Noetherian Algebras
Noetherian Algebras
Noetherian Algebras
1. Basic Theory
Dually, we have the descending chain condition and the minimum condition;
these are equivalent to each other.
1.4. Definition. An R-module satisfying (i), (ii), (iii) of Lemma 1.3 is Noetherian.
The ring R is left Noetherian if it is Noetherian as a left R-module.
An R-module satisfying the descending chain condition is said to be Artinian.
The ring R is left Artinian if it is Artinian as a left R-module.
We have similar definitions on the right hand side. Note that if the ring is
commutative, there is no difference between left and right.
Examples. (1) Any division ring D is right and left Noetherian (as well as
right and left Artinian)
(2) Any principal ideal domain R is Noetherian - for example, R = Z or R =
Q[x].
Proof. Suppose each Mi is Noetherian. By an easy induction using (1.5), the di-
rect sum N = M1 M2 Mn is Noetherian. Now there is an R-module
homomorphism f : N M such that f (m1 , . . . , mn ) = m1 + . . . + mn . Since
M = M1 + M2 + . . . + Mn , f is onto so M
= N/ ker f is Noetherian by (1.5).
3
Corollary. Let R be a left Noetherian ring. Then any finitely generated left R-
module is left Noetherian.
Proof. Any finitely generated module is a sum of cyclic modules. The result follows
from (1.2) and (1.6).
1.7. Proposition. Let R be a left Noetherian ring and let S be a ring containing
R such that S is a finitely generated left R-module under left multiplication by R.
Then S is Noetherian.
Proof. By Corollary 1.6, S is Noetherian as a left R module. Since any left ideal
of S is necessarily an R-submodule, S satisfies the ACC on left ideals. Hence S is
Noetherian.
Examples. (1) Any (associative) k-algebra R becomes a Lie algebra under the
commutator bracket [x, y] = xy yx.
(2) gln (k), the set of all n n matrices over k with the commutator bracket.
(3) sln (k), the set of traceless n n matrices over k with commutator bracket.
(4) If V is any vector space, we can define the trivial bracket [x, y] = 0 for all
x, y V . This is the abelian Lie algebra.
Definition. The universal enveloping algebra U(g) of the Lie algebra g is defined
to be
U(g) := khx1 , . . . xn i/I
where {x1 , x2 , . . . , xn } is a basis for g and I is the ideal of khx1 , . . . xn i generated
by the set {xi xj xj xi [xi , xj ], 1 i, j n}.
For example, if g is abelian, then U(g) is just the polynomial algebra k[x1 , . . . , xn ].
Proposition. U(g) is left (and right) Noetherian for any finite dimensional Lie
algebra g.
Now consider the polynomial algebra A = k[x1 , . . . , xn ] and the k-linear maps
xbi : A A and xi : A A
f
f 7 xi f f 7 xi
,
xi xj xj xi 1 i, j n,
y i yj y j yi 1 i, j n,
yi xi xi yi 1 1 i n,
xi yj yj xi i 6= j.
When n = 1, we obtain
Proposition. The Weyl algebras are left (and right) Noetherian for all n 1.
Definition. Let G be a group and let R be a ring. The group algebra RG consists
of formal linear combinations
X
rg g,
gG
where rg R for all g G and all but finitely many rg are zero. Addition and
multiplication is given by
X X X
( rg g) + ( sg g) = (rg + sg )g
gG gG gG
X X X X
( rh h)( sk k) = ( rh sk )g.
hG kG gG h,kG
hk=g
6
0 0 1
1 / G1 / G2 / G3 = G
1 0 Z 1 Z Z
where G1 = 0 1 0 and G2 = 0 1 0 .
0 0 1 0 0 1
Proof. Omitted.
Theorem (McConnell, 1968). Let S be a ring, R a left Noetherian subring and let
x S.
(1) If R + xR = R + Rx and S = hR, xi, then S is left Noetherian.
(2) Suppose Aut(R) is such that rx = xr for all r R. If S = hR, xi,
then S is left Noetherian.
(3) Suppose x is a unit in S such that x1 Rx = R. If S = hR, x, x1 i, then S
is left Noetherian.
Consequences:
(a) The set of all elements of S of the form
r0 + xr1 + . . . + xn rn , n0 ()
forms a subring of S. Since it contains both R and x and S = hR, xi, we see
that S is the ring of all such polynomials. Note that elements of S need not
be uniquely expressible in the form ().
(b) The set of polynomials of degree n, namely R + Rx + . . . + Rxn , is both a
left and a right R-submodule of S.
(c) For each r R and n 0 there exists r0 R such that r0 xn = xn r + s where
deg s < n.
Now, let I be a left ideal in S. We will show that I is finitely generated. Let
In = {rn R : there exists s I such that s = r0 + xr1 + . . . + xn rn }.
Its clear that In is closed under addition. Let r R. By part (c) above, we can
find r0 R such that r0 xn xn r has degree < n. Since I is a left ideal, r0 s I, and
r0 s r0 xn rn xn (rrn )
modulo terms of degree < n. Hence rrn In so In is a left ideal of R.
Pn Pn+1
Next, if s = i=0 xi ri I, then xs = i=1 xi ri1 I so rn In+1 . Hence
In In+1 for all n 0.
Since R is left Noetherian, the increasing chain
I0 I1 . . . In . . .
8
2. Ideal structure
This definition is not symmetrical, and it is known that left primitivity does not
imply right primitivity. Nonetheless, the attribute left is usually omitted. Note
that the annihilator I of any module M is always an ideal of R.
10
Proof. Note that this largest two-sided ideal K exists, since the sum of all two-sided
ideals contained in L is itself a two-sided ideal contained in L. Certainly I L, so
I K. Now KM = KRx Kx Lx = 0 since K is two-sided, so K I.
It will be shown in Corollary 2.7 that this definition is left-right symmetric. Note
that J(R) is the set of elements of R which annihilate every simple left R-module.
Proof. Let I be a maximal left ideal. Then P = AnnR (R/I) is primitive, so J(R)
P I by Lemma 2.4. Hence J(R) K.
Now let P = AnnR (M ) be a primitive ideal, where M is a simple R-module.
Note that P = 06=xM ann(x) is an intersection of maximal left ideals, so K P .
It follows that K J(R) as required.
Lemma. Let M be a finitely generated nonzero left R-module and let J = J(R).
Then JM is strictly contained in M .
Proposition.
J(R) = {x R : 1 axb is a unit for all a, b R} =: K.
Now let x J(R). Since J(R) is a two-sided ideal, to show that x K its
sufficient to show 1 x is a unit. Now, if R(1 x) is a proper left ideal, we can find
a maximal left ideal L containing it by Lemma 2.3. By Lemma 2.5, x J(R) L
and 1 x L so 1 L, a contradiction. Hence there exists y R such that
y(1 x) = 1.
Now, 1 y = yx J(R), so by the above argument applied to 1 y, we can find
z R such that
z(1 (1 y)) = zy = 1.
Hence zy(1 x) = 1 x = z so zy = 1 and yz = 1, meaning that z = 1 x is a
unit. as required.
This result shows that J(R) is the largest ideal A of R such that 1 A consists
entirely of units of R.
Corollary. The Jacobson radical is left-right symmetric. It follows that the inter-
section of all maximal left ideals of R is equal to the intersection of all maximal
right ideals.
Proposition. Let R be a left Noetherian ring and let I / R be a proper ideal. Then
(1) There exist primes P1 , . . . , Pn containing I such that P1 Pn I.
(2) The set of minimal primes over I is equal to the set of minimal primes in
{P1 , . . . , Pn }.
12
Proof. Suppose that (1) is false. Since R is left Noetherian, we can choose a maximal
counterexample I. Thus I contains no finite product of prime ideals containing I,
and I is maximal with respect to this property.
Claim: I is prime.
If I is not prime, we can find A, B / R such that AB I but A 6 I and B 6 I.
By maximality of I, I + A contains the product of primes P1 , . . . , Pn containing
I + A, and similarly Q1 Qm I + B for some primes Q1 , . . . , Qm containing
I + B. Hence
P1 Pn Q1 Qm (I + A)(I + B) I 2 + AI + IB + AB I,
so I itself contains a finite product of primes containing it. This contradicts the
definition of I, so in fact I is prime.
Thus we have a contradiction, and (1) follows.
Hence we have a finite set of primes P1 , . . . , Pn containing I such that P1 Pn
I. Let {X1 , . . . Xm } be the distinct minimal primes of {P1 , . . . , Pn }. Thus each Pj
contains some Xij so I contains some product of the Xk s, possibly with repetition:
Xi1 Xin P1 Pn I.
Now, suppose Q is any prime containing I. Then Xi1 Xi2 Xin I Q which
forces Xij Q for some j. If Q is a minimal prime over I, Q must equal Xij .
Finally, we show that each Xk is a minimal prime over I. If I Q Xk then
Xj Q Xk for some j by the above. But the Xs are minimal in {P1 , . . . , Pn },
so Xj = Q = Xk and (2) follows.
Definition. The prime radical N (R) of R is the intersection of all prime ideals of
R. R is semiprime if N (R) = 0. An ideal I / R is semiprime if its an intersection
of some collection of prime ideals, or equivalently, if N (R/I) = 0.
Proposition. Let R be a left Noetherian ring with nilradical N = N (R) and let
P1 , . . . , Pn be the minimal primes of R. Then
(1) N = P1 . . . Pn .
(2) N is nilpotent.
Proof. Part (1) is clear. By the proof of Proposition 2.9, 0 contains a product of k
of the Pi s, possibly with repetition, so N k = 0, as required for (2).
Proposition. Let R be a commutative ring. Then the set K of all nilpotent ele-
ments of R is an ideal and equals N (R).
13
contradicting P S.
Hence x
/ K implies x / P for some prime ideal P , whence x / N (R).
3. Artinian rings
3.1. Recall that a right Artinian ring R is one which is satisfies the DCC as a
right module.
Proposition. Suppose S be a ring containing a right Artinian ring R such that S
is a finitely generated right R-module under right multiplication by R. Then S is
right Artinian.
Proof. This can be proved in exactly the same way as Proposition 1.7. In fact, the
Artinian analogues of Propositions 1.5 and 1.6 are also valid.
Examples. The following rings are all right Artinian:
(1) The group algebra kG of a finite group G over a field k.
(2) The matrix ring Mn (D), where D is a division ring.
(3) Any finite ring, for example Z/mZ, m 6= 0.
(4) k[x]/(x2 ), k a field.
The structure of a right Artinian ring is quite well understood. The following is
a summary of whats known:
Theorem. Let J = J(R) be the Jacobson radical of the right Artinian ring R.
(1) (Artin-Wedderburn) R/J is isomorphic to a finite direct sum of matrix
rings over division rings D1 , . . . , Dk :
R/J
= Mn1 (D1 ) Mn2 (D2 ) . . . Mnk (Dk )
(2) J is nilpotent and J = N (R).
(3) (Hopkins) R is right Noetherian.
3.2. Semisimple modules.
Definition. A right R-module is said to be semisimple or completely reducible if
M is a direct sum of simple submodules.
Lemma. Let M be a semisimple R-module. The following are equivalent:
(1) M is Noetherian.
(2) M is Artinian.
(3) M is a direct sum of finitely many simple modules.
Proof. By Proposition 1.6 and its Artinian analogue, (3) implies (1) and (2). If M
is not a finite direct sum, we can find a submodule N of M which is a countably
infinite direct sum N = M1 M2 . . . of nonzero submodules Mi . Now the chains
M1 < M1 M2 < M1 M2 M3 < . . .
and
M1 M2 M3 > M2 M3 > . . .
show that M is not Noetherian nor Artinian.
15
3.4. Proposition. Suppose R is semiprimitive and right Artinian. Then the right
R-module RR is semisimple.
Proof. This is best seen by writing elements of M as column vectors and thinking
of R-module endomorphisms acting by matrix multiplication on the left of these
column vectors.
Let j : Mj , M and i : M Mi be the canonical injections and projections.
Formally, we can define a map : EndR (M ) S by setting the (i, j) element of
(f ) to be the composition
j f
Mj M M i Mi ;
16
3.7. Artin-Wedderburn.
Theorem. Let R be a semiprimitive right Artinian ring. Then R is isomorphic to
a direct sum of matrix rings over some division rings D1 , . . . , Dk :
R
= Mn1 (D1 ) Mn2 (D2 ) . . . Mnk (Dk )
Proof. Consider : R HomR (RR , RR ) given by (x)(y) = xy. This is a ring
homomorphism. If (x) = 0, (x)(1) = x = 0, so is an injection. Also, if f : R
R is a right module map, then f (r) = f (1)r = (f (1))(r) so that f = (f (1)) and
is an isomorphism.
By Proposition 3.4, we can write R as a direct sum of simple right R-modules.
Grouping these together, we may write
RR n
= An1 1 An2 2 . . . Ak k
where the Ai are pairwise nonisomorphic simple modules. Applying Proposition
3.5, we obtain
EndR (An1 1 ) 0 0
0 EndR (An2 2 ) 0
R= EndR (RR )
=
nk
0 0 EndR
(Ak )
Mn1 (D1 ) 0 0
0 Mn2 (D2 ) 0
=
0 0 Mnk (Dk )
where each Di = EndR (Ai ) is a division ring, by Schurs Lemma (3.6).
17
As an exercise, the reader should check directly that a ring of the form
Mn1 (D1 ) Mn2 (D2 ) . . . Mnk (Dk )
is semisimple Artinian.
Note that if we were dealing with left R-modules, we would obtain that the op-
posite ring Rop is isomorphic to the direct sum of matrix rings. Since Mn (Dop )op
=
Mn (D) and the opposite ring of a division ring is also a division ring, the theorem
holds with right replaced by left. This justifies the term semisimple Artinian
ring, without reference to side.
3.8. Before we can prove Hopkins Theorem, we will need some results on semisim-
ple modules.
Since S contains singletons {} for all A, S is nonempty. Its easy to check that
unions of chains in S are again in S, so by Zorns Lemma, S contains a maximal
element A0 .
Now, if X = A0 M 6= M , we can find M such that M ( X. Hence
M X = 0 and the sum X + M is direct. Therefore A0 {} S, contradicting
the maximality of A0 . Hence X = M is a direct sum of simple modules, so M is
semisimple.
3.9. Lemma. Suppose the ring R is semisimple Artinian. Then every R-module
is semisimple.
Proof. Since every primitive ideal is prime by Lemma 2.8(2), N (R) J(R) in
any ring. On the other hand if R is right Artinian then J(R) is nilpotent by
Proposition 3.10, so J(R) P for any prime ideal P by Lemma 2.8(1). Hence
J(R) N (R).
3.11. Hopkins Theorem. Let R be a right Artinian ring. Then R is right Noe-
therian.
Its tempting to think that the left-right symmetry is so strong that every right
Artinian ring is left Artinian. This is not the case, however.
19
4. Commutative Rings
N (R) = J(R).
Proof. Its clear that each ma is a maximal ideal, being of codimension 1. Now if I
is maximal, then the R/I = k so Xi + I 7 ai for some ai k. Hence Xi ai I for
all i = 1, . . . , n so ma I. Since the former ideal is maximal, the result follows.
I(Z(A)) = A.
4.4. Proof of Theorem 4.2. We will only deal with the case when k is an un-
countable field in this course. For the general case, see Atiyah and Macdonald, (7.8)
and (7.9).
Suppose R is not algebraic. Then we can find R such that k() = k(t),
the field of fractions of the polynomial ring k[t]. Choose an uncountable subset
{ci : i I} of k and suppose 1 , . . . , m k are such that
1 2 m
+ + ... + = 0.
ci1 ci2 cim
Clearing denominators gives
1 ( ci2 )( ci3 ) ( cim ) + . . . + m ( ci1 )( ci2 ) ( cim1 ) = 0.
Substituting 1 = ci1 into this equation gives 1 (ci1 ci2 )(ci1 ci3 ) (ci1 cim ) = 0
so 1 = 0 and similarly j = 0 for all j = 1, . . . , m.
1
Hence { c i
k() : i I} is an uncountable linearly independent subset of
R. But R is a quotient of the polynomial ring k[X1 , . . . , Xn ] and as such has a
countable spanning set, consisting of the images of the monomials in R. This is a
contradiction, proving (1).
Now, a finitely generated field extension of k which is algebraic must necessarily
be finite dimensional. If k = k then R is a finite field extension of k so R = k as
required for (2).
4.5. Lemma. Let k be a field and let R be a finite dimensional kalgebra which
is a domain. Then R is a field.
4.6. Proof of Theorem 4.1. As we have observed in the proof of Corollary 3.10,
N (R) J(R) for any ring R. Suppose a / N (R) so a is not nilpotent. By Sheet 1
Exercise 6(2), 1 ax is not a unit in the polynomial ring R[x]. By Lemma 2.3 we
can find a maximal ideal M of R[x] containing 1 ax.
Since R is a finitely generated kalgebra, so is R[x]/M . By Theorem 4.2, R[x]/M
is algebraic over k and is hence finite dimensional over k. Now, we have a klinear
injection R/(R M ) , R[x]/M , so R/(R M ) is a finite dimensional kalgebra
21
4.7. Associated primes. From now on well be dealing with a commutative Noe-
therian ring R and finitely generated Rmodules M .
Example. Let R = Z and let M be a finite abelian group. Then M has an element
of order p for all primes p dividing |M |, so
so Ass(M ) is finite.
This result shows that any finitely generated Rmodule M can be thought of
as being built up from finitely many modules of the form R/Pi for some primes Pi .
Its enough to show that P1 Ass(R). Let M = P2s2 Pnsn if n > 1 and M = R
otherwise.
By Theorem 4.10(1) we can choose a chain
where Mi /Mi1 = Qi for some primes Q1 , . . . , Qt . Note that P1s1 kills each
Mi /Mi1 so P1s1 Qi and hence P1 Qi for all i = 1, . . . , t by the primality
of Qi .
Also, Q1 Q2 Qt M = 0 implies Q1 Q2 Qt P2s2 Pnsn P1 . Since P1 is prime
and Pi ( P1 for all i 2, Qk P1 for some k, so in fact Qk = P1 for some k.
Pick k 1 least such that Qk P1 . Hence if k > 1 then Q1 Qk1 ( P1 . If
k = 1 let r = 1 and if k > 1 choose r Q1 Qk1 \P1 . Since Ann(Mk /Mk1 ) =
Qk = P1 and r / P1 we can find x Mk such that rx / Mk1 .
Claim: P1 = ann(rx). Note that rP1 x Q1 Q2 Qk1 Qk x = 0 so P1
ann(rx). If srx = 0 then s(rx + Mk1 ) = 0. We chose x so that rx / Mk1 , and
hence s P1 because Mk /Mk1 = R/Q k =
R/P 1 is a domain. Hence ann(rx) =
P1 Ass(M ) as required.
Proof. Since any minimal prime over Q must be in Ass(R/Q) by Proposition 4.11,
we see that R/Q has precisely one minimal prime, namely P/Q. Hence N (R/Q) =
Q/Q = P/Q so P = Q.
23
For (2), note that P/Q is the prime radical of R/Q and hence is the only prime
ideal of R/Q, being maximal. Hence Ass(R/Q) must be {P }, being nonempty by
Lemma 4.8, so Q is P primary.
Examples. (1) The primary ideals of Z are the prime powers and 0.
(2) Let R = k[x, y], Q = (x, y 2 ). Then Q = (x, y) =: P is maximal so Q is
primary. However, Q is not a prime power since P 2 ( Q ( P .
4.14. Theorem.
(1) (Existence) Every proper ideal has a minimal primary decomposition.
(2) (Uniqueness) Let I = Q1 . . . Qk be a minimal primary decomposition
of I, with Qi a Pi primary ideal. Then
Ass(R/I) = {P1 , . . . , Pn }.
Proof. Omitted.
24
5. Commutative localisation
5.3. Construction.
(r, s) (u, v)
if and only if
rv = us.
This is clearly reflexive and symmetric. If (r, s) (u, v) and (u, v) (a, b) then
rv = us and ub = av, so (rb as)v = rvb asv = ubs avs = 0 so rb as
ass(S) = 0 and (r, s) (a, b). Hence is an equivalence relation.
Let RS = R S/ as a set. Write r/s for the equivalence class of (r, s) in RS
and define addition and multiplication on RS by
r/s + u/v = (rv + us)/(sv)
r/s u/v = (ru)/(sv)
r/s + u/v = (rv + us)/(sv) and (r0 /s0 ) + (u/v) = (r0 v + us0 )/(s0 v),
but (rv + us)(s0 v) = (rs0 )v 2 + uss0 v = (r0 s)v 2 + us0 sv = (r0 v + us0 )(sv), so
5.4. The general case. If ass(S) is not necessarily zero, let R = R/ ass(S) and
let S denote the image of S in R. Then ass(S) = 0 so we can form the localisation
RS , together with a ring homomorphism : R RS as constructed above. Let
: R RS
Note that (r)(s)1 = (u)(v)1 iff (rv us) = 0 iff rvt = ust for some
t S. So we could have constructed RS by imposing the equivalence relation on
R S instead, where
5.5. Examples.
(1) If R is an integral domain then S = R\{0} is a m.s. set. Then RS is just
the field of fractions of R. This is a special case of (2).
(2) If R is arbitrary and P is a prime ideal, then S = R\P is m.c. The
localisation RS is usually denoted by RP and is called
The ring Z(p) from Example 1.8 (4) is a special case of this construction.
(3) For any x R, the set {1, x, x2 , . . .} is m.c., and the localisation is usually
denoted by Rx .
(4) If 0 S then ass(S) = R so RS = 0.
(5) If I / R then S = 1 + I is m.c.
Of these, Example (2) is the most important.
MS = {m/s : m M, s S}
6. Noncommutative localisation
Now we drop the requirement that R be commutative and try to generalize the
results established so far about localisation.
As before, ass(S) has to be zero in any ring in which S consists of units, but we
have to be careful about what kind of annihilators we take.
We are interested in right fractions, i.e. those of the form rs1 . Hence the term
right localisation. One can play this game on the left, too. Note that if RS exists,
then ass(S) has to be a two-sided ideal - this is not true in general.
Hence (rb sa) = 0 so there exists t S such that r(bt) = s(at). Note that bt S
as S is m.c.
Definition. A m.c. set S is a right Ore set, or satisfies the right Ore condition if
and only if
Assuming S is a right Ore set, we may factor out by the two-sided ideal ass(S)
and obtain a ring R such that the image S of S in R consists of left regular elements.
If the right localisation RS exists, then R is a subring of RS and (S) consists of
units. Hence every element of S must actually be regular (and not just left regular).
This is another necessary condition on the set S for RS to exist.
Theorem (Ore, 1930). Let S be a m.c. subset of the ring R. Then the right
localisation RS exists if and only if
(1) S is a right Ore set, and
(2) S consists of regular elements in R = R/ ass(S).
Proof. We have shown in the above discussions the necessity of these conditions.
Assuming (1) and (2), it remains to construct the ring RS satisfying the conditions
of Definition 6.1.
By Lemma 6.2, ass(S) is a two-sided ideal in R. Its easy to check that the
image S also satisfies the right Ore condition. As before, by passing to the quotient
R = R/ ass S and using condition (2), we may assume that S is a right Ore set in
R consisting of regular elements.
We can now construct RS as a set of equivalence classes in R S under a
certain equivalence relation, in a very similar spirit to the construction of RS in the
commutative case (5.3). The construction is very tedious and is non-examinable.
See handout.
Definition. A m.c. set S is said to be a right divisor set if conditions (1), (2) are
satisfied.
By Theorem 6.3, its clear that if R is a right Ore domain, then the localisation
RS exists. Moreover, every nonzero element r/s RS is invertible, so RS is a
division ring, known as the division ring of fractions of R.
Examples. Let k be a field. The following rings are all right Ore domains and
hence have division rings of fractions:
(1) The Weyl algebras An (k).
(2) kG, where G is a torsionfree polycyclic group.
(3) U(g) where g is a f.d. Lie algebra over k.
(4) The Iwasawa algebra G if G is a torsionfree compact padic Lie group.
6.5. Modules.
Definition. Let S be a right divisor set in R and let M be a right Rmodule. The
localisation of M at S is defined to be the set of equivalence classes
MS = {m/s : m M, s S}
in M S under the equivalence relation given by
(m, s) (n, t) if and only if mt0 u = ns0 u for some u S,
where s0 , t0 R are such that st0 = ts0 S. The S-torsion submodule of M is
defined to be
assM (S) = {m M : ms = 0 for some s S}.
Theorem (Goldie, 1958, 1960). The ring R has a classical right ring of quotients
which is semisimple Artinian if and only if
(1) N (R) = 0,
(2) R contains no infinite direct sum of right ideals, and
(3) R satisfies the ACC on right annihilators.
We will only prove (), the other direction being easier and less interesting. For
the remainder of this chapter, assume that R is a ring satisfying conditions (1), (2)
and (3) of Theorem 6.6. Let S denote the set of all regular elements of R.
There is a connection between right regular elements and essential right ideals.
Let a = rann(a) for any a R. Note that x = 0 iff x is right regular and x = R
iff x = 0.
Proof. By Theorem 6.3 its sufficient to show that S is a right Ore set. Let r R
and s S; since s is regular, sR is essential in R by Proposition 6.7(). Let
T = {a R : ra sR}.
T is a right ideal in R; well show that T is essential. Suppose I /r R is nonzero. If
rI = 0 then I T so I T 6= 0. Otherwise sR rI 6= 0 since sR is essential. If
0 6= rx sR rI then 0 6= x I T so I T 6= 0 as claimed.
By Proposition 6.7(), we can find a T S. Then ra = sb for some b R
and S is a right Ore set.
Lemma. Let J /r R.
(a) If J is nil then J = 0.
(b) If J 6= 0 then there exists y J such that yR y = 0.
Proof. (a) Suppose for a contradiction that there exists nonzero a J. Then
aR J is nil, and hence Ra is also nil, because (ax)n = 0 (xa)n+1 = 0.
Let S = {b : 0 6= b Ra}. By (3), we can pick 0 6= b Ra so that b is maximal
in S. Since N (R) = 0 by (1), bRb 6= 0, so we can find x R such that bxb 6= 0.
Since Ra is nil and 0 6= xb Ra, there exists m 2 such that (xb)m1 6= 0 but
(xb)m = 0. Now ((xb)m1 ) S and b ((xb)m1 ) as by = 0 (xb)m1 y = 0.
By maximality of b , xb ((xb)m1 ) = b so bxb = 0, a contradiction.
(b) By (a), J is not nil. Choose z J which is not nilpotent. Now, the chain
z (z 2 ) (z 3 )
34
Draw a picture!
Note that the axioms imply that R0 is a subring of R and that each Ri is a left
and right R0 module. Note also that iZ Ri is always an ideal in R.
for some additive subgroups Si S, satisfying Si .Sj Si+j for all i, j Z and
1 S0 . Si is called the ith homogeneous component of S, and an element s S
is homogeneous iff it lies in some Si .
A graded right ideal J /r S is a right ideal of the form iZ Ji with Ji Si .
Definition. Let R be a filtered ring with filtration (Ri )iZ . Define the abelian group
M
gr R = Ri /Ri1 .
iZ
(r) = r + Ri1 gr R.
xi xj xj xi k{1, x1 , . . . , xn }
for all i, j = 1, . . . , n.
: k[X1 , . . . , Xn ] gr R
meaning that (xi ) and (xj ) commute. If one of xi , xj lies in k then this is also
true. Hence the kalgebra map exists.
To show that is surjective, its sufficient to show that u + Rt1 lies in Im for
any u Rt \Rt1 . Write u as a sum of words of length at most t in the generators
{x1 , . . . , xn }. Since were interested in u + Rt1 , we can assume that each word
actually has length t.
Pm (j) (j)
Writing u = j=1 j xi(j) xi(j) xi(j) for some i1 , . . . , it {1, . . . , n}, we have
1 2 t
m
X
u + Rt1 = j (Xi(j) )(Xi(j) ) (Xi(j) ) Im .
1 2 t
j=1
Example. Let h2n+1 be the Heisenberg Lie algebra of dimension 2n + 1 with gen-
erators {x1 , . . . , xn , y1 , . . . , yn , z} and relations [yi , xi ] = z for i = 1, . . . , n, all the
other brackets being zero. Then there is a surjective map of kalgebras
: U(h2n+1 ) An (k)
h2n+1
= gln+2 (k).
0 0 0 k
0 0 0 0
39
Definition. Let R be a filtered ring with filtration (Ri )iZ and let M be a right
Rmodule. A filtration on M is a set (Mi )iZ of additive subgroups of M satisfying
Mi Mi+1 for all i Z,
Mi .Rj Mi+j for all i, j Z,
iZ Mi = M .
Positive, negative and separated filtrations are defined analogously to (7.1). Fil-
tered left modules are defined similarly.
Definition. Let R be a filtered ring and let M be a filtered right Rmodule with
filtration (Mi )iZ . Define the abelian group
M
gr M = Mi /Mi1 .
iZ
Proof. If I J are right ideals of R, equip I and J with the subspace filtrations
and J/I with the quotient filtration. Proposition 7.7 shows that gr I and gr J are
right ideals of gr R, and moreover that gr I gr J and gr(J/I)
= gr J/ gr I.
Suppose
I1 I2 I3 . . .
is an increasing chain of right ideals of R. Applying gr we obtain an increasing
chain of right ideals of gr R
gr I1 gr I2 gr I3 . . .
which stops because gr R is right Noetherian. So its sufficient to prove that if
gr I = gr J for right ideals I J then I = J.
41
Let (Ri )iZ be the filtration on R. If I < J then I Ri < J Ri for some i.
Since the filtration on R is positive, we can choose a least such i. Choose some
x (J Ri )\I. Then x / Ri1 as I Ri1 = J Ri1 by minimality of i. Now
(x) = x + Ri1 gr J = gr I
so there exists y I such that (x) = (y) = y + Ri1 . Hence x y Ri1 . But
x y J since y I J, so x y J Ri1 = I Ri1 . Hence x y I so
x I, a contradiction. The result follows.
Corollary. Any almost commutative algebra R is right and left Noetherian. This
includes An (k) and U(g).
7.9. Topologies. Assume until the end of this chapter that R is a negatively fil-
tered ring and M is a negatively filtered right Rmodule with filtration (Mi ).
We can define a topology on M called the filtration topology, by choosing the open
subsets to be unions of sets of the form m + Mi . Then as required, the collection
of open sets is closed under arbitrary unions, and also under finite intersections -
since the intersection of m1 + Mi(1) , . . . , mn + Mi(n) is a union of sets of the form
m + Mi where i = min{i(1), . . . , i(n)}. Another way of phrasing this is to say that
the cosets m + Mi form a basis for the filtration topology.
If the filtration is separated, we can define a metric on M : fix a real number
c > 1 and set
d(x, y) = inf{ck : x y Mk }.
Note that we need Mi = 0 to ensure that d(x, y) = 0 x = y. It can be checked
that the topology we get on M from the metric is the filtration topology, so the
choice of the constant c is irrelevant.
Two filtrations (Mi ) and (Mi0 ) give the same filtration topology on M if and
0
only if for all i there exist s(i) such that Ms(i) Mi and t(i) such that Mt(i) Mi0 .
The filtrations are then said to be topologically equivalent.
Definition. A Cauchy sequence on M is a sequence (xn ) n=0 such that for any
open set U containing 0, there exists c(U ) N such that xn xm U whenever
n, m c(U ). Two Cauchy sequences (xn ) and (yn ) are said to be equivalent if
xn yn 0. Equivalently, given any open set U containing 0, there exists N (U ) Z
such that xn yn U whenever n N (U ).
42
Exercise. Check that this gives an equivalence relation on the set of all Cauchy
sequences on M .
Let [xn ] denote the equivalence class of the Cauchy sequence (xn ) under this
equivalence relation, and let M
c denote the set of all such equivalence classes. The
operations
[xn ] + [yn ] = [xn + yn ]
[xn ] . [yn ] = [xn yn ] [xn ], [yn ] R
b
turn R
b into an ring, and
turn M
c into a right Rmodule.
b There is a ring homomorphism
R R b
r 7 [r]
where [r] is the equivalence class of the constant sequence with value r. The kernel
of this map is Ri . This also enables us to view M c as a right Rmodule, and we
have an analogous map of Rmodules
M M c
m 7 [m]
ci = {[xn ] : ci N
M such that xn Mi whenever n ci }
7.11. Inverse limits. There is another way of approaching completions. Let r < s
and let rs : M/Mr M/Ms be the natural projection. Define
Y
lim M/Mi = {(yn +Mn )nZ M/Mn : rs (yr +Mr ) = ys +Ms whenever r < s.}
iZ
This becomes a right Rmodule with pointwise addition and scalar multiplica-
tion. When M = R, these operations turn lim R/Ri into a ring.
43
Lemma. M
c is isomorphic to lim M/Mi as Rmodules.
Proof. If (xn ) is a Cauchy sequence in M , then for each i Z there exists c(Mi ) N
such that xn xm mod Mi for all n, m c(Mi ). We can choose the c(Mi ) to
satisfy
Note that this doesnt depend on the choice of the numbers c(Mi ): if d(Mi ) are
such that xn xm mod Mi whenever n, m d(Mi ), then
for all i Z.
Now, if [xn ] = [yn ] M
c, then xn yn 0 as n . Hence for all i Z there
exists t(Mi ) such that xn yn Mi whenever n t(Mi ). Let c(Mi ), d(Mi ) be the
integers corresponding to [xn ] and [yn ]. Letting e(Mi ) = max(c(Mi ), d(Mi ), t(Mi )),
we have
for all i Z. Hence ([xn ]) doesnt depend on the choice of Cauchy sequence inside
the equivalence class [xn ], so is well defined.
Q
Equip each M/Mi with the discrete topology and iZ M/Mi with the (Ty-
chonov) product topology. Then the map constructed above is actually a home-
omorphism, where we give M c the filtration topology and lim M/Mi Q
iZ M/Mi
the subspace topology.
Also, in the case when M = R, is a ring isomorphism.
The advantage of inverse limits is that the elements are easy to manipulate; the
disadvantage is that everything is dependent on the filtration.
44
7.12. Examples.
(1) Zp , the ring of padic integers is the completion of R = Z with respect to
the padic filtration Ri = pi Z for i 0. Thus Zp = lim Z/pi Z.
(2) More generally, if R is any ring and I / R is a two-sided ideal, we can
complete R with respect to the Iadic filtration:
b = lim R/I i .
R
b
(3) If R = k[X1 , . . . , Xn ] and I = (X1 , . . . , Xn ) / R then R = k[[X1 , . . . , Xn ]],
the ring of power series. Here we identify a power series with the sequence
of partial sums. For example, if n = 1,
a0 + a1 X + a2 X 2 + . . . (a0 + I, a0 + a1 X + I 2 , . . .)
:R Fp
P P
gK g g 7 gK g
7.13. Proposition. Let R be a ring and I / R an ideal such that R/I is a division
ring. Then the completion R b of R with respect to the Iadic filtration is a local
ring, that is, it has a unique maximal left and right ideal.
y = 1 + x + x + x + . . . = (1 + I, 1 + x2 + I 2 , 1 + x3 + x23 + I 3 , . . .) R
2 3 b
Theorem. Let R be a complete negatively filtered ring such that gr R is right Noe-
therian. Then R is right Noetherian.
Since (y) and the si are homogeneous, we may assume that ti Rjn(i) /Rjn(i)1 ,
i = 1, . . . , m. If ti = 0 set ri1 = 0, otherwise choose ri1 Rjn(i) \Rjn(i)1 such
that ti = ri1 + Rjn(i)1 . Hence
m
X
y xi ri1 mod Rj1 .
i=1
Pm
Now let y1 = y i=1 si ri1 I Rj1 . If y1 = 0, stop; if not, repeat the
above with y1 in place of y. This gives a sequence (ri1 , ri2 , ri3 , . . .) such that rik
Rjn(i)k+1 and
Xm Xk
yk = y xi ril I Rjk
i=1 l=1
Pk
for every k 1. But now ( l=1 ril )
k=1 is a Cauchy sequence in R which converges
to some ri R because R is complete. Also, yk 0 since yk Rjk for all k.
P
Hence, letting ri = l=1 ril , we obtain
m
X m
X
y= xi ri xi R
i=1 i=1
as required.
46
8. Weyl algebras
: An Endk (A)
xi 7 (f 7 Xi f )
f
yi 7 (f 7 X i
).
We can therefore think of A as a left An module, via
8.2. Notation.
When = (1 , . . . , n ) Nn , write x = x n 1
1 xn and y = y1 yn .
1 n
Proof. Any word in the generators can be brought to a linear combination of the
x y s by pushing xs to the left, using the relations yi xi = xi yi + 1. Hence
{x y } spans An .
Next, suppose r = ,Nn x y An is zero but some coefficient is
P
r.X =
P P P
||<|| x y .X + ||||,6= x y .X + Nn ! X
= ! Nn X 6= 0
P
Corollary. Equip An with the positive filtration (An )m given in Example 7.1(3).
Then
gr An
= k[Z1 , . . . , Z2n ].
Proposition. Suppose R is a filtered ring with a separated filtration (Ri )iZ , such
that gr R is a domain. Then R is a domain also.
Proof. Suppose for a contradiction that r, s R are nonzero but rs = 0. Since the
filtration is separated, there exist i, j Z such that r Ri \Ri1 and s Rj \Rj1 .
But now (r) = r + Ri1 and (s) = s + Rj1 are nonzero in gr R and
(r).(s) = rs + Ri+j1 = 0,
contradicting the assumption that gr R is a domain.
8.5. PBW Theorem. Let g be a finite dimensional kLie algebra with basis
{x1 , . . . , xn }. The enveloping algebra U(g) contains the standard monomials
x = x1 2 n
1 x2 xn
for all Nn .
Proof. Omitted.
8.6. Proposition.
(1) Z(An ) = k.
(2) An is a simple ring.
[ab, c] = a[b, c] + [a, c]b and [a, bc] = b[a, c] + [a, b]c.
Next, we prove that a, b satisfy [a, b] = 1 then [am , b] = mam1 . This is true when
m = 1 so assume inductively that m > 1 and [am1 , b] = (m 1)am2 . Then
[x y , yi ] = i xei y .
and similarly deg[r, yi ] deg r 1 for all i. Since I is a two-sided ideal [r, xi ] and
[r, yi ] must lie in I for all i. By minimality of m, these must all be zero and hence
r Z(An ) = k. Since r 6= 0 and k is a field, 1 I and hence I = An .
Note that by Lemma 8.7, D(R) is a subring, and moreover, (Di (R))iZ is a
positive filtration on D(R), provided we assume Dk (R) = 0 for negative k.
8.8. Derivations.
Proof. (1) Since R is commutative, a D0 (R) for all a R. On the other hand,
if D D0 (R) then D(a) = D(a.1) = (D.a)(1) = (a.D)(1) = a(D(1)) = D(1)(a).
[
[ as required.
Hence D = D(1)
(2) If D Der(R) and a R, then
[ Der(R)+D0 (R)
so P (ab) = aP (b)+bP (a) and P Der(R). Hence Q = P + Q(1)
as required.
D1 (R) = R + R1 + . . . + Rn .
R
P
Theorem. D(R) = Nn = An .
Proof. Omitted.
8.10. An modules.
y .f = y .X = ! 6= 0
52
Part (3) is a special case of Bernsteins Inequality. (1) and (3) are false in positive
characteristic, and only (2) remains valid.
One can also look for solutions in other spaces B, which are An modules - the
solution space will be HomAn (MP1 ,...,Pm , B).
For example, when k = R, we can take B = C (U ), the set of all infinitely
differentiable functions f : U R, for some open set U Rn . This is a left
An module, with the action given by
(xi .f )(a1 , . . . , an ) = ai f (a1 , . . . , an ) and
f
yi .f = .
ai
53
9. Dimensions
9.3. Rees rings and modules. Let R be a filtered ring with filtration (Ri )iZ ,
and let M be a filtered left Rmodule with filtration (Mi )iZ .
R i ti M j tj Mi+j ti+j
r i ti , mj tj 7 mi rj ti+j .
Note that t R
e is a central regular element, since 1 R1 always. There is a
certain amount of interplay between the Rees ring of R and the associated graded
ring gr R.
Proof. We will only prove the result for the rings, leaving the modules as an exercise.
(1). We have an isomorphism of abelian groups
i
L
e = L iZ Ri t = gr R.
M
R/t
e R
i
Ri /Ri1 =
iZ Ri1 t iZ
Proof. (1) If the graded module M f is finitely generated, it has a finite homogeneous
generating set {tk1 m1 , . . . , tks ms } say, with mj Mkj . Then the ith homogeneous
component of M f is
One consequence of this result is that any finitely generated module has some
good filtration: just take a generating set X = {x1 , . . . , xs } and set the integers ki
to be 0, so that Mi = Ri x1 + . . . + Ri xs = Ri X (the standard filtration on M ) is
good.
At first sight, this depends on the choice of good filtration (Mi ) on M . However,
let (Mi0 ) be another good filtration on M and let 0 (t) be the corresponding Samuel
polynomial. By Proposition 9.5(2), there exists c N such that
0 0
Mjc Mj Mj+c for all j Z,
58
0 0
so dim Mjc dim Mj Mj+c . Hence, for large enough j,
0 (j c) (j) 0 (j + c).
Since the behaviour of a polynomial for large enough j is determined by its leading
term, we see that dimension and multiplicity is well defined.
9.7. Examples.
(1) Let R = k[X1 , . . . , Xn ] and let (Ri ) be the usual positive filtration on R,
given by Ri = k.{X : || i}. Then dim Ri is the number of monomials
in n variables of length at most i which is well known to be
in
n+i (i + n)(i + n 1) (i + 1)
dim Ri = = = + o(in1 ).
n n! n!
So if we view R as an Rmodule by multiplication, we can read off d(R) = n
and m(N ) = 1.
(2) Let R = An with the usual filtration. Viewing R as a filtered module
over itself with the same filtration, we see that gr R = k[X1 , . . . , X2n ] by
Corollary 8.3, so d(R) = 2n and m(R) = 1.
(3) Let R = An and let M = k[X1 , . . . , Xn ] be the natural An module with
the usual filtration. Then d(M ) = n and m(M ) = 1.
(4) Let R be any almost commutative kalgebra and M a finitely generated
Rmodule. Then d(M ) = 0 if and only if dim M < .
Proof. Let (Mi ) be a good filtration on M . Give N the subspace filtration (Ni )
and M/N the quotient filtration (M/N )i . First, we prove that these filtrations are
good.
Note that Ni Mi and Mi /Ni = Mi /(N Mi ) = (Mi + N )/N . A computation
similar to the proof of Proposition 7.7 shows that
g
M f e
N M and
e f = M /N
N
as left Rmodules.
e
Now R e is left Noetherian by Lemma 9.4, so both N ^ are finitely gener-
e and M/N
ated Rmodules,
e as (Mi ) is good. Hence the filtrations on N and M/N are good
and we can use them to compute dimensions and multiplicities.
Next, we have
Nj (M/N )j Mj
dim + dim = dim
Nj1 (M/N )j1 Mj1
59
Proof. From Theorem 9.8(1), we see that d(Rm ) = d(R) = n for any m 1.
Since M is finitely generated, its a left Rmodule quotient of Rm for some m, so
d(M ) d(Rm ) = n, proving (1).
Next, pick 0 6= x I. Since R/I R/Rx, we see that d(R/I) d(R/Rx) by
Theorem 9.8(1). Suppose for a contradiction that d(R/Rx) = d(R) = n. Since
R is a domain, Rx = R as left Rmodules, so d(Rx) = d(R/Rx). But now by
Theorem 9.8(2), m(Rx) + m(R/Rx) = m(R) so m(R/Rx) = 0, a contradiction.
Hence d(R/Rx) < n as required.
Proof. (Joseph)
Pick a finite generating set X and let (Mi ) be the standard filtration on M
given by Mi = (An )i X. This filtration is good and positive. Let (t) be the
Samuel polynomial of gr M with respect to this filtration. Then for large enough i,
(i) = dim Mi by the remarks in (9.6).
Claim: dim Ri dim Homk (Mi , M2i ) = (dim Mi )(dim M2i ).
Assuming the claim is true, for large enough i, dim Ri (i)(2i). Since
dim Ri = i+2n
2n is a polynomial in i of degree 2n, (t)(2t) must be a polyno-
mial of degree 2n. But it has degree 2d(M ), so d(M ) n as required.
To prove the claim, consider the klinear map
Note that when n = 1, this says that any nonzero finitely generated An module
is infinite dimensional over k, by Example 9.7(4). This is also the content of Propo-
sition 8.10(3).