I of R, let I −1 denote the set
{r ∈ K ∗ | rI ⊆ R} .
Then I −1 is clearly an R-submodule of K. If r ∈ I is nonzero, then rI −1 , by definition, is contained in R,
so I −1 is contained in 1r · R and is thus a fractional ideal. Warning that in a general ring, I 7→ I −1 is not
always a good inverse operation. II −1 ⊆ R but need not equal R. For a prime ideal p of R, and n ∈ Z>0 ,
−n −1 n
define p = p . We then have the following.
Theorem 11.1.5. Let R be a Dedekind domain. Then
• the set of fractional ideals of R form a group under multiplication I, J 7→ IJ, identity R, and inverse
I 7→ I −1 , and
• moreover, any fractional ideal I of R factors uniquely as
pn1 1 . . . pns s , ni ∈ Z,
where the pi are nonzero prime ideals.
50
M3P8 Algebra III 11 Dedekind domains
The proof of this statement will occur in several steps. We first show the following. Lecture 23
Monday
Proposition 11.1.6. Let I be a nonzero ideal of a Noetherian ring R. Then there exist nonzero primes 26/11/18
p1 , . . . , pr and n1 , . . . , nr ∈ Z>0 such that I contains
pn1 1 . . . pnr r .
Proof. Suppose the claim fails for some I. Then there exists an ideal I such that
• I does not contain a product of primes, but
• every ideal containing I does.
Suppose not. Then let I0 be an ideal satisfying 1. Since 2 does not hold for I0 there exists I1 ) I0 such that
1 holds for I1 . Then 2 cannot hold for I1 , so there exists I2 ) I1 such that 1 holds for I2 , etc, so get infinite
increasing chain
I0 ( I1 ( . . . ,
contradicting Noetherianness of R. Fix such an I. Certainly I cannot be prime. So there exist a, b ∈ R with
ab ∈ I but a and b not in I. Then the ideals I + hai and I + hbi both strictly contain I, so the claim holds
for both of these ideals by 2, so
I + hai ⊇ p1 . . . pr , I + hbi ⊇ q1 . . . qs ,
for pi and qj prime ideals. Then it also holds for their product
p1 . . . pr q1 . . . qs ⊆ (I + hai) (I + hbi) ⊆ I
so x satisfies a monic polynomial with coefficients in R. Thus x is integral over R. Since R is integrally
closed and x does not lie in R this is a contradiction.
When R is not integrally closed this is false.
51
M3P8 Algebra III 11 Dedekind domains
√
1+ −3
√
Example. If R = Z −3 , I ⊂ R, and x = 2 , then
√ √ √
xI = 1 + −3, −1 + −3 = 2, 1 + −3 = I.
Proposition 11.1.8. Let p be a nonzero prime ideal of a Dedekind domain R. Then p−1 · p = R.
Proof. We first show that there is an element x ∈ p−1 such that x ∈
/ R. Let a be an element of p, and a 6= 0,
so that we have hai ⊂ p. Choose a minimal set of nonzero primes p1 , . . . , pr such that
p1 . . . pr ⊆ hai .
by
xy = , by ∈ p1 . . . pr ⊆ hai .
a
Thus xy lies in R. By definition, this means x lies in p−1 but not in R. Now consider p−1 · p. By definition
this is contained in R. Since p ⊆ R, 1 ∈ p−1 so p−1 · p contains p. Since p is a nonzero prime ideal it is
maximal, so we must have either p−1 · p = R or p−1 · p = p. Suppose the latter holds. Then in particular
multiplication by x sends p to p. This contradicts Lemma 11.1.7 above, so p−1 · p ) p.
Proposition 11.1.9. Let I be a nonzero ideal of a Dedekind domain R. Then there exists a fractional ideal
J of R such that IJ = R.
Proof. Suppose otherwise. Then there is a maximal nonzero ideal I of R for which no such J exists.
Proposition 11.1.8 shows that I is not a maximal ideal, so I is properly contained in some maximal ideal p
of R. Then p−1 is contained in I −1 . We thus have inclusions
I ⊆ Ip−1 ⊆ II −1 ⊆ R.
Suppose that Ip−1 = I. By Proposition 11.1.8 there exists x ∈ p−1 not in R, so we would have xI ⊂ I
contradicting Lemma 11.1.7 above. Thus Ip−1 strictly contains I and thus has an inverse
−1
J = Ip−1 .
II −1 = 1r J −1 · rJ = JJ −1 = R.
Lecture 24
Lecture 24 is a problem class. Wednesday
28/11/18
52
M3P8 Algebra III 11 Dedekind domains
It remains to show that every fractional ideal of R factors uniquely as a product of prime powers. The hard Lecture 25
part is showing such factorisations exist, and we make heavy use of the fact that the fractional ideals are a Friday
group. Uniqueness is then almost an afterthought. 30/11/18
Proposition 11.1.11. Every fractional ideal in a Dedekind domain R is uniquely expressible as a product
of, possibly negative, prime powers
pn1 1 . . . pns s , ni ∈ Z,
where the pi are primes.
Proof. We first show that every nonzero ideal I in R is a product of nonnegative prime powers.
• Suppose otherwise, that for some ideal I of R, I cannot be expressed as a product of primes. Claim
that there is a largest ideal I 0 such that I 0 cannot be expressed as a product of primes but all J ) I
can. Take I0 = I if there exists I1 ) I that cannot be expressed as a product of primes, either all ideals
properly containing I1 can be or there exists I2 ) I1 that cannot be expressed as a product of primes.
Since R is Noetherian, the process terminates. Now let I 0 be as in the claim. Certainly I 0 6= R, and I 0
is not prime, since every maximal ideal of R is certainly such a product I 0 cannot be a maximal ideal.
Thus I 0 is properly contained in a maximal ideal p. Then
J = p−1 · I 0
But then
p · J = I 0 = ppn1 1 . . . pns s
is also a product of prime powers, contradicting our assumption.
• Now suppose that I is a fractional ideal. Then I = 1r J for some nonzero ideal J of R and some nonzero
element r of R. Since
hri = qm mt
1 . . . qt ,
1
J = pn1 1 . . . pns s
factor as products of prime powers, so does
−1
I = 1r J = hri J = pn1 1 . . . pns s q−m
1
1
. . . qt−mt .
It remains to show that such factorisations are unique. Suppose otherwise for a fractional ideal I. Then we
have a finite collection of distinct primes p1 , . . . , ps and q1 , . . . , qt , and two sequences
n1 , . . . , ns , m1 , . . . , mt ∈ Z,
such that
I = pn1 1 . . . pns s = qm mt
1 . . . qt ,
1
and we must show that mi = ni for all i. Suppose this is not the case. We can make all prime powers ni , mj
m
involved positive by cancelling pni i and qj j from both sides of the equation. We then get an expression of
the form
pn1 1 . . . pns s = qm mt
1 . . . qt ,
1
where the primes {pi } , {qj } are all distinct and all powers ai and bj are positive. Claim that both products
must be empty under these assumptions. Recall that if R is Noetherian, p prime, I, J ideals, then p contains
IJ implies that p contains I or p contains J. If one product, say the pi ’s, is nonempty, then since p1 divides
the left hand side it also divides the right hand side, and thus contains one of the qi ’s for some i. Since pi
and qj are maximal in a Dedekind domain, p1 = qi , which contradicts disjointness and is impossible.
53
M3P8 Algebra III 11 Dedekind domains
−1 1
(rR) R
=
r
(rR) (sR) = rsR.
Denote this subgroup by P (R). We can then form a quotient group called the ideal class group of R.
Definition 11.2.1. The ideal class group of R, denoted A (R), is
I (R)
A= .
P (R)
54
M3P8 Algebra III 12 Integers in number fields
The matrix of course depends on the basis β1 , . . . , βd chosen, but its trace and determinant are elements of
K that depend only on α. We denote the trace of Mα by T rL/K (α) ∈ K and call it the trace of α with
respect to L/K. Similarly, the determinant of Mα is denoted NL/K (α) ∈ K and called the norm of α.
The map
NL/K : L −→ K
α 7−→ NL/K (α)
is multiplicative, so
NL/K (αα0 ) = NL/K (α) NL/K (α0 ) .
Proof. Since λ ∈ K, distributivity of multiplication over addition shows that, with respect to a fixed basis
of L over K, Mλα+α0 = λMα + Mα0 , so
NL/K (αα0 ) = det (Mαα0 ) = det (Mα ) det (Mα0 ) = NL/K (α) NL/K (α0 ) .
We will prove that if R is a PID or a Dedekind domain, such as Z, K is its field of fractions, L/K is finite
such that T rL/K is not the zero map, then the integral closure of R in L is Dedekind.
55
M3P8 Algebra III 12 Integers in number fields
Lecture 26
Proposition 12.2.2. Let L/K be a finite extension, and α an element of L. Let
Monday
Q (X) = X n + an−1 X n−1 + · · · + a0 03/12/18
56
M3P8 Algebra III 12 Integers in number fields
Remark 12.2.3. If L/K is finite, T rL/K (1) = [L : K], considered as an element of K. The map T rL/K :
L → K is sometimes the zero map. However, this does not happen if K has characteristic char (K) = 0 or
if the degree d = [L : K] is relatively prime to the characteristic of K, since the above Proposition 12.2.2
shows that T rL/K (1) = d 6= 0.
Example. Let K = F2 (t), the rational functions with coefficients in F2 .
K [X] 1
L= 2
= F 2 t2 .
hX − ti
T rL/K is K-linear. T rL/K (1) = 0 and T rL/K (X) = 0, so T rL/K (aX + b) = 0 for all a, b ∈ K.
Proposition 12.2.4. If L/K are finite fields, then T rL/K is not zero map.
R K
.
S L
Lecture 27
To do this, choose a K-basis β1 , . . . , βd for L over K. We have seen that for each i there exists ri ∈ R∗ such Wednesday
that ri βi ∈ S, so, replacing βi by ri βi , we may assume that the βi all lie in S. Let M ⊆ S be the R-module 05/12/18
in S spanned by the βi , and let M ∗ denote
M ∗ = x ∈ L | ∀m ∈ M, T rL/K (xm) ∈ R ⊆ L.
57
M3P8 Algebra III 12 Integers in number fields
58
M3P8 Algebra III 12 Integers in number fields
In particular −a0 lies in the ideal generated by s, and hence in p. Moreover, since Q (X) is irreducible, a0
is a nonzero element of R, so q is nonzero since a0 ∈ q. Since R is Dedekind, thus q is a maximal ideal of R.
Have R ⊆ S → S/p, where kernel is R/q, a field. So get R/q → S/p. The ring S/p is an integral domain
containing the field R/q. Moreover, since S is a finitely generated R-module, if s1 , . . . , sr generate S as an
R-module, then they generate S/p as an R/q-module, so S/p is a finitely generated R/q-module. That is,
S/p is a finite dimensional R/q-vector space. We now show the following.
Lemma 12.3.7. Let K be an integral domain and let R be an integral domain containing K that is finite
dimensional as an K-vector space. Then R is a field.
Proof. Let r be a nonzero element of R. Have map
K [X] −→ R
.
x 7−→ r
Let I be the kernel. Since R is finite dimensional I = 0. Since R is an integral domain I is prime. So
K [X]
−→ R
I
x 7−→ r
is an injection, with K [X] /I a field. Then x has an inverse in K [X] /I, and this inverse maps to a
multiplicative inverse for r in R.
Lemma 12.3.7 shows that S/I is a field, so I is maximal. We have thus shown that S is Noetherian, integrally
closed, and that every nonzero prime ideal in S is maximal, so S is indeed a Dedekind domain. In particular,
for any finite extension K/Q, the integral closure OK of Z in K is a Dedekind domain, and thus has unique
factorisation of ideals. Another class of examples comes by taking K a field, letting L be a finite extension
of K (t) such that T rL/K(t) is nonzero, and letting R be the integral closure of K [t] in L. The field L is
called a function field, and the ring R is the ring of regular functions on a smooth affine algebraic
curve. Such rings R are also Dedekind domains, and they are of considerable interest in algebraic geometry.
They of course also have the unique factorisation property for ideals, and just like in ring of integers one can
consider the ideal class group. In this context, the ideal class group is also known as the Picard group.
It has a geometric interpretation in terms of line bundles on algebraic curves. Unlike in the number field
setting, the Picard group is often not a finite group.
59
M3P8 Algebra III 13 Introduction to algebraic geometry
In other words, Z (S) is the set of common zeros of all the polynomials in S.
Example.
• S = {y − x (x − 1) (x + 1)} is y = x (x − 1) (x + 1).
• S = {y (y − x (x − 1) (x + 1))} is y = 0 or y = x (x − 1) (x + 1).
• S = {y, y − x (x − 1) (x + 1)} is y = 0 and y = x (x − 1) (x + 1).
• S = y 2 − x (x − 1) (x + 1) is connected.
60
M3P8 Algebra III 13 Introduction to algebraic geometry
P = Q1 S1 + · · · + Qr Sr , Qi ∈ K [X1 , . . . , Xn ] , Si ∈ S,
Example.
61
M3P8 Algebra III 13 Introduction to algebraic geometry
One might thus hope that similarly I (Z (J)) = J for any ideal J of K [X1 , . . . , Xn ]. This cannot be literally
true, for the following reason. Recall that
rad (I) = {r ∈ K [X1 , . . . , Xn ] | ∃m, rm ∈ I} .
An ideal I is radical if rad (I) = I.
Note. In fact, for any T , the ideal I (T ) is a radical ideal.
m
P m ∈ I (T ) ⇐⇒ ∀t ∈ T, P (t) =0 ⇐⇒ ∀t ∈ T, P (t) = 0 ⇐⇒ P ∈ I (T ) .
Thus if J is not a radical ideal we cannot have I (Z (J)) = J.
However, one does have the following.
Theorem 13.2.5 (Hilbert’s Nullstellensatz). Let K be an algebraically closed field. For any ideal J of
K [X1 , . . . , Xn ], we have I (Z (J)) = rad (J).
Note. This can only possibly hold over algebraically closed fields.
Example. For n = 1, let P (X) ∈ K [X].
I (Z (P (X))) = rad (P (X)) 6= R,
unless P (X) is constant, so Z (P (X)) 6= ∅.
Corollary 13.2.6. If J ∈ K [X1 , . . . , Xn ] an ideal is not the unit ideal, then Z (J) 6= ∅.
Proof. I (Z (J)) = rad (J) 6= h1i, so Z (J) 6= ∅.
Lecture 29
We will prove this theorem later on. For now, we note that since rad (rad (J)) = rad (J) for all ideals J, the Monday
maps X 7→ I (X) and J 7→ Z (J) define a bijection 10/12/18
affine algebraic subsets of AnK .
radical ideals J ⊆ K [X1 , . . . , Xn ] !
This bijection is inclusion-reversing, and it is interesting to ask what geometric properties of X are carried
to algebraic properties of I (X) via this bijection. For instance, the following holds.
Proposition 13.2.7. Let J ⊆ K [X1 , . . . , Xn ] be a radical ideal. Then J is maximal if and only if Z (J) is
a single point.
Proof. Suppose Z (J) = {p}, where p = (p1 , . . . , pn ). I (Z (J)) = I ({p}), so
X1 − p1 , . . . , Xn − pn ∈ I ({p}) .
Then
mp = hX1 − p1 , . . . , Xn − pn i ⊆ I ({p}) ( K [X1 , . . . , Xn ] .
Note that mp is maximal. Consider the map
K [X1 , . . . , Xn ] −→ K
,
Xi 7−→ pi
and K 7→ X. The kernel of this map is hXi − pi i = mp , so
K [X1 , . . . , Xn ] ∼
= K.
mp
So mp is maximal, must have I ({p}) = mp . Thus
J = rad (J) = I (Z (J)) = mp .
Conversely, suppose J is maximal. Z (J) 6= ∅, since I (Z (J)) = J but I (∅) is the unit ideal. So there exists
p ∈ Z (J) such that
J = I (Z (J)) ⊆ I ({p}) = mp ,
so J = mp .
62
M3P8 Algebra III 13 Introduction to algebraic geometry
Proof. Let p ∈ AnK be a point of Z (J1 + J2 ). Then every element of J1 +J2 vanishes at p, so since J1 ⊆ J1 +J2
we have that p ∈ Z (J1 ). Similarly p ∈ Z (J2 ), so
Q = R + S, R ∈ J1 , S ∈ J2 .
Proof. Follows from Z (J1 + J2 ) = Z (J1 ) ∩ Z (J1 ) and using the Nullstellensatz I (Z (J)) = rad (J).
Definition 13.2.10. An affine algebraic set X is irreducible if X cannot be written as the union Y ∪ Z of
two proper affine algebraic subsets Y and Z, that is Y, Z 6= X, ∅.
Example. Let
X = Z ({y (y − x (x − 1) (x + 1))}) .
Then
Z ({y (y − x (x − 1) (x + 1))}) = Z ({y}) ∪ Z ({y − x (x − 1) (x + 1)}) ,
so X is not irreducible.
Proposition 13.2.11. An affine algebraic set X is irreducible if and only if I (X) is prime.
Proof. Suppose X is irreducible, and let f, g be elements of K [X1 , . . . , Xn ] such that f g ∈ I (X). Then
X ⊆ Z (f g) = Z (f ) ∩ Z (g) .
In particular
X = (X ∩ Z (f )) ∪ (X ∩ Z (g)) .
Since X is irreducible we must have
X = X ∩ Z (f ) ⊆ Z (f ) ,
in which case f ∈ I (X), or
X = X ∩ Z (g) ⊆ Z (g) ,
in which case g ∈ I (X). So I (X) is prime. Conversely, suppose I (X) is prime, and that X = Y ∪ Z, where
Y and Z are affine algebraic subsets of X. Then
I (X) = I (Y ) ∩ I (Z) ,
so I (X) contains the product I (Y ) I (Z). Since I (X) is prime, either I (X) contains I (Y ), in which case
X = Y , or I (X) contains I (Z), in which case X = Z.
63
M3P8 Algebra III 13 Introduction to algebraic geometry
Definition 13.2.12. An affine algebraic set X has Krull dimension d if d is the length of the largest
increasing tower of irreducible affine algebraic subsets
X0 ( · · · ( Xd ⊆ X.
Example.
• Points have dimension zero.
• In A1 , affine algebraic sets are zeros of polynomials, so irreducibles in A1 are points on A1 , so A1 has
Krull dimension one.
• In fact, An has Krull dimension n for every n.
Definition 13.2.13. A ring R has Krull dimension d if d is the length of the largest increasing tower
p0 ( · · · ( pd
of prime ideals of R.
Example. If R is a domain,
• dimension zero implies that R is a field, and
• dimension one implies that every nonzero prime ideal is maximal.
Lecture 30
Lecture 30 is a problem class. Wednesday
The Hilbert basis theorem then gives us the following. 12/12/18
Proposition 13.2.14. Let X be an affine algebraic set. Then X can be written uniquely as a finite union
X1 ∪ · · · ∪ Xr such that each Xi is an irreducible affine algebraic set and no Xi is contained in Xj for i 6= j.
Proof. We first show that if X is not irreducible, then X can be written as Y ∪ Z with Y irreducible
and Z 6= X affine algebraic. Certainly we can write X = Y1 ∪ Z1 with Y1 , Z1 proper subsets of X. If
Y1 is irreducible we are done. Otherwise write Y1 = Y2 ∪ Z2 . Again, if Y2 is irreducible we can write
X = Y2 ∪ (Z1 ∪ Z2 ) and we are done. Otherwise, supposing this never terminates, we obtain
Y1 ) Y2 ) . . . , I (Y1 ) ( I (Y2 ) ( . . . ,
which is impossible since K [X1 , . . . , Xn ] is Noetherian. Now given X, if X is not irreducible we can write
X = Y1 ∪ Z1 with Y1 irreducible and Z1 6= X. If Z1 is not irreducible we write Z1 = Y2 ∪ Z2 with Y2
irreducible and Z2 6= Z1 . If this process ever terminates we have written X as a finite union of irreducibles.
Otherwise, we have
Z1 ) Z2 ) . . . ,
and as above this is impossible since K [X1 , . . . , Xn ] is Noetherian. For uniqueness, suppose we have
X = Y1 ∪ · · · ∪ Yr , X = Z1 ∪ · · · ∪ Zs ,
with the Yi and Zj irreducible, and with no Yi , or Zi , contained in Yj , or Zj , when i 6= j. Then I (Yi ) and
I (Zi ) are prime for all i. In particular I (Y1 ) is prime. Since Y1 ⊂ X, I (X) ⊂ I (Y1 ), so I (Y1 ) contains
I (Z1 ∪ · · · ∪ Zs ) = I (Z1 ) ∩ · · · ∩ I (Zs ) .
It follows that I (Y1 ) contains the product I (Z1 ) . . . I (Zs ). Thus I (Y1 ) contains I (Zj ) for some j. Similarly
I (Zj ) contains I (Yi ) for some i. Then I (Y1 ) ⊆ I (Yi ), so Yi ⊆ Y1 and we must have i = 1. Then Y1 = Zj .
Proceeding we show that each Yi is equal to some Zj and vice versa, proving uniqueness.
Translating this to a statement about ideals, we find the following.
Corollary 13.2.15. Every radical ideal in K [X1 , . . . , Xn ] is uniquely expressible as a finite intersection of
prime ideals, none of which contains any of the others.
This is a special case of a very general ring-theoretic phenomenon known as primary decomposition,
which was discovered via the sort of geometric considerations we see above.
64
M3P8 Algebra III 13 Introduction to algebraic geometry
1 − T P (X1 , . . . , Xn ) .
In particular if (x1 , . . . , xn , t) lies in Z Je , then (x1 , . . . , xn ) lies in Z (J). Since P ∈ I (Z (J)) we have
P (x1 , . . . , xn ) = 0, so
1 − tP (x1 , . . . , xn ) = 1.
Thus Z Je is empty. By the weak Nullstellensatz, Je is the unit ideal, so there are polynomials Q0 , . . . , Qs
in K [X1 , . . . , Xn , T ], and R1 , . . . , Rs ∈ I, such that
1 = Q0 (1 − T P ) + Q1 R1 + · · · + Qs Rs .
It remains to prove the weak Nullstellensatz. This requires some new ideas. The following approach is due
to Emmy Noether.
65
M3P8 Algebra III 13 Introduction to algebraic geometry
Definition 13.3.2. Let R be a K-algebra, that is a ring together with a map K → R. We say that elements
y1 , . . . , ys of R are algebraically independent over K if there is no nonzero polynomial P (X1 , . . . , Xs ) ∈
K [X1 , . . . , Xs ] such that P (y1 , . . . , ys ) = 0. Equivalently, y1 , . . . , ys are algebraically independent if and
only if the map
K [X1 , . . . , Xs ] −→ R
Xi 7−→ yi
is injective.
Proposition 13.3.3 (Noether’s normalisation lemma). Let K be a field, and let R be a finitely generated
K-algebra. Then there exists s ∈ Z≥0 , and algebraically independent elements y1 , . . . , ys of R such that R is
integral over K [y1 , . . . , ys ].
Proof. Write
K [X1 , . . . , Xm ]
R= .
I
We proceed by induction on m. The base case m = 0 is clear. Fix m and assume the claim is true for
m − 1. If I = 0 then the statement is also clear, with yi = Xi for all i. Otherwise let P (X1 , . . . , Xm ) ∈ I.
Renumbering the variables if necessary, we may assume P is a nonconstant polynomial in Xm with coefficients
in X1 , . . . , Xm−1 . Let d be the total degree of P . That is, the largest value of a1 + · · · + am for any monomial
i
cX1a1 . . . Xm am
appearing in P . Let ni = (1 + d) and Yi = Xi − Xm ni
for each i in 1, . . . , m − 1. Define
n1 nm−1
Q (X1 , . . . , Xm ) = P (X1 + Xm , . . . , Xm−1 + Xm , Xm ) .
Then Q (Y1 , . . . , Ym−1 , Xm ) is zero in R. We now claim that, up to a factor c ∈ K ∗ , Q (X1 , . . . , Xm ) is monic
when considered as a polynomial in Xm . Let cX1a1 . . . Xm am
be a monomial appearing in P (X1 , . . . , Xm ).
This monomial contributes the following terms to Q (X1 , . . . , Xm ).
n1 a1 nm−1 am−1 am
c (X1 − Xm ) . . . (Xm−1 − Xm ) Xm .
Moreover, each ni is greater than d and hence greater than the sum of the aj . It is thus clear that the term
N
of highest degree in the above expression is cXm , where
m−1
N = n1 a1 + · · · + nm−1 am−1 + am = a1 (1 + d) + · · · + am−1 (1 + d) + am .
Since 1 + d is greater than the sum of the exponents a appearing in any monomial of P (X1 , . . . , Xm ), the
N
terms cXm appearing in different monomials are all of different degree and thus cannot cancel. It follows
N
that the term of the form cXm of highest degree is the highest degree term in Q (X1 , . . . , Xm ), so that
(1/c) Q (X1 , . . . , Xm ) is monic in Xm . Write
N
1 X
n
Q (X1 , . . . , Xm ) = Hn (X1 , . . . , Xm−1 ) Xm .
c n=0
That is, Xm is integral over the subalgebra S = K [Y1 , . . . , Ym−1 ] of R. Since Xm generates R over S, it
follows that R is integral over S. On the other hand we have a map
K [Z1 , . . . , Zm−1 ] −→ S
.
Zi 7−→ Yi
Let J be the kernel. Then
K [Z1 , . . . , Zm−1 ]
S= .
J
Then by the inductive hypothesis there are algebraically independent elements y1 , . . . , ys ∈ S such that S is
integral over K [y1 , . . . , ys ]. Since R is integral over S, it follows that R is integral over K [y1 , . . . , ys ] and we
are done.
66
M3P8 Algebra III 13 Introduction to algebraic geometry
hX1 − p1 , . . . , Xn − pn i , p1 , . . . , pn ∈ K.
Proof. Let I be a maximal ideal of K [X1 , . . . , Xn ], and consider R = K [X1 , . . . , Xn ] /I. Then R is a field.
On the other hand, by Noether normalisation, there exist y1 , . . . , ys algebraically independent such that R
is integral over S = K [y1 , . . . , ys ]. Let x be a nonzero element of K [y1 , . . . , ys ]. Then x−1
lies in R. Since R
is integral over S there is a monic polynomial P with coefficients in S such that P x−1 = 0. We thus have
d−1
d X
x−1 = ai x−i ,
i=0
so that x−1 is also in S. Thus S is a field. But since the yi are algebraically independent, S is also a
polynomial ring in s variables. Since no such ring is a field unless s = 0 we must have s = 0 and R is
integral over K. But then R is a finite dimensional K-vector space, hence a finite extension of K. Since K
is algebraically closed, the inclusion of K in R is an isomorphism. Thus for each i there is an element pi of
K such that Xi is equal to pi in R. Then Xi − pi is in I for all i, so I contains the ideal
hX1 − p1 , . . . , Xn − pn i .
hX1 − p1 , . . . , Xn − pn i .
67