Lectures On Communicative Algebra
Lectures On Communicative Algebra
Sudhir R. Ghorpade
Indian Institute of Technology Bombay
2
Contents
2 Noetherian Rings 18
2.1 Noetherian Rings and Modules . . . . . . . . . . . . . . . . . 18
2.2 Primary Decomposition of Ideals . . . . . . . . . . . . . . . . 20
2.3 Artinian Rings and Modules . . . . . . . . . . . . . . . . . . . 24
2.4 Krull’s Principal Ideal Theorem . . . . . . . . . . . . . . . . . 28
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3 Integral Extensions 33
3.1 Integral Extensions . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2 Noether Normalization . . . . . . . . . . . . . . . . . . . . . . 36
3.3 Finiteness of Integral Closure . . . . . . . . . . . . . . . . . . 39
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4 Dedekind Domains 45
4.1 Dedekind Domains . . . . . . . . . . . . . . . . . . . . . . . . 46
4.2 Extensions of Primes . . . . . . . . . . . . . . . . . . . . . . . 51
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
References 63
3
Chapter 1
In this chapter, we shall review a number of basic notions and results con-
cerning rings and modules. First, let us settle the basic terminology and
notation that we shall use throughout these notes.
By a ring we mean a commutative ring with identity. Given a ring A,
we denote by A∗ the set of all nonzero elements of A and by A× the set of
all (multiplicative) units of A. For sets I, J, we write I ⊆ J to denote that
I is a subset of J and I ⊂ J to denote that I is a proper subset of J, that
is, I ⊆ J and I 6= J. We denote the set of nonnegative integers by N and
for any n ∈ N, by Nn we denote the set of all n-tuples of elements of N. We
sometimes use the abbreviation ‘iff’ to mean ‘if and only if’.
4
by definition, is the set of all a ∈ A such that φ(a) = 0) is an ideal of A;
moreover, if I = ker φ denotes the kernel of φ, then A/I is isomorphic to
the image of φ. In short, residue class ring and homomorphic image are
identical notions. If I is an ideal of a ring A, then there is a one-to-one
correspondence between the ideals of A containing I and the ideals of A/I
given by J 7→ q(J) = J/I and J ′ 7→ q −1 (J ′ ).
An easy way to generate examples of ideals is to look at ideals gen-
erated by a bunch of elements of the ring. Given a ring A and elements
a1 , . . . , an ∈ A, the set
(a1 , . . . , an ) := { a1 x1 + · · · + an xn : x1 , . . . , xn ∈ A}
5
the set of all maximal ideals of A is denoted by Max (A). It is easy to see that
I is a prime ideal if and only if A/I is an integral domain, and also that I is
a maximal ideal if and only if A/I is a field. Using this (or alternatively, by
a simple direct argument), we see that every maximal ideal is prime, that
is, Max (A) ⊆ Spec (A).
Examples 1.1. (i) If A is the zero ring, then Spec (A) = ∅ = Max (A).
Remark 1.3. An easy alteration of the above proof shows that Proposition
1.2 holds under the weaker hypothesis that I is a subset of A closed under
addition and multiplication, and P1 , . . . , Pn are ideals of A such that at least
n − 2 of them are prime. If A contains a field, then Proposition 1.2 can be
proved, by elementary vector space arguments, without assuming any of
the Pi ’s to be prime.
(i) I1 I2 . . . In = I1 ∩ I2 ∩ · · · ∩ In .
6
(iii) The map x + I1 I2 · · · In 7→ (x + I1 , . . . , x + In ) defines an isomorphism of
A/I1 I2 . . . In onto the direct sum A/I1 ⊕ A/I2 ⊕ · · · ⊕ A/In .
7
I is any ideal in a ring A, then A \ I is multiplicatively closed if and
only if I is a prime ideal.
Corollary 1.8. Let I be a nonunit ideal of a ring A. Then there is a maximal ideal
m of A such that I ⊆ m. In particular, every nonzero ring has a maximal ideal and
the spectrum of a nonzero ring is nonempty.
Proof. The first assertion follows from Lemma 1.7 with S = {1}. To prove
the second assertion, take I = (0).
1
= 1 + a + · · · + an−1 ,
1−a
which shows that 1 − a is a unit in A. In fact, a similar argument shows
that if a ∈ Apis nilpotent, then 1 − ab is a unit in A for any b ∈ A. It
follows that (0) ⊆ {a ∈ A : 1 − ab is a unit for every b ∈ A}. The set
{a ∈ A : 1 − ab is a unit for every b ∈ A} will be denoted by J (A) and
called the Jacobson radical of A. The following result shows that the Jacobson
8
radical of A is an an ideal of A and it is, in fact, the intersection of all the
maximal ideals of A. It also gives an alternative proof of the fact that the
nilradical is contained in the Jacobson radical.
9
generated by homogeneous polynomials (resp: monomials). Henceforth,
when we use a notation such as k[X1 , . . . , Xn ], it will be tacitly assumed
that k denotes a field and X1 , . . . , Xn are independent indeterminates over
k (and, of course, n ∈ N).
The process of localization described next generalizes the construction
of the field of fractions of an integral domain, which in turn, is a general-
ization of the formal construction of rational numbers from integers.
Localization: Let A be a ring and S be a multiplicatively closed subset
of A. Define a relation ∼ on A×S as follows. Given any (a, s), (b, t) ∈ A×S,
10
The map φ is not injective, in general, and its kernel is given by
[
ker φ = {a ∈ A : sa = 0 for some s ∈ S} = (0 : s).
s∈S
(iv) The prime ideals of S −1 A are in one-to-one correspondence with the prime
ideals of A which do not meet S.
11
that (x/s)(y/t) ∈ S −1 p, then uxy ∈ p for some u ∈ S. Since p is prime and
p ∩ S = ∅, it follows that x ∈ p or y ∈ p, which implies that x/s ∈ S −1 p or
y/t ∈ S −1 p. Thus S −1 p is a prime ideal of S −1 A. So, in view of (i) and (iii)
above, it follows that the processes of contraction and extension set up the
desired one-to-one correspondence.
1.3 Modules
Let A be a ring. An A-module is simply a vector space except that the
scalars come from the ring A instead of a field. Apart from vector spaces
over a field, basic examples of A-modules are: ideals I of A, residue class
rings A/I, polynomial rings A[X1 , . . . , Xn ] and localizations S −1 A. The
notions of submodules, quotient modules, direct sums of modules and iso-
morphism of modules are defined in an obvious fashion. The concept of
localization (w.r.t. multiplicatively closed subsets of A) also carries over
to A-modules. A direct sum of (isomorphic) copies of A is called a free A-
module. The finite direct sum
An := A
| ⊕ ·{z
· · ⊕ A}
n times
(M1 : M2 ) := {a ∈ A : aM2 ⊆ M1 }.
12
In general, and unlike in the case of vector spaces, if an A-module M is
finitely generated, then a submodule of M need not be finitely generated.
For example, if A = C[X1 , X2 , . . . ] is the polynomial ring in infinitely many
variables and I = (X1 , X2 , . . . ) is the ideal of A generated by all the vari-
ables, then A is finitely generated as an A-module, but the A-submodule I
of A is not finitely generated. We shall see in Chapter 2 that if every ideal
of A is finitely generated, then every submodule of a finitely generated
A-module is finitely generated. Meanwhile let us note here a very useful
property of finitely generated A-modules.
Lemma 1.13 (Nakayama’s Lemma). Let M be a finitely generated A-module
and I be an ideal of A such that IM = M . Then (1 − a)M = 0 for some a ∈ I.
In particular, if I 6= A and if A is a local ring, then M = 0.
P
Proof. Write M = Ax1 + · · · + Axn . Then xi = nj=1 aij xj , for some aij ∈ I.
Let d = det(δij − aij ). Then d = 1 − a, for some a ∈ I, and, by Cramer’s
rule, dxj = 0 for all j.
13
V ⊆ Spec (A), we define
\
I(V ) := {f ∈ A : f ∈ P for all P ∈ V } = P.
P ∈V
Clearly, I(V ) is an ideal of A (in fact, a radical ideal); it is called the vanishing
ideal of V . For example, if A = C[X] and a is a nonzero ideal of A, then a =
(h(X)) for some nonzero polynomial h(X) ∈ C[X]. Further, if α1 , . . . , αm
are the distinct roots of h(X) in C, then V = V (a) = {α1 , . . . , αm } and
I(V ) = {f ∈ C[X] : f (αi ) = 0 for i = 1, . . . m} is the ideal of polynomials
which vanish at the roots of h(X).
The following result may be viewed as a version of Hilbert’s Nullstel-
lensatz (see Corollary 3.19 in Chapter 3). However, here the proof is trivial.
We shall use here the following notation and terminology.
Proposition 1.14. Let A be a ring.
(i) If a is a nonunit ideal of A, then V (a) is nonempty.
√
(ii) If a is an ideal of A and V = V (a) then I(V ) = a.
Proof. Corollary 1.8 implies (i) and Corollary 1.9 implies (ii).
form a base for the the Zariski topology on Spec (A). The Zariski topology
is almost never Hausdorff (or T2 ). In fact, as part (ii) of the proposition
below shows, a singleton set can be dense. On the other hand, part (iv) of
the same proposition shows that Spec (A) with the Zariski topology is a T0
topological space.
Proposition 1.15. Let A be a ring and p, q ∈ Spec (A). We have the following:
(i) p ∈ Max (A) ⇐⇒ {p} is closed subset of Spec (A).
14
ideal a of A. Consequently a ⊆ p, and hence V (p) ⊆ V (a). It follows that
V (p) is the smallest closed set containing {p}, and thus {p} = V (p). In case
A is a domain, (0) is a prime ideal and we have V ({(0)}) = Spec (A), that
is, {(0)} = Spec (A).
(iii) If p 6= q, then p * q or q * p. Suppose p * q. Then there is f ∈ p \ q;
now Df is an open set that contains q but does not contain p.
Proof. Given an open cover of Spec (A), we can refine it to an open cover
of Spec (A) consisting of principal open sets. This implies that Spec (A) is
the union of Df where f vary over a subset Λ of A. Consequently, the
intersection of V (f ) as f varies over Λ is empty, and so V (a) = ∅, where
a is the ideal of A generated by Λ. By part (ii) of Proposition 1.14, we see
that a = A, and hence there are f1 , . . . , fm ∈ Λ and g1 , . . . , gm ∈ A such that
f1 g1 + · · · + fm gm = 1. This implies that {Dfi : i = 1, . . . , m} cover Spec (A),
and hence the given open cover of Spec (A) has a finite subcover.
Exercises
Throughout the following exercises A denotes a ring.
15
4. Consider ideals a, b, c of A and the following three equalities.
Give examples to show that these results do not hold (for finite fami-
lies) if intersections are replaced by products.
8. Suppose A is not the zero ring and let N be the nilradical of A. Show
that the following are equivalent.
16
12. Given any f ∈ A, let S = {f n : n ∈ N} and Af = S −1 A. Show that Af
is isomorphic to A[X]/(Xf − 1).
13. Let S and T be multiplicatively closed subsets of A with S ⊆ T and
let U denote the image of T under the natural map φ : A → S −1 A.
Show that T −1 A is isomorphic to U −1 (S −1 A).
14. Show that localization commutes with taking homomorphic images.
More precisely, if I is an ideal of a ring A and S is a multiplica-
tively closed subset of A, then show that S −1 A/S −1 I is isomorphic
−1
to S (A/I), where S denotes the image of S in A/I.
15. Let A be an integral domain. Fix a quotient field K of A and consider
the localization Ap, where p ∈ Spec (A), as subrings of K. Show that
\ \
A = Ap = Am.
p∈Spec (A) m∈Max (A)
16. Consider the following ring-theoretic properties that A can have: (i)
integral domain, (ii) field, (iii) PIR, (iv) PID, and (v) UFD. For each of
these, determine if the property is preserved under the passage from
A to a (i) residue class ring, (ii) polynomial ring, or (iii) localization.
17. Let M be an A-module and S be a multiplicatively closed subset of
A. Define carefully the localization S −1 M of M at S. With ideals re-
placed by A-submodules, determine which of the notions and results
in Section 1.2 has an analogoue in the setting of modules.
18. Let (A, m) be a local ring [which means that A is a local ring and m
is its unique maximal ideal] and M be a finitely generated A-module.
For x ∈ M , let x denotes the image of x in the A/m-module M/mM .
Given any x1 , . . . , xr ∈ M , show that {x1 , . . . , xr } is a minimal set
of generators of M if and only if {x1 , . . . , xr } is a basis for the A/m-
vector space M/mM . Deduce that any two minimal set of generators
of M have the same cardinality, namely, dimA/m M/mM .
19. Assume that A is not the zero ring and let m, n ∈ N. Use Exercise 18
to show that Am and An are isomorphic as A-modules iff m = n.
20. Given any f, g ∈ A, show that the principal open sets Df and Dg of
Spec A satisfy the following.
(i) Df = ∅ ⇐⇒ f is nilpotent, (ii) Df = Spec (A) p
⇐⇒ f ispa unit,
(iii) Df ∩ Dg = Df g , and (iv) Df = Dg ⇐⇒ (f ) = (g).
21. Given any f ∈ A, show that the principal open set Df is quasi-
compact. Further show that an open subset of Spec (A) is quasi-
compact if and only if it is a finite union of principal open sets.
17
Chapter 2
Noetherian Rings
In this chapter we shall study a class of rings and modules named after E.
Noether (1921), who first realized their importance. Basic results proved
in this chapter include the Basis Theorem that goes back to the work of
Hilbert (1890) on C[X1 , . . . , Xn ] and the Primary Decomposition Theorem
for ideals that goes back to the work of Lasker (1905) also on C[X1 , . . . , Xn ].
18
A ring which satisfies either (and hence all) of the three equivalent con-
ditions in Proposition 2.1 is said to be noetherian. The class of noetherian
rings has a remarkable property that it is closed w.r.t. each of the three fun-
damental processes mentioned in Section 1.2. Indeed, if A is a noetherian
ring, then it is trivial to check that A/I is noetherian for every ideal I of A
and also that S −1 A is noetherian, for every multiplicatively closed subset
S of A. Moreover, the following basic result implies, using induction, that
if A is noetherian, then A[X1 , . . . , Xn ] is noetherian.
Proposition 2.2 (Hilbert Basis Theorem). If A is noetherian, then so is A[X].
Proof. Let I be any ideal of A[X]. For 0 6= f ∈ I, let LC(f ) denote the
leading coefficient of f , and let J := {0} ∪ {LC(f ) : f ∈ I, f 6= 0}. Then
J is an ideal of A, and so we can find f1 , . . . , fr ∈ I \ {0} such that J =
(LC(f1 ), . . . , LC(fr )). Let d = max{deg fi : 1 ≤ i ≤ r}. For 0 ≤ i < d, let
Ji = {0} ∪ {LC(f ) : f ∈ I, deg f = i}; then Ji is an ideal of A, and so we
can find fi1 , . . . , firi ∈ I \ {0} such that Ji = (LC(fi1 ), . . . , LC(firi )). Let I ′
be the ideal of A[X] generated by {f1 , . . . , fr }∪{fij : 0 ≤ i < d, 1 ≤ j ≤ ri }.
Clearly, I ′ ⊆ I and for any 0 6= f ∈ I, we easily see that there is f ′ ∈ I ′ such
that deg(f − f ′ ) < deg f . Thus an inductive argument yields I = I ′ .
19
(i) (Finite Generation Condition) Every submodule of M is finitely generated.
Proof. The proof of Proposition 2.1 carries over verbatim with ideals re-
placed by A-submodules.
20
Definition 2.8. An ideal q in a ring A is said to be primary if q 6= A and for
any a, b ∈ A,
/ q =⇒ an ∈ q for some n ≥ 1.
ab ∈ q and b ∈
√
If q is a primary ideal, then its radical q is readily seen to be a prime
√
ideal; if p = q, then we say that q is p–primary or that q is a primary ideal
belonging to p or that q is primary to p.
√
Remark 2.9. If q is an ideal such that q is prime, then q needn’t be pri-
mary; in fact, even a power of a prime ideal can fail to be primary. [Exam-
ple: p2 , where p is the image in k[X, Y, Z]/(XY − Z 2 ) of (X, Z).] However,
√
if q is a maximal ideal m, then √q is easily seen to be m–primary. [Indeed,
√
if ab ∈ q and a 6∈ m = q, then q + Aa ⊆ m + Aa = A, and so 1 ∈ q + Aa,
which implies b ∈ q.] On the other hand, if q is p–primary, then q needn’t be
a power of p, even when p is maximal. [Example: q = (X 2 , Y ) in k[X, Y ].]
It may be noted, however, that if A is a noetherian ring and q is a p–primary
ideal of A, then q does contain some power of p.
Proof: Classical proof of (i) is in two steps. First, one considers irreducible
ideals, viz., nonunit ideals that are not finite intersections of strictly larger
ideals. The maximality condition readily implies that every ideal of A is a
finite intersection of irreducible ideals. Next, if I is irreducible and ab ∈ I
are such that b ∈ / I and no power of a is in I, then we consider the chain
(I : a) ⊆ (I : a ) ⊆ . . . . By a.c.c., (I : an ) = (I : an+1 ) for some n, and now
2
21
have yci ∈ I, that is, y ∈ (I : ci ), which is a contradiction. This proves the
claim. Conversely, suppose (I : x) is a prime ideal p for some x ∈ A. Then
p = (∩qi : x) = ∩(q p i : x), and thus (qi : x) ⊆ p for some i. Hence x ∈ / qi
√
and pi = qi ⊆ (qi : x) ⊆ p. Also, a ∈ p ⇒ ax ∈ I ⊆ qi ⇒ a ∈ pi . Thus
p = pi . Finally, the uniqueness of the primary component qi corresponding
to a minimal prime pi can be proved by localizing at pi . 2
22
x = z 2 /y is in p2 Ap ∩ A. In fact, it can be seen that p(2) = (x, z 2 ) and
that p2 = (x, z 2 ) ∩ (x2 , y, z) is a primary decomposition of p2 in A.
The following result is a nice application of primary decomposition
(Proposition 2.10) and Nakayama’s Lemma (Proposition 1.13).
Proposition 2.13 (Krull’s Intersection Theorem). Suppose A is noetherian
and I is any ideal of A. Then there exists a ∈ I such that (1 − a) ∩∞ n
n=0 I = 0. In
∞ n
particular, if I 6= A, and A is a local ring, then ∩n=0 I = 0.
Proof. Let J = ∩∞ n
n=0 I . Write IJ = q1 ∩· · ·∩qr ∩qr+1 ∩· · ·∩qh , where qi are pi -
primary ideals with pi ⊇ I for 1 ≤ i ≤ r and pj 6⊇ I for r < j ≤ h. Fix some
yj ∈ I\pj for r < j ≤ h. Then x ∈ J ⇒ xyj ∈ IJ ⇒ xyj ∈ qj ⇒ x ∈ qj . Thus
J ⊆ qr+1 ∩ · · · ∩ qh . Also, since I ⊆ pi for 1 ≤ i ≤ r, there exists m ≥ 1 such
that I m ⊆ q1 ∩ · · · ∩ qr . Since J ⊆ I m , it follows that J ⊆ q1 ∩ · · · ∩ qh = IJ.
Thus IJ = J. Now apply Nakayama’s Lemma.
It may be remarked that Krull’s Intersection Theorem is usually proved
using a result known as the Artin-Rees Lemma. Importance of Krull Inter-
section Theorem stems from the fact that it paves the way for a Hausdorff
topology on a noetherian local ring. See [AM, Ch. 10] for more on this.
We end this section by introducing the notions, ascribed to Krull (1928),
of the dimension of a ring and the height of a prime ideal.
Definition 2.14. Let A be a ring. The (Krull) dimension of A is defined as the
maximum of the lengths of chains of prime ideals of A, that is,
dim A := max {n : ∃ p0 , p1 , . . . , pn ∈ Spec A such that p0 ⊂ p1 ⊂ · · · ⊂ pn } ;
in case A has no prime ideals (which happens only when A is the zero ring),
we set dim A = −1, and in case the set of lengths of chains of prime ideals
of A is unbounded, we set dim A = ∞. Given any prime ideal p of A, the
height of p is defined by ht p = dim Ap. Equivalently, ht p is the maximum
of the lengths of chains p0 ⊂ p1 ⊂ · · · ⊂ pn of prime ideals of A with pn = p.
If A is a noetherian ring, the concept of height can be extended to any
nonunit ideal I of A by putting ht I = min{ht p : p ∈ Ass(A/I)}. Note that
ht I = min{ht p : p ∈ Min(A/I)} = min{ht p : p ∈ V (I)}.
Examples 2.15. (i) A field has dimension 0. A PID which is not a field
(such as, for example, Z or k[X]) has dimension 1.
(ii) If R is a local ring and I is primary for the unique maximal ideal of
R, then dim(R/I) = 0. More generally, if I is any nonunit ideal of a
ring A and p is a minimal prime of I in A, then dim(Ap/IAp) = 0.
(iii) In the polynomial ring k[X1 , . . . , Xn ] we have a chain of prime ideals
of length n given by (0) ⊂ (X1 ) ⊂ (X1 , X2 ) ⊂ · · · ⊂ (X1 , . . . , Xn );
consequently, dim k[X1 , . . . , Xn ] ≥ n. Similarly, ht (X1 , . . . , Xr ) ≥ r.
23
2.3 Artinian Rings and Modules
Throughout this section A denotes a ring and M denotes an A-module. By
a submodule of M , we mean an A-submodule of M . Here is a result similar
to Proposition 2.1 with descending chains instead of ascending chains.
Proof. The proof of “(i) ⇒ (ii)” is identical to that of “(ii) ⇒ (iii)” in Propo-
sition 2.1 with ideals replaced by submodules and “⊂” replaced by “⊃”.
Conversely, if N1 ⊇ N2 ⊇ . . . is a descending chain of submodules of M
and Nm is a minimal element of {Ni : i ≥ 1}, then Nn = Nm for n ≥ m.
Examples (i) and (iii) above show that an analogue of Hilbert Basis The-
orem is not true for artinian rings. However, it is easy to see that if A is
artinian, then so is A/I as well as S −1 A for any ideal I of A and a mul-
tiplicatively closed subset S of A. In the case of modules, we have the
following analogues of the results in Section 2.1 for noetherian modules.
24
(ii) Mi is artinian for i = 1, . . . , n =⇒ ⊕ni=1 Mi is artinian.
Proof. (i), (ii) and (iii) follow from arguments similar to those in the proofs
of Proposition 2.5, Corollary 2.6 and Proposition 2.7, respectively.
For a finite dimensional vector space over a field, there are several ways
of defining the dimension. For example, as the cardinality of (i) a basis or
(ii) a minimal generating set or (iii) a maximal linearly independent set.
Yet another way is to define the dimension as the length of maximal chains
of subspaces of V . Indeed, if dim V = n, then there is a maximal chain
V = V0 ⊃ V1 ⊃ · · · ⊃ Vn of subspaces and its length (= the number of
inclusion signs) is n. Moreover, every maximal chain looks like this. It turns
out that the notion of length extends to a slightly more general setting.
In general, a chain of submodules of M is a strictly descending sequence
M = M0 ⊃ M1 ⊃ · · · ⊃ Mn = 0 of submodules of M ; the length of such a
chain is n. A maximal chain, i.e., a chain in which no extra submodules can
be added, is called a composition series of M . Note that a chain M = M0 ⊃
M1 ⊃ · · · ⊃ Mn = 0 of submodules of M is a composition series of M if and
only if Mi−1 /Mi is simple (that is, it has no submodules other than 0 and
itself) for each i = 1, . . . , n. We define the length of M , denoted by ℓA (M )
or simply by ℓ(M ), to be the minimum of the lengths of composition series
of M . In case M has no composition series, we set ℓ(M ) = ∞.
Proposition 2.19. Assume that M has finite length. Then we have the following.
(i) Every submodule N of M has finite length. Also, N ⊂ M ⇒ ℓ(N ) < ℓ(M ).
25
(iii) If M = M0′ ⊃ M1′ ⊃ · · · ⊃ Mr′ = 0 is a composition series of M ,
then r ≤ n, by (ii). Also, r ≤ n by the definition of length. Thus r = ℓ(M ).
Conversely, if r = ℓ(M ) and the chain is not a composition series, then we
can insert some terms to get a chain of length > ℓ(M ).This contradicts (ii).
(iv) In view of (iii), if a chain is not a composition series, then new terms
can be added to it until its length becomes n, as desired.
26
We will now use the previous proposition to derive a useful and in-
teresting characterization of artinian rings. But first, let us observe some
simpler properties of artinian rings.
Proposition 2.24. Let A be an artinian ring. Then we have the following.
(i) A is a domain ⇐⇒ A is a field.
p
(ii) Spec (A) = Max (A). In particular, (0) = J (A).
(iii) Max (A) is finite.
p
(iv) If N = (0) is the nilradical of A, then Nm = (0) for some m ∈ N.
Proof. (i) Given any a ∈ A, applying the d.c.c. to the chain (a) ⊇ (a2 ) ⊇ . . . ,
we see that (an ) = (an+1 ) for some n ≥ 1. Hence an = ban+1 for some
b ∈ A. Now if A is a domain and a 6= 0, then an 6= 0 and hence ba = 1. Thus
an artinian domain is a field. The converse is trivial.
(ii) Given any p ∈ Spec p (A), apply (i) to A/p. This gives Spec (A) =
Max (A). Consequently, (0) = J (A), by Propositions 1.9 and 1.10.
(iii) If A is the zero ring, there is nothing to prove. Assume that A 6= 0.
Consider the set F of all finite intersections of maximal ideals of A. This
is nonempty and hence has a minimal element, say m1 ∩ · · · ∩ mr . Now
for any m ∈ Max (A), m ∩ m1 ∩ · · · ∩ mr ∈ F. Hence by the minimality
of m1 ∩ · · · ∩ mr , we see that m ⊇ m ∩ m1 ∩ · · · ∩ mr = m1 ∩ · · · ∩ mr , and
this implies m ⊇ mi for some i. Since mi is maximal, we obtain m = mi . It
follows that Max (A) = {m1 , . . . , mr }.
(iv) Applying d.c.c. to the chain N ⊇ N2 ⊇ . . . , we see that there is
m ≥ 1 with Nn = Nm for n ≥ m. Suppose, if possible, I := Nm is
nonzero. Then the set {J : J an ideal of A with IJ 6= (0)} is nonempty;
hence it has a minimal element, say J0 . Since IJ0 6= (0), there is a ∈ J0 such
that I(a) 6= (0). Hence J0 = (a), by the minimality of J0 . Further I 2 = I
and so I(I(a)) = I(a) 6= (0). Hence I(a) = (a), again by the minimality of
J0 = (a). It follows that a =pua for some u ∈ I, and consequently, a = uj a
for all j ≥ 1. But, u ∈ I ⊆ (0), and hence uj = 0 for some j ≥ 1. Thus,
a = 0, which is a contradiction. It follows that Nm = I = (0).
Proposition 2.25. A ring is artinian iff it is noetherian and zero dimensional.
Proof. Suppose A is an artinian ring. Then by part (ii) of Proposition 2.24,
dim A = 0. Further, by parts (iii) and (iv) of Proposition 2.24, we see that
(m1 m2 · · · mr )m = (m1 ∩ · · · ∩ mr )m = 0. for some m ≥ 1 and (finitely
many) maximal ideals m1 , . . . , mr of A. Hence by Proposition 2.23, A is
noetherian. Conversely, if A is noetherian and dim A = 0, then every
p prime
ideal of A is maximal, and hence in view of Proposition 2.10, (0) is an
intersection m1 ∩ · · · ∩ mr of maximal ideals of A. Since A is noetherian,
(m1 m2 · · · mr )m = (m1 ∩ · · · ∩ mr )m = 0 for some m ≥ 1, and hence by
Proposition 2.23, A is artinian.
27
2.4 Krull’s Principal Ideal Theorem
In this section we shall prove an important result in dimension theory of
commutative rings, due to Krull (1929), called the Principal Ideal Theorem
or the Hauptidealsatz. The proof given here is along classical lines and
will use the notion of symbolic powers introduced in Section 2.2. For a
more modern approach, using the so called Dimension Theorem, one may
consult [AM] or [M2].
Proof. By localizing at p, we can and will assume that A is a local ring with
p as its unique maximal ideal. Let p1 be any prime ideal of A such that
p1 ⊂ p. It suffices to show that ht p1 = 0. Since p is a minimal prime of (a)
and the unique maximal ideal of A, we see that dim A/(a) = 0 and a 6∈ p1 .
Consider
(2) (3)
(a) + p1 ⊇ (a) + p1 ⊇ (a) + p1 ⊇ . . . ,
(n)
where p1 denotes the nth symbolic power of p1 . This corresponds to a
descending chain of ideals of A/(a). By Proposition 2.25, A/(a) is artinian,
(n) (n+1)
and hence there is n ≥ 1 such that (a) + p1 = (a) + p1 . In particular,
(n) (n+1)
given any x ∈ p1 , we can write x = ab + y for some b ∈ A and y ∈ p1 .
(n) (n) (n)
Now ab ∈ p1 and since a 6∈ p1 and p1 is p1 -primary, we see that b ∈ p1 .
(n) (n) (n+1)
Thus, p1 = (a)p1 + p1 . Hence by Nakayama’s Lemma (applied to
(n) (n+1) (n) (n+1)
the module p1 /p1 ), we obtain p1 = p1 . This implies that pn1 Ap1 =
n+1
p1 Ap1 = (p1 Ap1 )(pn1 Ap1 ). Applying Nakayama’s Lemma once again (this
time to the Ap1 -module pn1 Ap1 ), we obtain pn1 Ap1 = 0. It follows that p1 Ap1
is the only prime ideal of Ap1 and dim Ap1 = 0, that is, ht p1 = 0.
Proof. By localizing at p, we can and will assume that A is a local ring with
p as its unique maximal ideal. This implies that (a1 , . . . , ar ) is p-primary.
We shall now proceed by induction on r. Proposition 2.26 settles the case
r = 1. Suppose r > 1 and the result holds for r − 1. Let p1 be any prime
ideal of A such that p1 ⊂ p and p1 is maximal among prime ideals of A
strictly contained in p. It suffices to show that ht p1 ≤ r − 1. To begin
with, note that not all ai (1 ≤ i ≤ r) can be in p1 . Assume, without loss of
generality, that a1 6∈ p1 . Now, p1 ⊂ p1 + (a1 ) ⊆ p, and hence p is a minimal
prime of p1 + (a1 ). But p is also the unique maximal ideal of A. Hence
28
p1 + (a1 ) is p-primary, and since A is noetherian, there is m ≥ 1 such that
pm ⊆ p1 + (a1 ). In particular, for i = 2, . . . , r we can write am i = yi + xi a1
for some yi ∈ p1 and xi ∈ A. Now (a1 , . . . , ar ) is p-primary and am i ∈
(a1 , y2 , . . . , yr ) for 1 ≤ i ≤ r. Hence pN ⊆ (a1 , y2 , . . . , yr ) for some large
enough N ≥ 1. Hence p is a minimal prime of (a1 , y2 , . . . , yr ), and therefore
p/(y2 , . . . , yr ) is a minimal prime of (a1 , y2 , . . . , yr )A/(y2 , . . . , yr ). But the
latter is a principal ideal of A/(y2 , . . . , yr ). Hence by Proposition 2.26, the
height of p/(y2 , . . . , yr ) is ≤ 1, and therefore, the height of p1 /(y2 , . . . , yr )
is ≤ 0. It follows that p1 is a minimal prime of (y2 , . . . , yr ), and so by the
induction hypothesis, ht p1 ≤ r − 1. This proves that ht p ≤ r.
Of these assertions, the first is obvious from Proposition 2.27, while the
second follows from the first. To see the third, recall from Example 2.15 (iii),
that ht (X1 , . . . , Xr ) ≥ r; the other inequality ht (X1 , . . . , Xr ) ≤ r follows
from Proposition 2.27. Finally, the fourth assertion that gives a character-
ization of UFD’s can be proved as follows. If A is a UFD and p a height
1 prime ideal of A, then there is a ∈ p such that a is irreducible (check!);
moreover, since A is a UFD, (a) is a nonzero prime ideal of A with (a) ⊆ p.
But since ht p = 1, we must have p = (a), that is, p is principal. Conversely,
suppose every height 1 prime ideal of A is principal. Since A is a noetherian
domain, it satisfies a.c.c. on principal ideals of A, and hence every nonzero
element of A can be factored as a product of irreducible elements. Thus it
suffices to show that an irreducible element of A is prime. Let a ∈ A be
irreducible. If p is a minimal prime of (a), then ht p ≤ 1. Moreover, since p
contains a 6= 0, it follows that ht p = 1. Hence by our hypothesis, p is princi-
pal, that is, p = (b) for some b ∈ A. Now, a is irreducible and (a) ⊆ (b) 6= A
implies (a) = (b) = p. Hence (a) is a prime ideal of A and so a is prime.
29
Let R be a local ring with m as its unique maximal ideal. By a system of
parameters for R we mean elements x1 , . . . , xd of R such that (x1 , . . . , xd ) is
m-primary, that is, mn ⊆ (x1 , . . . , xd ) for some n ∈ N. Define
s(R) := min{d ∈ N : ∃ a system of parameters for R with d elements}.
Then s(R) = dim R. Indeed, if (x1 , . . . , xd ) is m-primary, then m is a min-
imal prime of (x1 , . . . , xd ), and so by Proposition 2.27, dim R = ht m ≤ d.
Thus dim R ≤ s(R). On the other hand, if dim R = ht m = d, then by the
(above mentioned) converse of Krull’s Principal Ideal Theorem, there exist
x1 , . . . , xd ∈ R such that m is a minimal prime of (x1 , . . . , xd ). But since m
is the unique maximal ideal of R, it follows that (x1 , . . . , xd ) is m-primary.
This shows that s(R) ≤ d = dim R.
Exercises
Throughout the following exercises A denotes a ring.
1. Let A = k[X1 , X2 , . . . ] be the polynomial ring in infinitely many vari-
ables with coefficients in a field k. Prove that A is not noetherian.
√
2. Let q be an ideal of A and p = q. Show that if A is noetherian, then
pn ⊆ q for some n ∈ N. Is this result valid if A is not noetherian?
Justify your answer.
3. Let q be a nonunit ideal of A. Show that q is primary if and only if
every zerodivisor in A/q is nilpotent.
4. Let q be a p-primary ideal and x be an element of A. Show that if
x ∈ q, then (q : x) p
= (1), whereas if x 6∈ q, then (q : x) is p-primary,
and in particular, (q : x) = p. Further show that if x 6∈ p, then
(q : x) = q.
√
5. Show that if q is an ideal of A such that q ∈ Max (A), then q is
primary.
6. Let A = Z[X] and consider the ideals q = (4, X) and m = (2, X).
Show that m is a maximal ideal of A and q is m-primary, but q is not
a power of m.
7. Let A = k[X, Y ] and I = (X 2 , XY, Y 2 ). Show that I = (X 2 , Y ) ∩
(X, Y 2 ) is a primary decomposition of I. Is this an irredundant pri-
mary decomposition of I? Justify your answer.
8. Let I be a radical ideal and I = q1 ∩· · ·∩qh be an irredundant primary
decomposition of I, where qi is pi –primary for 1 ≤ i ≤ h. Show that
I = p1 ∩ · · · ∩ ph . Deduce that I has no embedded component, and
that qi = pi for 1 ≤ i ≤ h.
30
9. Given an ideal I of A, define Z(A/I) := {x ∈ A : (I : x) 6= I} ∪ {0}.
Show that Z(A/I) is the union of the associated primes of I, that is,
[ [
Z(A/I) = p and deduce that Z(A) = p,
p∈Ass(A/I) p∈Ass(A/(0))
13. Given an ideal I of A, let I[X] denote the set of all polynomials in
A[X] with coefficients in I. Show that IA[X] = I[X] and also that
14. Let ∆ be a simplicial complex with vertex set V = {1, 2, . . . , n}, and
let F1 , F2 , . . . , Fm be the facets (that is, maximal faces) of ∆. Let I∆
be the ideal of k[X1 , . . . , Xn ] generated by the monomials Xi1 . . . Xir
for which {i1 , . . . , ir } ∈/ ∆. Given any face F of ∆, let PF be the
ideal of k[X1 , . . . , Xn ] generated by the variables Xj1 , . . . , Xjs , where
{j1 , . . . , js } = V \ F . Prove that each PF is a prime ideal and that
I∆ = PF1 ∩ · · · ∩ PFm is an irredundant primary decomposition of I∆ .
31
15. Let J be a monomial ideal of k[X1 , . . . , Xn ] and let u, v be relatively
prime monomials in k[X1 , . . . , Xn ]. Show that (J, uv) = (J, u) ∩ (J, v).
Also show that if e1 , . . . , en are positive integers, then (X1e1 , . . . , Xnen )
is (X1 , . . . , Xn )–primary. Use these facts to determine the associated
primes and a primary decomposition of the ideal (X 2 Y Z, Y 2 Z, Y Z 3 )
of k[X, Y, Z].
17. Determine the dimension of the ring Z[X] of polynomials in one vari-
able with integer coefficients.
18. Suppose A is noetherian and I is any ideal of A. Show that dim A/I =
max{dim A/p : p ∈ Ass(A/I)} = max{dim A/p : p ∈ Min(A/I)}.
22. Show that every artinian ring is isomorphic to a direct sum of finitely
many artinian local rings. (Hint: Use Proposition 2.24 and the Chi-
nese Remainder Theorem.)
23. Give an example of a zero dimensional local ring that is not noethe-
rian.
32
Chapter 3
Integral Extensions
33
closed subset of A, B is integral over A, and J is an ideal of B, then S −1 B
(resp: B/J) is integral over S −1 A (resp: A/J ∩ A).
Example 3.2. Let B := k[X, Y ]/(Y − X 2 ), and let x, y denote the images of
X, Y in B so that B = k[x, y]. Let A = k[y]. Then x is integral over A, and
hence B is integral over A. On the other hand, if B = k[X, Y ]/(XY − 1) =
k[x, y], then x is not integral over A = k[y]. It may be instructive to note,
indirectly, that B ≃ k[Y, 1/Y ] is not a finite k[Y ]-module. These examples
correspond, roughly, to the fact that the projection of the parabola y = x2
along the x-axis onto the y-axis is a ‘finite’ map in the sense that the inverse
image of every point is at ‘finite distance’, whereas in the case of the hyper-
bola xy = 1, this isn’t so. Similar examples in “higher dimensions” can be
constructed by considering projections of surfaces onto planes, solids onto
3-space, and so on.
Basic results about integral extensions are as follows. In the seven re-
sults given below, B denotes an integral extension of A and p ∈ Spec (A).
A is a field ⇐⇒ B is a field.
p is maximal ⇐⇒ q is maximal.
Proposition 3.5 (Lying Over Theorem). There exists a prime ideal q of B such
that q ∩ A = p. In particular, pB ∩ A = p.
Proposition 3.8 (Going Down Theorem). Assume that A and B are domains
and A is normal. If q is a prime ideal of B such that q ∩ A = p, and p′ is a prime
ideal of A such that p′ ⊆ p, then there exists a prime ideal q′ of B such that q′ ⊆ q
and q′ ∩ A = p′ .
Corollary 3.9. Assume that A and B are domains and A is normal. Then for any
prime ideal q of B such that q ∩ A = p, we have ht p = ht q.
34
Sketch of Proofs of Propositions 3.3, 3.5, 3.6 and 3.8. Easy manipula-
tions with integral equations of relevant elements proves the first asser-
tion of Proposition 3.3; the second and third assertions follow from the first
one by passing to quotient rings and localizations respectively. To prove
Proposition 3.5, first localize at p and, then, note that if pB0 = B0 , where
B0 = (A \ p)−1 B, then pB ′ = B ′ for some f.g. A-algebra B ′ ; now B ′ is a
finite A-module and Nakayama’s Lemma applies. Proposition 3.6 follows
by applying Proposition 3.5 to appropriate quotient rings. To prove Propo-
sition 3.8, consider the multiplicatively closed subset S = (A \ p′ )(B \ q) =
{ab : a ∈ A \ p′ , b ∈ B \ q} of B and note that it suffices to prove p′ B ∩ S = ∅
[because, then there exists a prime ideal q′ of B such that p′ B ⊆ q′ and
q′ ∩ S = ∅, and this will have the desired properties]. To this end, let
x ∈ p′ B ∩ S. Let K and L denote the quotient fields of A and B respectively.
Let L be a normal extension of K containing L and B the integral closure of
A in L. Since p′ ⊆ A and x ∈ p′ B, all the conjugates of x w.r.t L/K are in p′ B.
Hence the coefficients of the minimal polynomial, say f (X), of x over K are
in p′ B ∩ A = p′ (since A is normal!). Write f (X) = X d + c1 X d−1 + · · · + cd ,
and x = ab, where c1 , . . . , cd ∈ p′ , a ∈ A \ p′ and b ∈ B \ q. Clearly,
X d + (c1 /a)X d−1 + · · · + (cd /ad ) is the minimal polynomial of b over K. But
A is normal and b is integral over A implies that ci = c′i ai for some c′i ∈ A
(1 ≤ i ≤ d). Since ci ∈ p′ and a 6∈ p′ , we have c′i ∈ p′ for 1 ≤ i ≤ d. Hence
bd ∈ p′ B ⊆ pB ⊆ q, and so b ∈ q, which is a contradiction. 2
For a more leisurely proof of the results above, see [AM, pp. 61–64] or
[ZS, pp. 257–264].
Remark 3.10. It may be noted that Corollary 3.7 is an analogue of the sim-
ple fact that if L/K is an algebraic extension of fields containing a common
subfield k, then tr.deg.k L = tr.deg.k K. Recall that if K is a ring contain-
ing a field k, then elements θ1 , . . . , θd of K are said to be algebraically in-
dependent over k if they do not satisfy any algebraic relation over k, i.e.,
f (θ1 , . . . , θd ) 6= 0 for any 0 6= f ∈ k[X1 , . . . , Xd ]. A subset of K is al-
gebraically independent if every finite collection of elements in it are alge-
braically independent. If K is a field then any two maximal algebraically
independent subsets have the same cardinality, called the transcendence de-
gree of K/k and denoted by tr.deg.k K; such subsets are then called transcen-
dence bases of K/k; note that an algebraically independent subset S is a tran-
scendence basis of K/k iff K is algebraic over k(S), the smallest subfield of
K containing k and S. If B is a domain containing k and K is its quotient
field, then one sets tr.deg.k B = tr.deg.k K. Finally, note that k[X1 , . . . , Xd ]
and its quotient field k(X1 , . . . , Xd ) are clearly of transcendence degree d
over k. On the other hand, if θ1 , . . . , θd are algebraically independent over k,
then k[θ1 , . . . , θd ] and k(θ1 , . . . , θd ) are k-isomorphic to the polynomial ring
k[X1 , . . . , Xd ] and the rational function field k(X1 , . . . , Xd ), respectively.
A good reference for this material is Chapter 2 of [ZS].
35
3.2 Noether Normalization
In this section we prove a basic result in Dimension Theory, known as
Noether’s Normalization Lemma. This result has a number of useful con-
sequences. For example, we will show that for a special class of rings, the
notions of Krull dimension and transcendence degree are closely related.
Also, we will prove a famous result known as Hilbert’s Nullstellensatz.
An essential ingredient in the proof of Noether’s Normalization Lemma
can be formulated and proved in the form of a lemma, which is of indepen-
dent interest and use. As an application, we shall compute the dimension
of k[X1 , . . . , Xn ] and the height of any prime ideal in k[X1 , . . . , Xn ] gener-
ated by a subset of the variables X1 , . . . , Xn .
The key idea in the proof of the following lemma may be explained
by revisiting the example of hyperbola xy = 1. We have noticed that the
projection from the hyperbola onto the y-axis is not ‘finite’. However, if we
tilt the axes a bit, e.g., via the coordinate change X ′ = X, Y ′ = Y − cX
for some c 6= 0, then the projection becomes a ‘finite’ map. A similar trick
works in general.
f = cX1m + g1 X1m−1 + · · · + gm
36
f, X2′ , . . . , Xn′ are algebraically independent over k, and A is isomorphic to
a polynomial ring in n variables over k. This implies that A is a normal do-
main and thus if f h ∈ A for some h ∈ B, then h ∈ A, being in the quotient
field of A and integral over A. Thus f B ∩ A = f A and consequently, B/f B
is integral over A/f A.
37
there exists r ≥ 1 such that J1 ∩ k[θ2 , . . . , θn ] = (θ2 , . . . , θr ). Hence A =
A′ [f ] is integral over k[f, θ2 , . . . , θn ], and therefore so is B. Consequently,
tr.deg.k k[f, θ2 , . . . , θn ] = tr.deg.k B = n, and so if we let θ1 = f , then
θ1 , . . . , θn are algebraically independent over k. Moreover, since f ∈ J1 .
and J1 ∩ k[θ1 , . . . , θn ] = (θ1 ) + J1 ∩ k[θ2 , . . . , θn ] = (θ1 , . . . , θr ).
If m > 1 and we assume that the result holds for m − 1, then for the
chain J1 ⊆ · · · ⊆ Jm−1 , there exist θ1′ , . . . , θn′ ∈ B and nonnegative integers
r1′ , . . . , rm−1
′ satisfying conditions such as (i), (ii) and (iii). Let r = rm−1 ′ .
Using the previous case (of m = 1), we can find θr+1 , . . . , θn in B ′′ = ′′ ′′
Corollary 3.15. If B is a domain and a f.g. algebra over a field k, and P is a prime
ideal of B, then dim B = ht P + dim B/P .
Proof. Apply Proposition 3.13 for the singleton chain P, and use Corollary
3.7 and Corollary 3.9.
38
(0) and q. By Noether’s Normalization Lemma, there is a corresponding
chain (0) ⊂ p0 ⊂ p1 ⊂ · · · pm of prime ideals of the polynomial ring A :=
k[θ1 , . . . , θd ] with qi ∩ A = pi and pi := (θ1 , . . . , θri ) for 0 ≤ i ≤ m and
some nonnegative integers r0 ≤ r1 ≤ · · · ≤ rm . Now, by Corollary 3.9 and
Corollary 3.12, we have ri+1 = ht pi+1 = ht qi+1 > ht qi = ht pi = ri for
0 ≤ i < m. Moreover, if ri+1 > ri + 1 for some i, then using the Going
Down Theorem (Proposition 3.8) and Corollary 3.9, we can find a prime
ideal q′ of B such that qi ⊂ q′ ⊂ qi+1 . This contradicts the maximality of
(0) = q0 ⊂ q1 ⊂ · · · qm = q. Thus, ri+1 = ri + 1 for 0 ≤ i < m. Further,
it is easily seen that r0 = 0, and so rm = m. It follows that ht q = m. This
completes the proof.
39
by Tr(α). It is easy to see that TrL/K is a K-linear map of L into K, and
Tr(a) = [L : K]a for all a ∈ K.
A finite field extension L/K is said to be algebraic if every α ∈ L satisfies
a nonzero polynomial in K[X]. Such a polynomial is unique if we require
it to be monic and irreducible; in that case it is called the minimal polynomial
of α over K and denoted by Irr(α, L/K). An algebraic extension L/K is
said to be separable if Irr(α, L/K) has distinct roots for every α ∈ L. Now
suppose L/K is finite separable of degree n := [L : K] and α ∈ L. The
degree, say d, of Irr(α, L/K) divides n, and the distinct roots of Irr(α, L/K),
each repeated n/d times, are called the conjugates of α w.r.t L/K. Thus,
α ∈ L has exactly n conjugates w.r.t. L/K. The sum of these n conjugates
is precisely the trace of α. Equivalently, TrL/K (α) is the sum of all σ(α)
as σ varies over all K-homomorphisms of L into an algebraic closure of K
containing L.
If L/K is a finite extension of degree n and {α1 , . . . , αn } is any K-basis
of L, then the determinant of the n × n matrix TrL/K (αi αj ) is called the
discriminant of L/K w.r.t. {α1 , . . . , αn }. The vanishing of this is indepen-
dent of the choice of a K-basis. In case L/K is separable, L has a K-basis
of the form {1, α, α2 , . . . , αn−1 } for some α ∈ L. With respect to this basis,
the discriminant is a Vandermonde determinant and hence nonzero. Con-
sequently, the bilinear form L × L → K given by (x, y) 7→ TrL/K (xy) is
nondegenerate.
Proof. (i) If α satisfies the monic polynomial X n +a1 X n−1 +· · ·+an ∈ K[X],
then we can find a common denominator c ∈ A such that c 6= 0 and ai =
ci /c for some ci ∈ A. Multiplying the above polynomial by cn , we get a
monic polynomial in A[X] satisfied by cα.
(ii) If α ∈ L is integral over A, then so is each of its conjugate. Hence
Tr(α) is integral over A. Since Tr(α) ∈ K and A is normal, it follows that
Tr(α) ∈ A.
We are now ready to prove the Finiteness Theorem in the separable case.
Proposition 3.21. Let A be a normal domain and K be its quotient field. Let L/K
be a finite separable extension of degree n and B be the integral closure of A in L.
40
Then B is contained in a free A-module generated by n elements. In particular, if
A is also noetherian, then B is a finite A-module and a noetherian ring.
Proof. Using part (i) of Lemma 3.20, we can find a K-basis {α1 , . . . , αn } of
L, which is contained in B. Let {β1 , . . . , βn } be a dual basis, w.r.t. the non-
degenerate bilinear form TrL/KP (xy), corresponding to {α1 , . . . , αn }. Given
any
P x ∈ B, we can write x = j bj βj for some bj ∈ K. Now Tr(αi x) =
j bj Tr(αi βj ) = bi . Moreover, since αi x is integral over A, it follows from
Lemma 3.20 that bi ∈ A. Thus B is contained in the A-module generated
by β1 , . . . , βn . This module is free since β1 , . . . , βn are linearly independent
over K.
Then I is an ideal of A and thus I = (a11 ) for some a11 ∈ A. Also, there exist
a12 , . . . , a1n ∈ A such that y1 ∈ N where y1 = a11 x1 + a12 x2 + · · · + a1n xn .
If n = 1, we have N = Ix1 = Ay1 , where y1 = a11 x1 and thus the result is
proved in this case. If n > 1, then let M1 = Ax2 + . . . Axn and N1 = N ∩ M1 .
By induction hypothesis, we can find P aij ∈ A for 2 ≤ i ≤ j ≤ n such that
N1 = Ay2 + · · · + Ayn where yi = j≥i aij xj for
X
N1 = Ay2 + · · · + Ayn where yi = aij xj for 2 ≤ i ≤ n.
j≥i
41
The above Corollary applied in the particular case of A = Z, shows that
the ring of integers of a number field always has a Z-basis. Such a basis is
called an integral basis of that ring or of the corresponding number field.
Now let us prove the general case of the Finiteness Theorem. Again it
may be useful to recall some pertinent facts from field theory.
A finite field extension L/K is said to be normal if Irr(α, K) has all its
roots in L for every α ∈ L. If L/K is a finite extension, then we can find
a normal extension N/K such that N contains L and no proper subfield
of N containing L is normal over K. Such an extension N is necessarily
finite over K and is unique upto a K-isomorphism; it is called the nor-
mal closure of L over K. Given any field extension L/K, the set of all K-
automorphisms of L is a group, called the Galois group of L/K and denoted
by Gal(L/K). From Galois theory, we know that if H is a finite subgroup
of Gal(L/K) and LH := {α ∈ L : σ(α) = α for all σ ∈ H} is the fixed
field of H in L, then L/LH is a finite, separable and normal extension with
Gal(L/LH ) = H. If L/K is a finite normal extension and H = Gal(L/K),
then every element of LH is purely inseparable over K, i.e, every α ∈ LH
satisfies αq ∈ K where q is some power of the characteristic of K. In other
words, LH /K is a purely inseparable extension.
Theorem 3.24. Let A be a domain and an affine k-algebra, and K be its quotient
field. Let L be a finite extension of K and B be the integral closure of A in L. Then
B is a finite A-module, and hence an affine k-algebra.
42
of B ′′ , and hence B ′ is a finite A-module. Finally, since B is also the inte-
gral closure of B ′ in L, in view of Proposition 3.21, we see that B is a finite
A-module.
Remark 3.25. Even the seemingly simplest case L = K of the above theo-
rem is not obvious. In fact, the general case can be easily reduced to this
case. In general, the integral closure of a noetherian domain in its quo-
tient field need not be noetherian. A theorem of Krull and Akizuki says
that for one dimensional noetherian domains A, the integral closure of A,
and more generally, its subrings containing A, are noetherian. In dimen-
sion two, Mori and Nagata prove that the integral closure of a noetherian
domain A is noetherian but it may have subrings containing A that are not
noetherian. In dimension three, there are counterexamples to show that the
integral closure of a noetherian domain need not be noetherian. However,
the integral closure of a noetherian domain in a finite extension of its quo-
tient field is always a Krull ring. A useful generalisation of affine k-algebras
is to an important class of rings known as Nagata rings or pseudo-geometric
rings or (noetherian) universally Japanese rings.1 For more on these matters,
see Matsumura [M1] or Nagata [Na].
Exercises
1. If a rational number satisfies a monic polynomial in Z[X], then show
that it must be an integer. Deduce that Z is a normal domain. More
generally, show that any UFD is a normal domain.
2. If B/A is an integral extension of rings, then show that B/J is integral
over A/J ∩ A for every ideal J of A. Further, if S is a multiplicatively
closed subset of A, then show that S −1 B is an integral extension of
S −1 A.
√ √
3. Consider the subring A = Z[ 5] of C and let α = (1 + 5)/2. Show
that α is in the quotient field of A and α is integral over A, but α 6∈ A.
4. If A is a normal domain and S is a multiplicatively closed subset of A
such that 0 6∈ S, then show that S −1 A is a normal domain.
5. Show that if A is a domain, then A is normal if and only if A[X] is
normal.
6. If A is a domain, then show that A is integrally closed in the polyno-
mial ring A[X]. What if A is not a domain?
1
A domain A is called a Krull ring if Ap is a PID for all p ∈ Spec A with ht p = 1 and
every nonzero principal ideal of A is intersection of primary ideals belonging to height 1
primes. A noetherian ring A is called a Nagata ring if for every p ∈ Spec A, the integral
closure of A/p in a finite extension of its quotient field is a finite A/p-module.
43
7. Let A = k[X, Y ]/(Y 2 − X 2 − X 3 ) and write A = k[x, y] where x, y
denote the images of X, Y in A, respectively. Show that A is a domain
and the element y/x is in the quotient field of A and it is integral over
A, but y/x 6∈ A. Can you determine the integral closure of A in its
quotient field?
8. Let A = k[X, Y ]/(Y 2 − X 3 ). Show that A is a domain, but A is not
normal. Further show that A is isomorphic to the subring k[t2 , t3 ]
of the polynomial ring k[t], and that the integral closure of A in its
quotient field is isomorphic to k[t].
9. Let A be a domain and consider the localizations Am of A at m ∈
Max (A) as subrings of a fixed quotient field K of A. Show that if Am
is normal for each m ∈ Max (A), then so is A.
10. Prove Corollaries 3.4, 3.7 and 3.9 using the propositions preceding
them.
11. Let {Aα : α ∈ Λ} be a family of subrings of a field K such that each
Aα is a normal domain. Prove that the intersection of the Aα , as α
varies over Λ, is normal.
12. If A is a normal domain, K is its quotient field, and x is an element of
a field extension L of K such that x is integral over A, then show that
the minimal polynomial of x over K has its coefficients in A.
13. Let k be an infinite field and f ∈ k[X1 , . . . , Xn ] be a nonzero polyno-
mial. Prove that there exist a1 , . . . , an ∈ k such that f (a1 , . . . , an ) 6= 0.
Further, show that if n ≥ 1 and f is nonconstant and homogeneous,
then there exist c2 , . . . , cn ∈ k such that f (1, c2 , . . . , cn ) 6= 0.
14. Suppose k is an algebraically closed field and f, g ∈ k[X, Y ]. Deter-
mine a primary decomposition of (f, g)k[X, Y ]. Show that if f and g
are not divisible by any nonconstant polynomial in k[X, Y ], then they
have only finitely many common zeros in k2 .
15. If A is an affine k-algebra, then show that the Jacobson radical of A
(which, by definition, is the pintersection of all maximal ideals of A)
coincides with the nilradical (0) of A.
16. If in Lemma 3.22, we further assume that A = Z and that N contains a
Q-basis of K, then show that the conclusion can be refined as follows.
The quotient module M/N is a finite set and we can choose aij ∈ A,
for 1 ≤ i ≤ j ≤ n, satisfying (3.1) and with the additional property
aii > 0 for 1 ≤ i ≤ n and |M/N | = a11 a22 · · · ann = det(aij ) (3.2)
where, by convention, aij = 0 for j < i.
44
Chapter 4
Dedekind Domains
1
Dedekind published his ideas as a supplement to Dirichlet’s lectures on Number The-
ory, which were first published in 1863. Dedekind’s supplements occur in the third and
fourth editions, published in 1879 and 1894, of Dirichlet’s Vorlesungen über Zahlentheorie.
Another approach towards understanding and extending the ideas of Kummer was devel-
oped by L. Kronecker, whose work was apparently completed in 1859 but was not pub-
lished until 1882. For more historical details, see the article “The Genesis of Ideal Theory”
by H. Edwards, published in Archives for History of Exact Sciences, Vol. 23 (1980), and
the articles by P. Ribenboim and H. Edwards in “Number Theory Related to Fermat’s Last
Theorem”, Birkhäuser, 1982.
2
The term Dedekind domains was coined by I.S. Cohen [Duke Math. J. 17 (1950), pp.
27–42]. In fact, Cohen defines a Dedekind domain to be an integral domain in which every
nonzero proper ideals factors as a product of prime ideals, and he notes that the uniqueness
of factorization is automatic, thanks to Matsusita [Japan J. Math. 19 (1944), pp. 97–110].
3
Abstrakter Aufbau der Idealtheorie in algebraischen Zahlund Funktionenkörpern, Math. Ann.
96 (1927), pp. 26–61. The Aufbau paper followed another famous paper Idealtheorie in Ring-
bereichen [Math. Ann. 83 (1921), pp. 24–66] in which rings with ascending chain condition
on ideals are studied; the term noetherian rings for such rings may have been coined by
Chevalley [Ann. Math. 44 (1943), pp. 690–708]. Emmy Noether had a great appreciation of
Dedekind’s work and her favorite expression to students was Alles steht schon bei Dedekind!
45
of abstract axioms for rings whose ideal theory agrees with that of ring
of integers of an algebraic number field. This leads to a characterization of
Dedekind domains. In the next section, we will take this abstract character-
ization as the definition of a Dedekind domain, and then prove properties
such as the unique factorization of ideals as a consequence. In the sub-
sequent sections, we study the phenomenon of ramification and discuss a
number of basic results concerning it.
46
Since Z is a Dedekind domain, we obtain as an immediate consequence
the following corollary.
and this subset is, in fact, a group. In case A is a PID, we see easily (from
Corollary 3.23, for example) that FA = PA , and in this case FA is a group.
We will soon show that more generally, if A is any Dedekind domain, then
FA is a group.
Lemma 4.6. Every nonzero ideal of a noetherian ring A contains a finite product
of nonzero prime ideals of A.
Proof. Assume the contrary. Then the family of nonzero nonunit ideals of
A not containing a finite product of nonzero prime ideals of A is nonempty.
Let I be a maximal element of this family. Then I 6= A and I can not be
prime. Hence there exist a, b ∈ A \ I such that ab ∈ I. Now I + Aa and
I+Ab are ideals strictly larger than I, and I ⊇ (I+Aa)(I+Ab). In particular,
I + Aa and I + Ab are nonzero nonunit ideals. So by the maximality of I,
both I + Aa and I + Ab contain a finite product of nonzero prime ideals,
and hence so does I. This is a contradiction.
47
Lemma 4.7. Let A be a noetherian normal domain and K be its quotient field. If
x ∈ K and I is a nonzero ideal of A such that xI ⊆ I, then x ∈ A.
Lemma 4.8. Let A be a Dedekind domain and K be its quotient field. If P is any
nonzero prime ideal of A, then
P ′ = (A :K P ) = {x ∈ K : xP ⊆ A}
48
Theorem 4.10. Let A be a Dedekind domain. Then every nonzero ideal I of A
can be factored as a product of prime ideals, and this factorization is unique up
to a rearrangement of the factors. More generally, every nonzero fractional ideal
J of A factors as J = pe11 · · · pehh , for some nonnegative integer h, distinct prime
ideals p1 , . . . , ph and nonzero integers e1 , . . . , eh .4 Furthermore, the prime ideals
p1 , . . . , ph and the exponents e1 , . . . , eh are uniquely determined by J.
Proof. Assume for a moment that the assertion for integral ideals is proved.
Then for any J ∈ FA , there exists d ∈ A, d 6= 0 such that dJ is a nonzero
ideal of A. Now if dJ = p1 · · · pk and (d) = q1 · · · ql , where pi and qj are
prime ideals then J = p1 · · · pk q−1 −1
1 · · · ql . Moreover, if we also have J =
P1 · · · Pm Q−1 −1
1 · · · Qn for some prime ideals Pi and Qj (necessarily nonzero
but not necessarily distinct), then p1 · · · pk Q1 · · · Qn = q1 · · · ql P1 · · · Pm and
the uniqueness for factorization of integral ideals can be used. This yields
the desired results for nonzero fractional ideals.
To prove the existence of factorization of nonzero ideals of A into prime
ideals, we can proceed as in the proof of Theorem 4.9. Thus, let I be a
nonzero ideal of A which can not be factored as a product of prime ideals
and which is maximal with this property. Then I 6= A and if P is a nonzero
prime ideal containing I, then IP −1 is an ideal of A which is strictly larger
than I. So by the maximality of I, the ideal IP −1 is a product of prime
ideals. Multiplying on the right by P , we find that I is also a product of
prime ideals. This is a contradiction.
To prove the uniqueness, let I be any nonzero ideal of A and suppose
I = p1 · · · pr for some r ≥ 0 and prime ideals p1 , . . . , pr . We induct on
r to show that any other factorization of I as a product of prime ideals
differs from p1 · · · pr by a rearrangement of factors. If r = 0, this is evident
since a nonempty product of prime ideals will be contained in any one of
the factors, which is a proper subset of A. Assume that r ≥ 1 and the
result holds for ideals which are products of r − 1 prime ideals. Now if
I = q1 · · · qs for some s ≥ 0 and prime ideals q1 , . . . , qs , then it is clear
that s > 0. Moreover, q1 · · · qs ⊆ p1 implies that qj ⊆ p1 for some j. But
since I 6= (0), each qj is a nonzero prime ideal and hence maximal. Thus
qj = p1 . Multiplying I by p−1 1 we find that p2 · · · pr = q1 · · · qj−1 qj+1 · · · qs .
Thus by induction hypothesis r − 1 = s − 1 and p2 , . . . pr are the same as
q1 , . . . , qj−1 , qj+1 , . . . , qs after a rearrangement. This implies that r = s and
p1 , . . . , pr equal q1 , . . . , qs after a rearrangement.
Remark 4.11. Either of the following four conditions can be taken as a def-
inition for an integral domain A to be a Dedekind domain.
49
(2) Nonzero fractional ideals of A form a group with respect to multipli-
cation.
(3) Every nonzero ideal of A factors uniquely as a product of prime ide-
als.
(4) Every nonzero ideal of A factors as a product of prime ideals.
Note that (3) ⇒ (4) is obvious and from Theorems 4.9 and 4.10, we have
(1) ⇒ (2) and (1) ⇒ (3). Moreover, if (2) holds, then A isP noetherian
because if I is a nonzero ideal of A, then II −1 = A implies that ni=1 ai bi =
1 for some ai ∈ I, bi ∈ I −1 , and consequently, I = (a1 , . . . , an ). Further, if
(2) holds, then as in the proof of Theorem 4.10, the existence of a nonzero
ideal of A which can not be factored as a product of prime ideals leads to a
contradiction. This shows that (2) ⇒ (4). Hence, to prove the equivalence
of (1), (2), (3) and (4) it suffice to show that (4) ⇒ (1). This can be done but
it needs a little bit of work; for details, we refer to [ZS, Ch. V, §6].
We have seen in Example 4.2 that a Dedekind domain need not be
a UFD. On the other hand, if a Dedekind domain A is a UFD and P is
any nonzero prime ideal of A, then P must contain an irreducible ele-
ment because otherwise there will be an infinite strictly ascending chain
(a1 ) ⊂ (a2 ) ⊂ · · · of principal ideals contained in P , contradicting that A is
noetherian. Now if p ∈ P is irreducible, then (p) is a nonzero prime ideal,
and hence maximal. Hence, P = (p). Next, by Theorem 4.10, every nonzero
ideal of A is a product of prime ideals and therefore, it is principal. Thus A
is a PID. Consequently, if a Dedekind domain A is a UFD, then FA = PA or
in other words, the quotient group FA /PA is trivial.
Definition 4.12. Let A be a Dedekind domain and K be its quotient field.
The ideal class group of A, denoted by CA , is defined to be the quotient
FA /PA . When K is a number field and A = OK is its ring of integers, CA is
often denoted by CK and called the ideal class group of K. The elements of
CK are called the ideal classes of K.
As remarked earlier, if A is a Dedekind domain, then
Thus the size of the ideal class group CA is a measure of how far A is from
being a UFD. In the case when K is a number field and A = OK , it turns
out that CK is a finite (abelian) group. The order of this group is denoted
by hK and is called the class number of K.
We end this section with a result which gives a sufficient condition for
a Dedekind domain to be a PID.
Proposition 4.13. A local Dedekind domain is a PID. More generally, a Dedekind
domain that has only finitely many maximal ideals is a PID.
50
Proof. Let A be a Dedekind domain with only finitely many maximal ideals,
say, P1 , . . . , Pr . Note that the ideals P1 , . . . , Pr , and more generally, their
powers P1m1 , . . . , Prmr are pairwise comaximal. Fix any i ∈ {1, . . . , r}. Note
that Pi 6= Pi2 (because otherwise Pi = A). So we can find ai ∈ Pi \ Pi2 . By
Chinese Remainder Theorem [cf. Prop. 1.4], there exists a ∈ A such that
Now, (a) is a nonzero ideal of A with (a) ⊆ Pi , and the factorization of (a)
into prime ideals can neither contain Pj for any j 6= i nor can it contain a
power of Pi with exponent 2 or more. Hence (a) = Pi . Since every nonzero
ideal of A is a product of the Pi ’s, it must be principal. Thus A is a PID.
Remark 4.14. A ring with only finitely many maximal ideals is sometimes
called a semilocal ring. Thus the above Proposition says that a semilocal
Dedekind domain is a PID. In the case of local Dedekind domains, we can,
in fact, say more. Namely, a local Dedekind domain is what is called a
discrete valuation ring or a DVR. An integral domain A with quotient field
K is a discrete valuation ring if there exists a map v : K \ {0} → Z with the
properties
51
In this section, we shall assume that A, K, L, B are as in the Extension
Theorem 4.3. We will also let n denote the degree of L/K.
52
In view of the observations above, we shall assume without loss of gen-
erality that A is a local Dedekind domain with p as its unique nonzero
prime ideal. Then, by the Corollary 3.23, B is a free A–module of rank
n = [L : K]. Write B = Ay1 + · · · + Ayn , where y1 , . . . , yn are some elements
of B. Now for the vector space B/pB over A/p, we clearly have
n
X
B/pB = (A/p) y i
i=1
where ai ∈ A and ai denotes its residue class mod p, and the last implication
follows since {y1 , . . . , yn } is a free A–basis of B. It follows that y 1 , . . . , y n
are linearly independent over A/p, and hence
dimA/p B/pB = n.
Now we count the same dimension by a different method. First, note that
e
since P1 , . . . , Pg are distinct maximal ideals, P1e1 , . . . , Pg g are pairwise co-
e
maximal. Since pB = P1e1 · · · Pg g , by Chinese Remainder Theorem, we get
an isomorphism (of rings as well as of (A/p)–vector spaces)
g
M
B/pB ≃ B/Piei .
i=1
Now let us find the dimension of the A/p–vector space B/P e where P = Pi
and e = ei for some i. First, we note that for any j ≥ 1, pP j ⊆ P j+1 , and
hence P j /P j+1 can be considered as a vector space over A/p. We claim that
we have an isomorphism
B/P e ≃ B/P ⊕ P/P 2 ⊕ · · · ⊕ P e−1 /P e .
To see this, use induction on e and the fact that for e > 1, we clearly have
B/P e
B/P e−1 ≃ .
P e−1 /P e
Next, we note that B is a Dedekind domain having only finitely many
prime ideals (in fact, (0) and P1 , . . . , Pg are the only primes of B), and so
B must be a PID. Let t be a generator of P , and consider the map
B/P → P j /P j+1
53
and consequently, from the above direct sum representations, we get
g
X g
X
dimA/p(B/pB) = dimA/p(B/Piei ) = ei fi ,
i=1 i=1
Examples:
1. Consider the quadratic field K = Q(i), where i denotes a square
root of −1. We know that OK is the ring Z[i] of Gaussian integers. If p is
a prime ≡ 1(mod 4), then we know (by a classical result of Fermat) that p
can be written as a sum of two squares. Thus there exist a, b ∈ Z such that
p = a2 + b2 = (a + bi)(a − bi). It can be seen that (a + bi) and (a − bi) are
distinct prime ideals inPOK . Thus for the prime ideal pZ, we have g = 2,
e1 = e2 = 1 and (since ei fi = 2) f1 = f2 = 1. On the other hand, it is not
difficult to see that a prime ≡ 3(mod 4) generates a prime ideal in Z[i] and
so for such a prime, we have g = 1 = e1 and f1 = 2. The case of p = 2 is
special. We have 2 = (1 + i)(1 − i). But (1 + i) and (1 − i) differ only by
a unit (namely, −i) and thus they generate the same prime ideal. So 2 is a
ramified prime and for it, we have g = 1 = f1 and e1 = 2.
The last example illustrates the following definition.
Exercises
1. If d is a√squarefree positive integer, then determine the units in the
ring Z[ −d].
√
2. Prove that in the
√ ring Z[ √−5], the two factorizations of 6 given by
(2)(3) and (1 + −5)(1 − −5) are indeed distinct, that is, the factors
are not associates of each other. Also show that the principal ideal (6)
admits the factorization
√ √ √ √
(6) = (2, 1 + −5)(2, 1 − −5)(3, 1 + −5)(3, 1 − −5)
and that the ideals on the right are distinct prime ideals.
√
3. Let i = −1 and consider the ring Z[i] of Gaussian integers. Prove
that Z[i] is a normal domain. Deduce that Z[i] is a Dedekind domain.
Is Z[i] a PID? Justify your answer.
54
4. Let A be a Dedekind domain with quotient field K. If S is any mul-
tiplicatively closed subset of A such that 0 ∈
/ S, then show that the
−1
localization S A of A at S is a Dedekind domain with quotient field
K. Moreover, if L is an algebraic extension of K, then show that the
integral closure of S −1 A in L is S −1 B.
6. With p and Pi as above, show that a prime ideal P of B lies over p iff
P = Pi for some i. Also show that pB ∩ A = p = Piei ∩ A. Deduce that
B/pB as well as B/Piei B can be regarded as vector spaces over the
field A/p. Further show that B/Pi is a field extension of A/p whose
degree is at most n.
55
Appendix A
Primary Decomposition of
Modules
56
Note that if p = (0 : x) ∈ Ass(M ), then the map a 7→ ax of A → M
defines an embedding (i.e., an injective A–module homomorphism) A/p ֒→
M . Conversely, if for p ∈ Spec A, we have an embedding A/p ֒→ M , then
clearly p ∈ Ass(M ). It may also be noted that if M is isomorphic to some
A–module M ′ , then Ass(M ) = Ass(M ′ ).
Lemma A.2. Let I be an ideal of A. Any maximal element of the family {(0 : y) :
y ∈ M, y 6= 0 and (0 : y) ⊇ I} is a prime ideal. In particular, if A is noetherian,
then Ass(M ) 6= ∅ iff M 6= 0.
Proof. Let F := {(0 : y) : y ∈ M, y 6= 0 and (0 : y) ⊇ I} and (0 : x)
be a maximal element of F. Then (0 : x) 6= A since x 6= 0. Moreover,
if a, b ∈ A are such that ab ∈ (0 : x) and a ∈ / (0 : x), then ax 6= 0 and
b ∈ (0 : ax) ⊇ (0 : x) ⊇ I. Since (0 : x) is maximal, (0 : ax) = (0 : x).
Thus b ∈ (0 : x). This proves that (0 : x) is a prime ideal. The last assertion
follows by taking I = (0).
57
Corollary A.5. Given any A-modules M1 , . . . , Mh , we have
h
M h
[
M≃ Mi =⇒ Ass(M ) = Ass(Mi ).
i=1 i=1
Note that the above result implies that if p ∈ Ass(M ), then pAp ∈
AssAp (Mp). So, in particular, Mp 6= 0.
58
Theorem A.9. Suppose A is noetherian and M is finitely generated. Then there
exists a chain 0 = M0 ⊆ M1 ⊆ · · · ⊆ Mn = M of submodules of M such that
Mi /Mi−1 ≃ A/pi , for some pi ∈ Spec A (1 ≤ i ≤ n). Moreover, for any such
chain of submodules, we have Ass(M ) ⊆ {p1 , . . . , pn } ⊆ Supp (M ); furthermore,
the minimal elements of these three sets coincide.
Proof. The case of M = 0 is trivial. Suppose M 6= 0. Then there exists
p1 ∈ Spec A such that A/p1 is isomorphic to a submodule M1 of M . If
M1 6= M , we apply the same argument to M/M1 to find p2 ∈ Spec A,
and a submodule M2 of M such that M2 ⊇ M1 and A/p2 ≃ M2 /M1 .
By (3.1), M has no strictly ascending chain of submodules, and therefore
the above process must terminate. This yields the first assertion. More-
over, Ass(Mi /Mi−1 ) = Ass(A/pi ) = {pi }, and so by Lemma A.4, we see
that Ass(M ) ⊆ {p1 , . . . , pn }. Since localisation commutes with homomor-
phic images, (Mi /Mi−1 )pi ≃ Api /pi Api 6= 0. Hence (Mi )pi 6= 0. Thus
{p1 , . . . , pn } ⊆ Supp (M ). Lastly, if p ∈ Supp (M ), then Mp 6= 0 and so
AssAp (Mp) 6= ∅. Now Lemma A.6 shows that there exists q ∈ Ass(M ) with
q ∩ (A \ p) = ∅, i.e., q ⊆ p. This implies the last assertion.
/ Q =⇒ an M ⊆ Q for some n ≥ 1.
ax ∈ Q and x ∈
59
Lemma A.13. Let Q be a submodule of M . If Q 6= M , then
p
Q is primary ⇐⇒ Z(M/Q) = Ann(M/Q) .
Proof. An easy exercise using Corollary A.3, Corollary A.10, Exercise 3 and
Lemma A.13.
60
Lemma A.16. If M is noetherian, then every submodule of M is a finite intersec-
tion of irreducible submodules of M .
Proof. Assume the contrary. Then we can find a maximal element, say Q,
among the submodules of M which aren’t finite intersections of irreducible
submodules of M . Now Q can’t be irreducible. Also Q 6= M (because M
is the intersection of the empty family of irreducible submodules of M ).
Hence Q = N1 ∩ N2 for some submodules N1 and N2 of M with N1 6= Q
and N2 6= Q. By maximality of Q, both N1 and N2 are finite intersections
of irreducible submodules of M . But then so is Q, which is a contradiction.
Remark A.19. Given any p ∈ Spec A and a submodule Q′ of Mp, the inverse
image of Q′ under the natural map M → Mp is often denoted by Q′ ∩
M . Thus Lemma A.18 can be expressed by saying that if Q is a p–primary
submodule of M , then Qp ∩ M = Q. Note that we have been tacitly using
the fact that if Q is any submodule of M and S is any multiplicativelyclosed
subset of A, then S −1 Q can be regarded as a submodule of S −1 M .
Theorem A.20 (Primary Decomposition Theorem for Modules). Suppose
A is noetherian, M is f. g., and N is any submodule of M . Then we have
61
(ii) In (i) above, Q1 , . . . , Qh can be so chosenpthat Qi 6⊇ ∩j6=i Qj for 1 ≤ i ≤ h,
and p1 , . . . , ph are distinct, where pi := Ann(M/Qi ).
(iii) If Qi and pi are as in (ii) above, then p1 , . . . , ph are unique; in fact, we have
{p1 , . . . , ph } = Ass(M/N ). Moreover, if pi is minimal among p1 , . . . , ph ,
i.e., pi 6⊇ pj for j 6= i, then the corresponding primary submodule Qi is also
unique; in fact, Qi = Npi ∩ M .
Proof. Clearly, (i) is a direct consequence of Lemma A.4 and Corollary A.5.
Given a decomposition as in (i), we can use Corollary A.3 to reduce it by
grouping together the primary submodules having the same associated
prime so as to ensure that the associated primes become distinct. Then we
can successively remove the primary submodules contained in the inter-
sections of the remaining submodules. This yields (ii). Now let Q1 , . . . , Qh
be as in (ii). Fix some i with 1 ≤ i ≤ h. Let Pi = ∩j6=i Qi . Clearly, N ⊂ Pi
and N 6= Pi . Thus we have
Pi Pi Pi + Qi
0 6= = ≃ ֒→ M/Qi ,
N Pi ∩ Qi Qi
h
[
Ass(M/N ) ⊆ Ass(M/Qj ) = {p1 , . . . , ph }.
j=1
62
Exercises
1. Let G be a finite abelian group of order n. Suppose n = pe11 . . . pehh ,
where p1 , . . . , ph are distinct prime numbers and e1 , . . . , eh are pos-
itive integers. Let Pi denote the pi –Sylow subgroups of G for 1 ≤
i ≤ h. Show that as a Z–module, we have Ass(Pi ) = pi Z for each
i = 1, . . . , h. Deduce that Ass(G) = {p1 Z, . . . , ph Z}. More generally,
if M is a finitely generated abelian group, then M = Zr ⊕ T for some
r ≥ 0 and some finite abelian group T . If r > 0, then show that
Ass(M ) = {l0 Z, l1 Z, . . . , ls Z}, where l0 := 0 and l1 , . . . ls are the prime
numbers dividing the order of T .
63
References
64