Vous êtes sur la page 1sur 23

Principal Ideal Domains

Waffle
Mathcamp 2009

Last week, Ari taught you about one kind of simple (in the nontechnical
sense) ring, specifically semisimple rings. These have the property that every
module splits as a direct sum of simple modules (in the technical sense).
This week, well look at a rather different kind of ring, namely a principal
ideal domain, or PID. These rings, like semisimple rings, have the property that
every (finitely generated) module is a direct sum of simple modules, though
here we use simple in the nontechnical sense of easy to understand. However,
while this property of modules was almost the definition of semisimple rings,
for PIDs it is much less obvious, and the bulk of our time will be devoted
to proving this classification of modules. This classification is very powerful,
and its applications include both a complete classification of finitely generated
abelian groups and a classification of matrices up to conjugation over C (or
any algebraically closed field). The main example of a PID we will focus on is
the integers Z, for which modules are just abelian groups. However, another
important example will be k[x], the ring of polynomials in one variable over a
field. Indeed, while k[x] and Z may look like fairly different rings at first, they
are in fact very similar just by both being PIDs.1
The first, most obvious difference between what I will do and what Ari did is
that PIDs are by definition commutative. Thus throughout these notes, all
rings will be assumed to be commutative. If a, b, c . . . R are elements
of a ring, we let (a, b, c, . . .) denote the ideal they generate. More generally,
if a, b, c . . . M are elements of an R-module, we let (a, b, c, . . .) denote the
submodule they generate.

1 Generalities on commutative rings


We start with some generalities on commutative rings that will lead up to the
notion of a PID.
Definition 1.1. A ring R is an integral domain (or domain, for short) if 0 6= 1
and whenever a, b R and ab = 0, either a = 0 or b = 0. A ring is a field if
1 There is even mathematicians who are trying to make sense of an imaginary field with

one element, or F1 , such that Z would be F1 [x] (or perhaps a similar ring formed from F1
that is not quite the same as F1 [x]; Im not an expert on this so I might be getting the details
wrong). If they can make enough sense of this, they may be able to use it to prove the
Riemann hypothesis!

1
0 6= 1 and every nonzero element is a unit, i.e. has a multiplicative inverse.
If ab = 0 and a, b 6= 0, we say a and b are zero-divisors. Thus a ring is a
domain iff it has no zero-divisors. A unit u is never a zero-divisor, since ub = 0
implies b = u1 ub = u1 0 = 0. Hence every field is a domain. Fields will
usually be denoted by k instead of R.2
Note that in a domain, multiplication allows cancellation: if ac = bc and
c 6= 0, then a = b. Indeed, we have (a b)c = 0, so since c 6= 0, a b = 0.
Example 1.2. The integers Z are a domain but not a field.

Example 1.3. The rationals Q, the reals R, and the complex numbers C are
all fields.
Example 1.4. Let C(R) be the ring of continuous functions f : R R (with
pointwise addition and multiplication). Then R is not a domain. For example,
we can construct continuous functions f and g such that f (x) = 0 for x 6 [0, 1]
and g(x) = 0 for x 6 [2, 3] but f (1/2) = g(5/2) = 1. Then f g = 0 but f, g 6= 0.
Example 1.5. For any prime p, Z/(p), the integers mod p, is a field. This is a
standard fact from number theory, that you can divide (except by 0) modulo a
prime.
Example 1.6. If n is not prime, Z/(n) is not a domain. Indeed, write n = ab
where a, b 6= 1. Then modulo n, a and b are nonzero but ab 0.
Example 1.7. Let R be a ring. Then we can form the ring R[x] of polynomials
(in one variable) with coefficients in R. A polynomial is a formal expression
an xn + an1 xn1 + . . . + a1 x + a0 where ai R, and polynomials are added and
multiplied according to the familiar rules from elementary algebra. If f (x) =
ak xk , the largest n such that an 6= 0 is the degree deg(f ) of f , and we call an
P
the leading coefficient of f .3 If R is a domain, then deg(f g) = deg(f ) + deg(g)
since the leading coefficients of f and g multiply to give a nonzero leading
coefficient of f g. It follows that if R is a domain, so is R[x]. Polynomials with
leading coefficient 1 are called monic.

Definition 1.8. An ideal I in a ring R is prime if R/I is a domain and maximal


if R/I is a field.
The name prime comes from the fact that, as we saw in Examples 1.5 and
1.6, the ideal (n) Z is prime iff n is a prime number. The name maximal
come from the following result.

Proposition 1.9. An ideal I R is maximal iff there is no strictly larger ideal


J such that I J R.
2 Blame the Germans for this.
3 We formally write deg(0) = .

2
Proof. Note that ideals J R containing I are equivalent to ideals of R/I.
Indeed, let p : R R/I and q : R R/J be the canonical maps. If I J,
then q(I) = 0 so we can factor q uniquely as f p for some f : R/I R/J, which
is surjective since q is. We then obtain an ideal J 0 = ker(f ) in R/I. Conversely,
if we start with J 0 R/I, we can define an ideal J I as the kernel of the
composition R R/I (R/I)/J 0 . It is easy to check that these are inverse to
each other.
Thus it suffices to show that k = R/I is a field iff its only (proper) ideal is
0. If k is a field and I k is an ideal and a I is nonzero, then 1 = a1 a I,
from which it follows that I = k, a contradiction. Conversely, if k is not a field
and a k is nonzero and not a unit, then 1 6 (a) so (a) is a nonzero proper
ideal.
It follows by the Axiom of Choice that any ideal is contained in a maximal
ideal (see Exercise 1.14). Intuitively, this makes sense, because we can keep
making an ideal bigger and bigger, and if we cant make it any bigger weve hit
a maximal ideal.4
As a general principle, domains and fields are much nicer than arbitrary
rings. A central idea of commutative algebra and algebraic geometry is that
rings can often be understood by studying the collection of their maximal or
prime ideals. We will not explore this in this class, but I think Mark will in his
Commutative Algebra class next week.

1.1 Exercises
Exercise 1.10. Show that an ideal I R is prime iff whenever ab I, either
a I or b I.

Definition 1.11. Let R be a ring. Then a sequence of homomorphisms


i p
0KM N 0

is a short exact sequence if i is injective, p is surjective, and ker(p) = Im(i).


Equivalently, up to isomorphism, K is a submodule of M and N = M/K, with
i and p the obvious maps.
i p
Exercise 1.12. Let 0 K M N 0 be a short exact sequence of
modules. Prove that the following are equivalent:

1. M
= K N , with i : K K N and p : K N N the canonical
maps.
2. There is a homomorphism q : M K such that qi = 1.
3. There is a homomorphism j : N M such that pj = 1.
4 This argument, when put into the right framework, is actually rigorous! Ask me about it

or take my set theory class to learn how.

3
If these conditions hold, we say that the short exact sequence splits.
The next exercise uses the following result, which is equivalent to the Axiom
of Choice.

Lemma 1.13 (Zorns Lemma). Let P be a partially ordered set, and suppose
that whenever C P is totally ordered, there is x P such that x c for
all c C (briefly, every chain in P has an upper bound). Then there is an
element z P which is maximal: there is no x P such that x > z.
Exercise 1.14.

(a): Let I R be an ideal. Show that there is a maximal ideal containing


I. (Hint: To get upper bounds, use that an ideal is proper iff it does not
contain 1).
(b): Show that an element a R is contained in a maximal ideal iff it is not a
unit.

Exercise 1.15.
(a): Show that the polynomial ring R[x] satisfies the following universal prop-
erty: a homomorphism from R[x] to another ring S is equivalent to a ho-
momorphism from R to S together with a chosen element in S (the image
of x).5
(b): Show that an R[x]-module is equivalent to an R-module M together with
an element of EndR (M ) (corresponding to multiplication by x). (Hint: An
S-module structure on an abelian group M is the same as homomorphism
from S to the (noncommutative) ring EndZ (M ).)

Exercise 1.16.
(a): Let R be a domain. Let k be the set of formal expressions a/b where
a, b R and b 6= 0, and we say a/b = c/d if ad = bc. Show that with
the usual addition and multiplication of fractions, k is a field. Show that
a 7 a/1 is an injective ring-homomorphism from R into k. We call k the
field of fractions of R.
(b): Show that a ring is a domain iff it is a subring of a field.
Exercise 1.17. Show that if a domain is finite, it is a field.

5 By
equivalent, we mean that there is a naturally defined bijection between the set of
homomorphisms R[x] S and the set of pairs (f, s) where f : R S and s S.

4
2 PIDs and prime factorization
We are now ready to define PIDs.
Definition 2.1. An ideal I in a ring R is principal if it is equal to (a) for some
a R.
A principal ideal domain (PID) is a domain in which every ideal is principal.
Note that every ideal in a PID is trivially finitely generated, so a PID is
Noetherian. Recall that this implies that a submodule of any finitely generated
module is finitely generated.
The most basic example of a PID is Z.
Proposition 2.2. The integers Z are PID.
Proof. Let I Z be an ideal; we want to show I is principal. If I = 0, this
is trivial. If I 6= 0, let n be the smallest positive integer in I (if m I, then
m I, so some positive integer is in I). We claim that I = (n). For any
m I, we can divide m by n with remainder to write m = qn + r where the
remainder r satisfies 0 r < n. But then r = m qn I, so minimality of n
forces r to be 0. Hence m = qn (n). Since m I was arbitrary, I = (n).
It is important to note that this proof used only the fact that we have
division with remainder and the remainder is strictly smaller than what you
were dividing by in a sufficiently nice sense. For polynomials over a field in a
variable x, we can do long division in base x to show that for any f, g k[x],
g = qf + r where r has strictly lower degree than f . It follows by a nearly
identical proof that k[x] is also a PID if k is a field.
Proposition 2.3. Let k be a field. Then the polynomial ring k[x] is a PID.
Proof. Let I k[x] be an ideal; we want to show I is principal. If I = 0, this
is trivial. If I 6= 0, let f I be a nonzero polynomial of minimal degree. We
claim that I = (f ). For any g I, we can divide g by f with remainder to write
g = qf + r where the remainder r has strictly smaller degree than f . But then
r = g qf I, so minimality of f forces r to be 0. Hence g = qf (f ). Since
g I was arbitrary, I = (f ).
In Exercise 2.10, you will generalize this to an abstract ring that has a good
notion of division with remainder, called a Euclidean domain.
Note that most rings, even most domains, are not PIDs. For example, in
the ring k[x, y] = k[x][y] of polynomials in two variables with coefficients in a
field, the ideal (x, y) cannot be generated by a single element (see Exercise 2.7).
Similarly, in the ring Z[x] of polynomials in one variable over Z, the ideal (2, x)
is not principal.
Prime numbers and the unique factorization of integers into primes is ex-
tremely useful in studying the integers and number theory. In fact, these are
not special to the integers, but work just as well in any PID. First, we need
some definitions.

5
Definition 2.4. Let R be a domain and a, b R. Then a divides b, or a|b, if
there exists c R such that b = ac. Equivalently, a|b if (b) (a).
We say p R is irreducible if p is not a unit and whenever p = ab, either
p|a or p|b. We say p is prime if the ideal (p) is a prime ideal.

Note that if a|b and b|a, then (a) = (b). In this case, we can write a = ub
and b = va. It follows that a = ub = uva, so by cancellatin uv = 1. Thus if
a|b and b|a, a and b differ by multiplying by a unit, and the converse clearly
also holds. In this case, we say a and b are associate. Whenever we talk about
divisibility properties, everything is always up to units; we will generally not
care to distinguish an element from any of its associates. It follows from this
discussion that if p is irreducible and p = ab then one of a and b is an associate
of p and the other is a unit.
If R = Z, then n and m are associate iff n = m, since the only units are
1. By our earlier discussion, we see that this definition of prime is equivalent
to the usual definition for integers, and is also equivalent to irreducibility. In
fact, the equivalence of primeness and irreducibility holds in any PID.
Irreducible elements of k[x] are called irreducible polynomials. Because de-
grees add when you multiply polynomials and 1 has degree 0, any unit must
have degree 0. Since k is a field, every degree 0 polynomial is a unit, so the
units are exactly k = k {0} k[x]. It follows that any degree 1 polynomial is
irreducible: if f = gh and deg(f ) = 1, then deg(g) = 0 or deg(h) = 0, so one of
them must be a unit. If k is algebraically closed, then every polynomial factors
into linear polynomials, so these are the only irreducible polynomials. However,
for example, if k = R, then x2 + 1 is also irreducible (see Exercise 2.8).
Theorem 2.5. Let R be a PID and let p R be nonzero. Then TFAE:

1. p is irreducible.
2. p is prime
3. The ideal (p) is maximal
Proof. (1 3): Suppose p is irreducible. Then whenever a|p, b is either a unit
or an associate of p. Since (p) (a) iff a|p, we get that (p) (a) only holds for
(a) = (p) or (a) = R. Since R is a PID, every ideal is (a) for some a, so this
says exactly that (p) is maximal in the sense of Proposition 1.9.
(3 2): Every maximal ideal is prime.
(2 1): Let p be prime and suppose p = ab. Then modulo (p), ab 0.
Since R/(p) is a domain, it follows that either a 0 or b 0, i.e. a (p) or
a (p). This is equivalent to either p|a or p|b, as desired.
Note that the proof of (2 1) (or Exercise 1.10) shows that that for p prime,
if p|bc then p|b or p|c. By induction, it follows that if p|b1 b2 bn , then p|bi for
some i. From this, we can prove that any PID has unique factorization into
primes. This is a generalization of the Fundamental Theorem of Arithmetic for
the case of Z.

6
First, we note that changing an element by a unit does not change its divis-
ibility properties. If we allow ourselves to change factors by units, factorization
is not even unique in Z. For example 6 = 2 3 = (2)(3), where 2, 3, 2,
and 3 are all prime. For the integers, we remedy this by arbitrarily choosing a
representative of each associate class, namely the positive one. For polynomials
over a field, we can similarly choose to use only monic polynomials (polynomials
with leading coefficient 1) in our factorizations. For a general PID, we will just
arbitrarily choose representatives of each associate class of primes, and call these
the chosen primes. Equivalently, we are picking a generator for every maximal
ideal.

Theorem 2.6 (Unique Factorization). Let R be a PID. Then every nonzero


a R can be factorized as upd11 pd22 pdnn , where u is a unit, the pi are chosen
primes, and di > 0. This factorization is unique up to permuting the pi .
Proof. First, we prove existence of the factorization. If a is a unit, this is
trivial. Otherwise, let P be any maximal ideal containing (a). Then P = (p1 )
for some chosen prime p1 , and p1 |a since (a) (p1 ). Write a = p1 a1 . Repeat
the argument with a1 in place of a to write a1 = p2 a2 if a1 is not a unit.
Continue by induction. If we ever get an = u to be a unit, we are done, since
a = p1 a1 = p1 p2 a2 = = p1 p2 pn u. If we never get a unit, we get S an infinite
ascending chain of ideals (a) (a1 ) (a2 ) (a3 ) . . .. Let I = (an ); then
I is an ideal. But then I = (b) since R is a PID, and b (an ) for some n. But
then I = (an ) (an+1 ), so (an ) = (an+1 ), a contradiction. Hence the process
must eventually stop with an a unit, and we get Pthe desired
P factorization.
Now we prove uniqueness by induction on di . If di = 0 P(i.e., n = 0),
then a = u is just a unit, and uniqueness is obvious. Now suppose di > 0 and
a = upd11 pd22 pdnn = vq1e1 q2e2 qm
em
are two different factorizations. Then p1
divides the product on the right-hand side and is prime, so p1 must divide one of
the factors. Since no two chosen primes are associate, this implies that p1 = qi
d1 1 d2 dn
for some i. Cancelling these common factors,
e1 e2 ei 1 em
P we get b = up1 p2 pn =
vq1 q2 qi qm . We have decreased di by one, so by induction the
factorization of b is unique, so these two factorizations are the same up to
permutation. It follows that the two original factorization of a were the same
up to permutation.

2.1 Exercises
Exercise 2.7. Let k be a field (though any ring will do for the first part). Show
that the ideal (x, y) k[x, y] is not principal. Conclude that k[x, y] is a domain
but not a PID.6
Exercise 2.8.
6 In fact, it can be shown that k[x, y] has unique factorization, showing that a domain with

unique factorization (or UFD) need not be a PID.

7
(a): Show that if a polynomial f (x) k[x] is irreducible and of degree greater
than 1, then f has no roots. (Hint: If f (a) = 0, show by division that
(x a)|f .)
(b): Show that the polynomial x2 + 1 is irreducible as an element of R[x].

(c): Show that the polynomial x4 4 is reducible as an element of Q[x], but


has no roots.
The next exercise gives another class of examples of PID which are important
in more advanced commutative algebra.

Exercise 2.9. A local ring is a ring R with a single maximal ideal P . By


Exercise 1.14(b), this means that every element of R P is a unit. A discrete
valuation ring (DVR) is a Noetherian local domain such that the maximal ideal
P is principal. Let R be a DVR and let p be a generator for P .
(a): Show that every nonzero element of R is of the form upn for u a unit.
(Hint: Imitate the proof of existence of prime factorizations in a PID and
use Noetherianness.)
(b): Show that R is a PID, and in fact that every nonzero ideal is (pn ) for some
n.

(c): Let p Z be a prime, and let Z(p) be the set of rational numbers whose
denominators are relatively prime to p. Show that Z(p) is a DVR with
maximal ideal (p).
P
(d): Let k be a field, and let k[[x]] = { n=0 an xn : an k} be the ring of
formal power series in one variable with coefficients in Pk. Show that k[[x]]
is a DVR with maximal ideal (x). (Hint: Show that an xn is a unit if
a0 6= 0, and use this to prove that (b) above holds for R = k[[x]] and p = x.)
The example of Exercise 2.9(c) is generalized in Exercise 3.15.
Exercise 2.10. A Euclidean domain is a domain R together with a function
d : R {0} N satisfying d(ab) max(d(a), d(b)) and such that if a, b R and
b 6= 0, then there exist q and r such that a = qb + r and r = 0 or d(r) < d(a).
For example, if R = Z we can let d(n) = |n| and if R = k[x] we can let
d(f ) = deg(f ). Show that any Euclidean domain is a PID.
Exercise 2.11. Let R be a PID and a, b R. Then a greatest common divisor
(gcd) of a and b is c R such that c|a, c|b, and if d|a and d|b then d|c. We write
c = gcd(a, b).
(a): Show that c is a gcd of a and b iff c generates the ideal (a, b). Conclude
that any two elements have a gcd, well-defined up to units.
(b): Show that if c = gcd(a, b), then there exist r, s R such that c = ra + sb.
Conversely, show that if c = ra + sb and c|a and c|b, then c = gcd(a, b).

8
(c): Suppose R is a Euclidean domain. Then show that the following algorithm,
known as the Euclidean algorithm, computes gcd(a, b) and allows one to
explicitly write it in the form ra + sb: Assume WLOG that d(b) d(a).
If a|b, stop and say the gcd is a. Otherwise, write b = qa + r with d(r) <
d(a). Now repeat the algorithm, but replace b with r. The algorithm will
eventually terminate because d(a) + d(b) decreases with each step. (To get
an idea of how the algorithm works, you may want to try it with R = Z
and a = 18, b = 26.)7

7 The Greeks invented the Euclidean algorithm to prove that Z is a PID (though they didnt

call it that!). Oddly, I have seen multiple references in which the example used to demonstrate
the Euclidean algorithm is 18 and 26.

9
3 Modules over a PID
In this section, R will always denote a PID, and primes will be taken to mean
chosen primes.
We can now get to the real meatclassifying finitely generated modules
over a PID. Note that a Z-module is the same as an abelian group, so this will
also give a complete classification of finitely generated abelian groups. This is
perhaps more impressive when you consider that is it generally accepted to be
hopeless to even classify finite nonabelian groups.
What about the case R = k[x]? In this case, by Exercise 1.15, an R-module is
a k-module V (i.e., vector space) together with an endomorphism A Endk (V )
(i.e., a linear transformation on V ). If (V, A) and (W, B) are k[x]-modules, a ho-
momorphism between them is a k-linear map T : V W that intertwines with
the endomorphisms: T A = BT (this just is the property that T (xv) = xT (v),
since A and B correspond to multiplication by x). It follows that (V, A) and
(V, B) are isomorphic iff there is a k-linear automorphism T of V such that
T A = BT , or B = T AT 1 . Thus by classifying k[x]-modules, we will classify
linear transformations on vector spaces up to conjugation by linear automor-
phisms. Or, in simpler language, we will classify matrices up to conjugation.
Lets now look at what the classification is.
Definition 3.1. An R-module M is cyclic if it is generated by a single element.
This generalizes the notion of a cyclic group. Note that if M is generated
by x, then f (a) = ax is a surjective homomorphism from R to M . Hence cyclic
modules are exactly those of the form R/I for some ideal I.
Here is the main theorem of this section.
Theorem (Classification of Modules over a PID). Let M be a finitely gen-
eratedLR-module. Then M is isomorphic to a direct sum of cyclic modules
Rn R/(qi ), where each qi = pdi i is a power of a prime. This decomposition
is unique up to permuting the factors.
The number n is sometimes called the Betti number of M and the qi are the
elementary divisors or torsion coefficients.
There are several steps to proving this theorem. Basically, we can prove it
separately for the summands Rn and for the summands R/(q) corresponding to
each prime.

3.1 Torsion modules


Definition 3.2. Let M be an R-module and x M . Then the annihilator of
x is (0/x) = {a R : ax = 0}. Note that (0/x) is an ideal. If x 6= 0 and
(0/x) 6= 0, we say x is a torsion element. If (0/x) = (a), we say a is the period
of x.
If every (nonzero) element of M is torsion, we say M is a torsion module. If
no element of M is torsion, we say M is torsion-free.

10
Note that the period of an element is only defined up to units; this should
not cause any confusion. If x has period a, then (x) M is isomorphic to
R/(a).
Example 3.3. Let R = Z. Then a module is a finitely generated torsion module
iff it is a finite abelian group. Indeed, it is clear that any element of a finite
abelian group must be torsion (otherwise its multiples would all be different
and there would be infinitely many of them). Conversely, if M is generated by
x1 , . . . , xn with periods a1 , . . . , an , then any element of M can be written as
b1 x1 + + bn xn where each bi satisfies 0 bi < ai , so M is finite.

Example 3.4. Let R = k[x]. If (V, A) is an k[x]-module and v V is torsion,


then (v) = k[x]/(f (x)) for some nonzero polynomial f , which (by changing by
a unit) we may assume to be monic. If f (x) = xn + an1 xn1 + . . . , then in
k[x]/(f ), xn = an1 xn1 . . . is k-linearly dependent on the lower powers
of x. We similarly can see that all higher powers of x can be spanned just by
B = {1, . . . , xn1 }. In fact, B is a basis for k[x]/(f ) as a k-vector space, since
every nonzero element of (f ) has degree at least n, so no linear combination of
elements of B can vanish in k[x]/(f ).
Thus if v is torsion, (v) is finite-dimensional. In fact, more generally, a k[x]-
modules is finitely generated torsion iff it is finite-dimensional as a vector space,
which you will prove in Exercise 3.11.

In this section, we will prove the classification theorem for torsion modules,
which is the general classification theorem in the case where the Betti number
is 0.
First, we show that we can treat each prime separately when analyzing a
torsion module.

Definition 3.5. Let p R be a prime and M be an R-module. Then an


element x M is p-torsion if its period is a power of p. If every element of M
is p-torsion, we say M is a p-module.
The name p-module generalizes the traditional term p-group for a group
whose order is a power of p.

Lemma 3.6. Let M be a finitely generated torsion R-module. For each p, let
Mp = {x M : xL is p-torsion}. Then Mp is a submodule of M , and M splits
as the direct sum Mp over all primes p.
Proof. First, we show Mp is a submodule. Let x, y Mp , with (0/x) = (pn )
and (0/y) = (pm ) and WLOG m n. Then pn (x + y) = 0, so pn (0/x + y).
Since p is prime, the only ideals containing (pn ) are (pk ) for k n. Hence the
period of x + y is a power of p, so x + y Mp . A similar argument shows that
Mp is closed under scalar multiplication.
L P
Now wePshow that M = Mp . This consists of two statements: M = Mp
and Mp q6=p M (q) = 0 for all p. That is, the Mp generate all of M and they
are linearly disjoint. We prove the second statement first.

11
P P
Suppose x Mp q6=p M (q) is nonzero; write x = q M (q).
yq for yP
Let r be the product of the periods of the yq . Then we have rx = ryq = 0,
so r (0/x). But x Mp , so (0/x) = (pn ) for some n. Since r, as a product
P of q for q 6= p, is relatively prime to p, this is impossible. Hence
of powers
Mp q6=p M (q) = 0. P
Now we show that M = Mp , i.e. the Mp together generate M . Let x M
be nonzero and factor the period of x as r = pd11 pdnn (ignoring units for
convenience). Let ai = r/(pni i ) and yi = ai x; i.e. multiply x by all of its period
except the pi part. Then pni i yi = rx = 0, and as above this implies yi M (pi ).
Now consider the ideal I = (a1 , . . . , an ) R. Since R is a PID, I = (a) for
some a. We then have a|ai for all i. By unique factorization, this implies P a = 1,
so I is all of R. In particular, 1 I = (a1 , . . . , an ), so we can write 1 = bi ai
for some bi R. But now we have
X  X
x= bi ai x = bi yi .
P
Since yi Mp , we conclude that x Mp , as desired.
You can think of this result as analogous to splitting a semisimple module
into its parts corresponding to each isomorphism class of simple modules (i.e.,
its Wedderburn components).
Remember that we want to prove that we can split a module as a direct
sum of modules of the form M/(q) where q is a prime power. By Lemma 3.6, it
suffices to show this for p-modules, in which case every q will be a power of p.
Heres the idea
Lbehind the proof. Supposing M is a p-module and we already
know that M = R/(pdi ), let d1 be the largest of the di . Then if we take any
element x M of period pdi , M will split as (x) M/(x). Why should we
expect this? Consider the example of R = Z, p = 2 and M = Z/(4) Z/(2), for
concreteness sake. Then, for example, if we pick the element x = (2, 0), which
has period 2, the short exact sequence

0 (x) M M/(x) 0

does not split, because there is an element y = (1, 0) satisfying 2y = x, which


clearly does not hold in the direct sum (x) M/(x) = Z/(2) Z/(2) Z/(2).
On the other hand, if we had chosen an element x of period 4, it would have
split. Indeed, this is obvious if x = (1, 0). If, say, x = (1, 1), note that
(x) = {0, (1, 1), (2, 0), (1, 1)} is still disjoint from the Z/(2) = {0, (0, 1)}. Thus
it seems that an obstacle to splitting is being able to divide x by p, and it will
turn out that this is the only obstacle. This cant happen if the period of x is
pd1 , since if py = x, y would have period pd1 +1 , contradicting maximality of d1 .
Thus the idea is, take an element x of M of maximal period, split M = (x)
M/(x), and repeat by induction on M/(x). To guarantee that this induction
works, we need a lemma.
Lemma 3.7. Let M be a finitely generated p-module, and let x M have
maximal period (i.e. period pd , where d is maximal). Let N = M/(x) and

12
f : M N be the canonical map. Then as a R/(p)-vector space, N/pN has
smaller dimension than M/pM .
Proof. First, note that M/pM and N/pN are vector spaces over the field R/(p);
this is straightforward to check. Also, they are finite-dimensional since M and
N are finitely generated as R-modules, and the same generators can be used.
Now note that by maximality of the period of x, x 6 pM . Also, note that
f induces a natural surjective map F : M/pM N/pN . Since F (x) = 0
but x is nonzero as an element of M/pM , ker(F ) 6= 0. Since dim(M/pM ) =
dim(N/pN ) + dim(ker(F )), were done.

Thus when we say by induction, we will mean by induction on dim(M/pM ).


We can now prove the classification theorem for p-modules.
Theorem 3.8. LetL M bedai finitely generated p-module. Then M is isomorphic
to a direct sum M/(p ). The numbers di are uniquely determined up to
permutation.

Proof. We prove existence of the splitting by (strong) induction on dim(M/pM ).


Let x M have maximal period, say period pd . Such an x exists since M
is finitely generated: if M is generated by elements xi with periods pdi , then
it is easy to see that every element of M has period at most pd where d is the
greatest of the di . By LemmaL3.7 and the induction hypothesis, we can write
N = M/(x) as a direct sum R/(pdi ). If we can show M = N (x), were
d
done, since (x) = R/(p ). We thus want to lift generators of N to generators of
M to induce a splitting. For this, we use the following lemma.
Lemma 3.9. Let M be a p-module, and let x M have maximal period (i.e.
period pd , where d is maximal). Let N = M/(x) and f : M N be the
canonical map, and let y N . Then there is y M such that f (y) = f (
y ) and
y has the same period as y.
Proof. Let pe be the period of y and let z M be any lift of y. Then f (pe z) =
pe y = 0, so pe z (x); write pe z = cpn x, where p does not divide c and n d.
Note that since x has period pd , pe z has period d n, so z has period d + e n.
By maximality of d, d + e n d, so e n. Now let y = z pne cx. Then
f (y) = y, and pe y = pe z pn cx = 0, as desired.
Now let yi be generators of N corresponding to the direct summands R/(pdi ).
Let yi M be lifts of yi having the same period, as in Lemma 3.9. We then
obtain a map r : N M given by r( yi ) = yi ; this is well-defined since N is
just a direct sum and yi has the same period as yi . Furthermore, for each yi ,
f (r(
yi )) = yi . Since the yi generate N , this implies that f r = 1. That is, r
induces a splitting of the short exact sequence 0 (x) M N 0. Hence
M = N (x), and were done.
R/(pdi ), it is clear that
L
Now we show uniqueness of the di . If M =
the
L dimension of M/pM is just the total number of di . Furthermore, pM =
R/(pdi 1 ). Thus the dimension of pM/p2 M is the number of di that are

13
greater than 1. Similarly, by looking at the dimension of pk M/pk+1 M for all k,
we can see how many di there are that are greater than k, for all k. This shows
that the di are uniquely determined by M , as desired.
Corollary 3.10 (Classification of finitely generated torsion modules). Let M
be
L a finitely generated torsiondiR-module. Then M is isomorphic to a direct sum
R/(qi ), where each qi = pi is a power of a prime. The qi are unique up to
permutation.
Proof. Existence of this splitting is immediate
L from Theorem
L3.8 and Theorem
3.6. For uniqueness, note that if M = R/(qi ), then Mp = p|qi R/(qi ). Thus
the uniqueness of the representation of Mp in Theorem 3.8 gives the uniqueness
for all of M .

3.1.1 Exercises
Exercise 3.11. Show that a k[x]-modules is finitely generated torsion iff it is
finite-dimensional as a k-vector space.
Exercise 3.12. Let f (x) = xn + an1 xn1 + + a0 . Write down the matrix
for multiplication by x in k[x]/(f (x)) with respect to the basis {1, x, . . . , xn1 }.
By Corollary 3.10, every linear transformation on a finite-dimensional vector
space is a direct sum of matrices of this form for some choice of basis.
Exercise 3.13. Let a R and factor a = pni i . Show that R/(a) R/(pni i ).
Q L
=
Exercise 3.14. Let M be a finitely generated torsion R-module. L Show that
M is isomorphic to a unique direct sum of cyclic modules R/(ai ) such that
ai+1 |ai for all i. These ai are called the invariant factors of M , and are some-
times more useful than the elementary divisors. (Hint: For each prime p, Qlet np
be the highest power of p appearing as an elementary divisor. Let a1 = p pnp .
Define the other ai in a similar way, and use Exercise 3.13.)
Exercise 3.15. Let R be a ring and P R be a prime ideal. Define the
localization of R at P to be the set RP of fractions a/s for s 6 P , where we
identify a/s and b/t if u(atbs) = 0 for some u 6 P . Informally, we are allowing
division by elements not in P .
(a): Show that RP is a well-defined ring with ordinary addition and multiplica-
tion of fractions, and that a 7 a/1 is a homomorphism from R to RP .
(b): Show that if R is a domain and P = 0, we just get the field of fractions of
R. Show that if R = Z and P = (p), we get the ring Z(p) of Exercise 2.9(c).
(c): Show that RP is a local ring with maximal ideal generated by the image of
P R under the map R RP .8 (Hint: It suffices to show that anything
not in this ideal is a unit.)
8 It can be shown that a localization of a Noetherian ring is Noetherian, and that if R is a

Noetherian domain and P is a minimal nonzero prime ideal, then RP is a DVR.

14
(d): If M is an R-module, we can define the localization MP similarly as frac-
tions x/s with x M and s R P , with the same identification. Show
that if R is a PID and M is torsion, the localization M(p) is naturally iso-
morphic to the module Mp of Lemma 3.6. (Hint: Localization preserves
direct sums. Use Lemma 3.6, and show (Mp )(p) = Mp and (Mq )(p) = 0 if
q 6= p.)
Localization is an extremely important technique in commutative algebra,
though we dont have time to discuss it in more depth here. Unless it turns out
that we do, in which case I might talk about it some more on Saturday.

15
3.2 Torsion-free modules
Now we look at the other extreme case of the classification, torsion-free modules.
We want to show that any torsion-free module M is free, i.e. isomorphic to
a direct sum of copies of R. Generators of M inducing such a direct sum
representation are called aPbasis for M . More concretely, {xi } is a basis for M
iff they generate M and ai xi = 0 implies ai = 0 for all i.PLike bases for a
vector space, this implies that any x M can be written as ai xi for unique
ai . The size of the basis {xi } is called the rank of M (like the dimension of
a vector space. Note that for any prime p, {xi } will also give a basis for the
R/(p)-vector space M/pM . This implies that the rank of M is well-defined.
The first step to proving all torsion-free modules are free is the following.
Lemma 3.16. Let M be a finitely generated free R-module, and let N M be
a submodule. Then N is free.
Proof. Let {x1 , . . . , xn } be a basis for M , let Mr = (x1 , . . . , xr ) and let Nr =
N Mr . We show by induction that Nr is free; since Nn = N this will imply
N is free. For r = 0, N0 = 0, so this is trivial. Now suppose Nr is free,
and let I be the set of a R such that some element of Nr+1 has a as its
coefficient for xr+1 . Equivalently, identifying (xr+1 ) with R, I is the ideal
(Mr + Nr+1 )/Mr Mr+1 /Mr = (xr+1 ).
If I = 0, then clearly Nr+1 = Nr , so Nr+1 is free. Otherwise, since R
is a PID, we can write I = (a) for a nonzero. Let w Nr+1 be such that
w = b1 x1 + . . . br xr + axr+1 . If y Nr+1 , then the xr+1 coefficient of y is ca for
some c R, so y cw Nr . Thus Nr+1 = Nr + (w). Furthermore, Nr (w) = 0
clearly, since any nonzero multiple of w has nonzero xr+1 coefficient. Thus
Nr+1 = Nr (w); since Nr and (w) = R are free, so is Nr+1 .
Note that this proof also shows that the rank of N is at most the rank of
M.
We can now prove that torsion-free modules are free.
Theorem 3.17. Let M be an R-module. Then M is free, i.e. isomorphic to
Rn for some n, and this n is unique.

Proof. Uniqueness of n is just the well-definedness of rank of free modules.


Now let {v1 , . . . , vnP } M be a maximal linearly independent set, i.e. a
maximal set such that ai vi = 0 implies ai = 0. Let N = (v1 , . . . , vn ); by linear
independence, N is free. For any y M , there is some nonzero a M such that
ay N . Indeed, if no such a existed, then y would be linearly independent from
{v1 , . . . , vn }, contradicting maximality.
Q In particular, let {y1 , . . . , ym } generate
M and ai yi N , and let b = ai . Then byi N for all i, so bM N . But
now multiplication by b is a homomorphism from M to N , and is injective since
M is torsion-free. Thus M is isomorphic to a submodule of N . By Lemma 3.16
and freeness of N , M is free.

16
3.3 Exercises
Exercise 3.18. Let k be a field and let M be the ideal (x, y) k[x, y]. Show
that M is a finitely generated torsion-free k[x, y]-module, but M is not free.
(Hint: Show that I = (x, y) and J = (x 1, y) are both maximal ideals and
k[x, y]/I
= k[x, y]/J
= k as rings. Then show that M/IM and M/JM have
different dimensions as k-vector spaces, and that this is impossible for a free
module.9 )
Exercise 3.19. Find an example of a (non-finitely generated) torsion-free Z-
module which is not free.
Exercise 3.20. Find an example of a (non-finitely generated) torsion-free k[x]-
module which is not free.
Exercise 3.21. Find an example of a (non-finitely generated) torsion Z-module
which is not a direct sum of cyclic modules.

Exercise 3.22. Let R be a PID. Show that a submodule of any (not necessarily
finitely generated) free R-module is free. (Hint: Let M be free and N M .
Use Zorns Lemma on the poset of linearly independent subsets of N which
generate free submodules F N such that N/F is torsion-free. Show that a
maximal such basis must generate N itself by imitating the proof of Lemma
3.16. To show that every chain has an upper bound, show that if (F ) are
S a
chain of nested submodules of N such that N/F is torsion-free, then N/ F
is torsion-free.)

3.4 Putting it all together


We can now combine Corollary 3.10 for torsion modules and Theorem 3.17 for
torsion-free modules to prove the full classification theorem.
Theorem 3.23 (Classification of Modules over a PID). Let M be a finitely
generated
L R-module. Then M is isomorphic to a direct sum of cyclic modules
Rn R/(qi ), where each qi = pdi i is a power of a prime. This decomposition
is unique up to permuting the factors.
Proof. Let T = {x M : x is torsion}, the torsion part of M . Then T is a
submodule of M ; the argument is the same as the argument in Lemma 3.6 that
Mp is a submodule for p a prime. Clearly T is torsion, and T is finitely
L generated
by Noetherianness, so Corollary 3.10 says it splits uniquely as R/(qi ) for qi
prime powers.
Let N = M/T , and let f : M N be the canonical map. Then N is
N is nonzero and a
torsion-free. Indeed, if x x = 0, let x M be such that
. Then ax T , so b(ax) = 0 for some nonzero b. But then (ba)x = 0
f (x) = x
so x T , contradicting the assumption that x 6= 0.
9 Here IM is the submodule of M generated by products ix for i I and x M .

17
Thus N is torsion-free and hence free by Theorem 3.17. If the short exact
sequence
0T M N 0
were to split, we would get the desired direct sum representation fo M . But a
splitting map r : N M is easily constructed: let { xi } be a basis for N , and let
r(
xi ) = xi for some xi such that f (xi ) = x
i . This is well-defined since
L N is free.
We clearly have f r = 1, so we get a splitting M = N T = Rn R/(qi ).
L Finally, we show uniqueness L of the representation of M . Given M = Rn
R/(qi ), it is clear that T = R/(qi ) and N = Rn . The uniqueness thus
follows from the uniqueness of the qi and n given in Corollary 3.10 and Theorem
3.17.

18
4 Applications to linear algebra
The case of R = Z, in which we have a classification of finitely generated abelian
groups, is already very powerful and has applications basically wherever abelian
groups come up. Somewhat less obvious but no less powerful are the applications
when R = k[x].
Recall that a k[x]-module is a k-vector space V together with a linear map A :
V V , and that V is finitely generated and torsion iff it is finite-dimensional.
In this case, we can choose a basis for V and think of A as a matrix. If we
identify the entire module with just the matrix, we have that two nn. matrices
correspond to isomorphic modules iff they are conjugate
If V = W1 W2 as a k[x]-module, then the matrix A of V is the direct
sum ofthe matrices
 Bi of the Wi . The direct sum matrix is the block diagonal
B1 0
matrix .
0 B2
Lets look at the simplest case, which is when k is algebraically closed, i.e.
every polynomial over k can be factored into linear polynomials. The Funda-
mental Theorem of Algebra says that the complex numbers C are algebraically
closed, so you can pretend k = C if youre not comfortable with an arbitrary
algebraically closed field.
As we noted earlier, in this case the primes of k[x] are exactly the linear
polynomials x a for a k (we can assume they are monic since multiplying by
a unit does not change things). The classification theorem then says that every
finitely generated torsion module is a direct sum of modules V = k[x]/((xa)n ).
What do the matrices for these modules look like? Well, a basis for V is given
by {1, x a, (x a)2 , . . . , (x a)n1 }. In this basis, it is easy to check that the
matrix looks like:
a 0 0 0 0 0
1 a 0 0 0 0

0 1 a 0 0 0

0 0 1 a 0 0

.. .. .. .. . . .. ..
. . . .
. . .
0 0 0 0 a 0
0 0 0 0 1 a
A matrix of this form is called a Jordan block. Note in particular that every
Jordan block is lower triangular.10 A matrix is said to be in Jordan normal
form if it is a direct sum of Jordan blocks. The classification theorem for
finitely generated torsion k[x]-modules thus gives:
Corollary 4.1. Let k be an algebraically closed field. Then every square matrix
over k is conjugate to a Jordan normal form matrix. That is, every linear
transformation of a finite-dimensional k-vector space is in Jordan normal form
with respect to some basis. Two matrices are conjugate iff their Jordan normal
forms are the same up to permuting the Jordan blocks.
10 By convention, the order of the basis is often reversed, so that Jordan blocks are upper

triangular instead of lower triangular.

19
Corollary 4.2. Let k be an algebraically closed field. Then every linear trans-
formation of a finite-dimensional k-vector space is lower triangular11 with re-
spect to some basis.
Proof. Jordan normal forms are lower triangular.

Jordan normal forms enormously simplify much of linear algebra. For ex-
ample, the determinant of a triangular matrix is just the product of its diagonal
entries.
If (V, A) k[x]/((x ai )ni ), the polynomials (x ai )ni are called
L
= Q the el-
ementary divisors of the linear transformation A on V . The product (xai )ni
is called the characteristic polynomial A (x). Note that if we were working over
Z instead of k[x], the product of the elementary divisors of a finitely generated
torsion module would be exactly its order as a group. Thus in some sense we
could consider the characteristic polynomial to measure the size of (V, A). For
example, it is easy to see that deg(A ) = dim V .

Proposition 4.3. Let k be algebraically closed and A be a matrix over k. The


characteristic polynomial is A (x) = det(xI A), where xI A is a matrix with
entries in k[x].
Proof. The matrix xI A over the ring k[x] is the same as the matrix of A,
except that its diagonal entries are x ai instead of ai . Its determinant is thus
(x ai )ni = A (x).
Q

Recall that an eigenvalue of a linear map A is a scalar k such that I A


is not invertible, i.e. there exists a nonzero v V such that Av = v. Note that
the eigenvalues of A are exactly the roots of the polynomial A (x): I A is
not invertible iff 0 = det(I A) = A (). Q
What are the roots of A ? Since A = i (xi ai )ni for (xi ai )ni the
elementary divisors of A, the roots are the ai . This is also easy to see quite
concretely from the Jordan normal form: if A is in Jordan normal form and
Av = v, explicit computations with a basis fairly easily show that is one of
the ai .
The following theorem is very trivial using Jordan normal forms.

Theorem 4.4 (Cayley-Hamilton over Algebraically Closed Fields). Let A be a


square matrix over an algebraically closed field k. Then A (A) = 0.
Proof. We may assume A is in Jordan normal form, and we can split it into
Jordan blocks. But on a (x ai )ni Jordan block, it is easy to explicitly compute
that (A ai I)ni = 0. Alternatively, note that A represents multiplication by x
on the module k[x]/((x ai )ni ), so Q
clearly (A ai I)ni represents multiplication
by 0. In any case, we get A (A) = i (A ai I)ni = 0.
All of this is well and good, but what if our field is not algebraically closed?
Then irreducible polynomials are not always linear, so we do not have Jordan
11 Or, as in the previous footnote, upper triangular

20
normal forms. However, we can embed k in a larger algebraically closed field K,
and then often we can answer questions over k by working over K. For doing
this, the invariant factors of Exercise 3.14 are often useful.
Proposition 4.5. Let A and B be matrices over a field k and let K k be a
larger field. Then if A and B are conjugate over K, they are conjugate over k.
Proof. For any degree n polynomial f over k, let A(f ) be the matrix for the
module k[x]/(f ) with respect to the basis {1, x, . . . , xn1 } (cf. Exercise 3.12).
Note that A(f ) is also the matrix for K[x]/(f ) with respect to the same basis.
Now Exercise 3.14 says that there are unique monic polynomials {fi } such that
fi+1 |fi and A is conjugate to the block diagonal matrix

A(f1 ) 0 0
0 A(f2 ) 0
D= .

. . ..
.. .. .. .
0 0 A(fn )
It follows that A and B are conjugate iff they have the same invariant
factors fi . Also, the invariant factors of A are the same whether we think of
it as a matrix over k or over K. Indeed, A is conjugate to D over k and hence
also over K, and by uniqueness of D, the invariant factors are the same over
both fields.
Now suppose A and B are conjugate over K. Then they have the same
invariant factors over K. But these are the same as the invariant factors over
k, so A and B are also conjugate over k.
Proposition 4.5 is quite useful. For example, we could have two real ma-
trices which we know are diagonalizable over C and have the same (complex)
eigenvalues. This implies that they are conjugate over C, so we can conclude
that they are conjugate over R.
We can also talk about characteristic polynomials over non-algebraically
closed fields. The elementary divisors (i.e., the torsion coefficients of (V, A) as a
k[x]-module) are no longer necessarily powers of linear polynomials, but we can
still define A as the product of As elementary divisors. Note that it is easy to
see that the product of the invariant factors is the same as the product of the
elementary divisors (since the elementary divisors are just the irreducible-power
factors of the invariant factors). Since the invariant factors are invariant under
extending the field, so is the product of the elementary divisors, i.e. A . We
thus obtain:
Proposition 4.6. Let k be any field and A be a matrix over k. The character-
istic polynomial is A (x) = det(xI A), where xI A is a matrix with entries
in k[x].
Proof. Let K be an algebraically closed field containing k. By the discussion
above, we can compute A over K instead of over k. However, det(xI A)
clearly also doesnt depend on what field were working over. Thus the result
follows by Proposition 4.3.

21
This could also be proven by explicitly writing down the matrices A(f ) and
computing that det(xI A(f )) = f (x).
By a similar method, we can prove the Cayley-Hamilton Theorem over any
field.
Theorem 4.7 (Cayley-Hamilton). Let A be a square matrix over a field k.
Then A (A) = 0.
Proof. Let K be an algebraically closed field containing k. Then by the dis-
cussion above, A is the same over k and over K. By Theorem 4.4, we have
A (A) = 0 over K, and hence also over k.
Lets look at some more concrete applications of all this theory.
Theorem 4.8. Diagonalizable matrices are dense in Mn (C). That is, for any
complex matrix A and any  > 0, there is a diagonalizable matrix B such that
every entry of B A is smaller than .
Proof. Since conjugation by any invertible matrix is continuous, we may assume
A is in Jordan normal form. By modifying A by less than  and only on the
diagonal, we can obtain a triangular matrix QB all of whose diagonal entries are
distinct; call them bi . But then B (x) = (x bi ), and it follows that Bs
elementary divisors are x bi so its Jordan normal form is diagonal.
Theorem 4.8 is sometimes useful when trying to prove a property of a con-
tinuous function of matrices, since it then suffices to check that property for
diagonalizable matrices.
As another application, lets classify all 22 real matricesup to conjugation.

a 0
Over C, the only possible Jordan normal forms are Dab = and Ta =
  0 b
a 1
. If Dab is conjugate to a real matrix, its trace a + b and determinant
0 a
ab must be real, and it easily follows that a and b must be either be conjugate
complex numbers or both be real. On the other hand, for any z= x + iy  C,
x y
there does indeed exist a real matrix conjugate to Dzz, namely .
y x
If Ta is conjugate to a real matrix, similar considerations show that a R.
Thus we have the following result:

 real 2 2 matrixis conjugate to exactly one matrix of


Theorem 4.9. Every
x 0 x y x 1
the form , , or (for x, y R).
0 y y x 0 x

4.1 Exercises
Exercise 4.10. Show that the matrices

0 1 0
A = 0 0 1
1 0 0

22
and
1 1 0
B = 1 0 0
0 3 1
are conjugate over Q. (Hint: Compute their characteristic polynomials to show
they have the same Jordan normal form over C (or to just show they have the
same elementary divisors over Q)).
   
1 1 1 2
Exercise 4.11. Show that the matrices and are conjugate
0 1 0 1
over Q but not over Z. Thus Proposition 4.5 does not hold for arbitrary rings.
Exercise 4.12. Show that a matrix (over an arbitrary field) is diagonalizable
iff its elementary divisors all have degree 1.
Exercise 4.13 (Cayley-Hamilton over an arbitrary ring). Let A be a square
matrix over a ring R and define A (x) = det(xI A).

(a): Assume R is a domain. Show that A (A) = 0. (Hint: Embed R in its field
of fractions and use Theorem 4.7.
(b): Show that A (A) = 0 even if R is not a domain. (Hint: Show that it
suffices to prove this when R = Z[xij ] and the entries of A are the variables
xij .)

23

Vous aimerez peut-être aussi