Part III Algebras
Based on lectures by C. J. B. Brookes
Notes taken by Dexter Chua
Lent 2017
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
The aim of the course is to give an introduction to algebras. The emphasis will be
on non-commutative examples that arise in representation theory (of groups and Lie
algebras) and the theory of algebraic D-modules, though you will learn something
ab out commutative algebras in passing.
Topics we discuss include:
Artinian algebras. Examples, group algebras of finite groups, crossed products.
Structure theory. Artin–Wedderburn theorem. Projective modules. Blocks.
K
0
.
No etherian algebras. Examples, quantum plane and quantum torus, differen-
tial operator algebras, enveloping algebras of finite dimensional Lie algebras.
Structure theory. Injective hulls, uniform dimension and Goldie’s theorem.
Hochschild chain and cochain complexes. Hochschild homology and cohomology.
Gerstenhab er algebras.
Deformation of algebras.
Coalgebras, bialgebras and Hopf algebras.
Pre-requisites
It will be assumed that you have attended a first course on ring theory, eg IB Groups,
Rings and Modules. Experience of other algebraic courses such as II Representation
Theory, Galois Theory or Number Fields, or III Lie algebras will be helpful but not
necessary.
Contents
0 Introduction
1 Artinian algebras
1.1 Artinian algebras
1.2 Artin–Wedderburn theorem
1.3 Crossed products
1.4 Projectives and blocks
1.5 K
0
2 Noetherian algebras
2.1 Noetherian algebras
2.2 More on A
n
(k) and U(g)
2.3 Injective modules and Goldie’s theorem
3 Hochschild homology and cohomology
3.1 Introduction
3.2 Cohomology
3.3 Star products
3.4 Gerstenhaber algebra
3.5 Hochschild homology
4 Coalgebras, bialgebras and Hopf algebras
0 Introduction
We start with the definition of an algebra. Throughout the course,
k
will be a
field.
Definition
(
k
-algebra)
.
A (unital) associative
k
-algebra is a
k
-vector space
A
together with a linear map
m
:
A A A
, called the product map, and linear
map u : k A, called the unit map, such that
The product induced by m is associative.
u(1) is the identity of the multiplication.
In particular, we don’t require the product to be commutative. We usually
write m(x y) as xy.
Example.
Let
K/k
be a finite field extension. Then
K
is a (commutative)
k-algebra.
Example.
The
n × n
matrices
M
n
(
k
) over
k
form a non-commutative
k
-algebra.
Example.
The quaternions
H
is an
R
-algebra, with an
R
-basis 1
, i, j, k
, and
multiplication given by
i
2
= j
2
= k
2
= 1, ij = k, ji = k.
This is in fact a division algebra (or skew fields), i.e. the non-zero elements have
multiplicative inverse.
Example. Let G be a finite group. Then the group algebra
kG =
n
X
λ
g
g : g G, λ
g
k
o
with the obvious multiplication induced by the group operation is a k-algebra.
These are the associative algebras underlying the representation theory of
finite groups.
Most of the time, we will just care about algebras that are finite-dimensional
as
k
-vector spaces. However, often what we need for the proofs to work is not
that the algebra is finite-dimensional, but just that it is Artinian. These algebras
are defined by some finiteness condition on the ideals.
Definition
(Ideal)
.
A left ideal of
A
is a
k
-subspace of
A
such that if
x A
and
y I
, then
xy I
. A right ideal is one where we require
yx I
instead.
An ideal is something that is both a left ideal and a right ideal.
Since the multiplication is not necessarily commutative, we have to make the
distinction between left and right things. Most of the time, we just talk about
the left case, as the other case is entirely analogous.
The definition we want is the following:
Definition
(Artinian algebra)
.
An algebra
A
is left Artinian if it satisfies the
descending chain condition (DCC ) on left ideals, i.e. if we have a descending
chain of left ideals
I
1
I
2
I
3
· · · ,
then there is some N such that I
N+m
= I
N
for all m 0.
We say an algebra is Artinian if it is both left and right Artinian.
Example. Any finite-dimensional algebra is Artinian.
The main classification theorem for Artinian algebras we will prove is the
following result:
Theorem
(Artin–Wedderburn theorem)
.
Let
A
be a left-Artinian algebra such
that the intersection of the maximal left ideals is zero. Then
A
is the direct sum
of finitely many matrix algebras over division algebras.
When we actually get to the theorem, we will rewrite this in a way that
seems a bit more natural.
One familiar application of this theorem is in representation theory. The group
algebra of a finite group is finite-dimensional, and in particular Artinian. We
will later see that Maschke’s theorem is equivalent to saying that the hypothesis
of the theorem holds. So this puts a very strong constraint on how the group
algebra looks like.
After studying Artinian rings, we’ll talk about Noetherian algebras.
Definition
(Noetherian algebra)
.
An algebra is left Noetherian if it satisfies
the ascending chain condition (ACC ) on left ideals, i.e. if
I
1
I
2
I
3
· · ·
is an ascending chain of left ideals, then there is some
N
such that
I
N+m
=
I
N
for all m 0.
Similarly, we can define right Noetherian algebras, and say an algebra is
Noetherian if it is both left and right Noetherian.
We can again look at some examples.
Example. Again all finite-dimensional algebras are Noetherian.
Example.
In the commutative case, Hilbert’s basis theorem tells us a polynomial
algebra
k
[
X
1
, · · · , X
k
] in finitely many variables is Noetherian. Similarly, the
power series rings k[[X
1
, · · · , X
n
]] are Noetherian.
Example.
The universal enveloping algebra of a finite-dimensional Lie algebra
are the (associative!) algebras that underpin the representation theory of these
Lie algebras.
Example.
Some differential operator algebras are Noetherian. We assume
char k
= 0. Consider the polynomial ring
k
[
X
]. We have operators “multipli-
cation by
X
and “differentiate with respect to
X
on
k
[
X
]. We can form the
algebra
k
[
X,
x
] of differential operators on
k
[
X
], with multiplication given by
the composition of operators. This is called the Weyl algebra
A
1
. We will show
that this is a non-commutative Noetherian algebra.
Example.
Some group algebras are Noetherian. Clearly all group algebras of
finite groups are Noetherian, but the group algebras of certain infinite groups
are Noetherian. For example, we can take
G =
1 λ µ
0 1 ν
0 0 0
: λ, µ, ν Z
,
and this is Noetherian. However, we shall not prove this.
We will see that all left Artinian algebras are left Noetherian. While there is
a general theory of non-commutative Noetherian algebras, it is not as useful as
the theory of commutative Noetherian algebras.
In the commutative case, we often look at
Spec A
, the set of prime ideals of
A
. However, sometimes in the non-commutative there are few prime ideals, and
so Spec is not going to keep us busy.
Example. In the Weyl algebra A
1
, the only prime ideals are 0 and A
1
.
We will prove a theorem of Goldie:
Theorem
(Goldie’s theorem)
.
Let
A
be a right Noetherian algebra with no
non-zero ideals all of whose elements are nilpotent. Then
A
embeds in a finite
direct sum of matrix algebras over division algebras.
Some types of Noetherian algebras can be thought of as non-commutative
polynomial algebras and non-commutative power series, i.e. they are deformations
of the analogous commutative algebra. For example, we say
A
1
is a deformation
of the polynomial algebra
k
[
X, Y
], where instead of having
XY Y X
= 0, we
have
XY Y X
= 1. This also applies to enveloping algebras and Iwasawa
algebras. We will study when one can deform the multiplication so that it remains
associative, and this is bound up with the cohomology theory of associative
algebras Hochschild cohomology. The Hochschild complex has rich algebraic
structure, and this will allow us to understand how we can deform the algebra.
At the end, we shall quickly talk about bialgebras and Hopf algebras. In a
bialgebra, one also has a comultiplication map
A AA
, which in representation
theory is crucial in saying how to regard a tensor product of two representations
as a representation.
1 Artinian algebras
1.1 Artinian algebras
We continue writing down some definitions. We already defined left and right
Artinian algebras in the introduction. Most examples we’ll meet are in fact
finite-dimensional vector spaces over
k
. However, there exists some more perverse
examples:
Example. Let
A =

r s
0 t
: r Q, s, t R
Then this is right Artinian but not left Artinian over
Q
. To see it is not left
Artinian, note that there is an ideal
I =

0 s
0 0
: s R
=
R
of
A
, and a matrix
r s
0 t
acts on this on the left by sending
0 s
0
0 0
to
rs
0
. Since
R
is an infinite-dimensional
Q
-vector space, one sees that there is an
infinite strictly descending chain of ideals contained in I.
The fact that it is right Artinian is a direct verification. Indeed, it is not
difficult to enumerate all the right ideals, which is left as an exercise for the
reader.
As in the case of commutative algebra, we can study the modules of an
algebra.
Definition
(Module)
.
Let
A
be an algebra. A left
A
-module is a
k
-vector space
M and a bilinear map
A M M
a m xm
such that (
ab
)
m
=
a
(
bm
) for all
a, b A
and
m M
. Right
A
-modules are
defined similarly.
An
A
-
A
-bimodule is a vector space
M
that is both a left
A
-module and a
right
A
-module, such that the two actions commute for
a, b A
and
x M
,
we have
a(xb) = (ax)b.
Example.
The algebra
A
itself is a left
A
-module. We write this as
A
A
, and
call this the left regular representation. Similarly, the right action is denoted
A
A
.
These two actions are compatible by associativity, so A is an A-A-bimodule.
If we write
End
k
(
A
) for the
k
-linear maps
A A
, then
End
k
is naturally a
k
-
algebra by composition, and we have a
k
-algebra homomorphism
A End
k
(
A
)
that sends
a A
to multiplication by
a
on the left. However, if we want
to multiply on the right instead, it is no longer a
k
-algebra homomorphism
A End
k
(A). Instead, it is a map A End
k
(A)
op
, where
Definition
(Opposite algebra)
.
Let
A
be a
k
-algebra. We define the opposite
algebra
A
op
to be the algebra with the same underlying vector space, but with
multiplication given by
x · y = yx.
Here on the left we have the multiplication in
A
op
and on the right we have the
multiplication in A.
In general, a left A-module is a right A
op
-module.
As in the case of ring theory, we can talk about prime ideals. However, we
will adopt a slightly different definition:
Definition
(Prime ideal)
.
An ideal
P
is prime if it is a proper ideal, and if
I
and J are ideals with IJ P , then either I P or J P .
It is an exercise to check that this coincides in the commutative case with
the definition using elements.
Definition
(Annihilator)
.
Let
M
be a left
A
-module and
m M
. We define
the annihilators to be
Ann(m) = {a A : am = 0}
Ann(M) = {a A : am = 0 for all m M} =
\
mM
Ann(m).
Note that
Ann
(
m
) is a left ideal of
A
, and is in fact the kernel of the
A
-module
homomorphism
A M
given by
x 7→ xm
. We’ll denote the image of this map
by Am, a left submodule of M, and we have
A
Ann(m)
=
Am.
On the other hand, it is easy to see that
Ann
(
M
) is an fact a (two-sided) ideal.
Definition
(Simple module)
.
A non-zero module
M
is simple or irreducible if
the only submodules of M are 0 and M.
It is easy to see that
Proposition.
Let
A
be an algebra and
I
a left ideal. Then
I
is a maximal left
ideal iff A/I is simple.
Example. Ann(m) is a maximal left ideal iff Am is irreducible.
Proposition.
Let
A
be an algebra and
M
a simple module. Then
M
=
A/I
for some (maximal) left ideal I of A.
Proof.
Pick an arbitrary element
m M
, and define the
A
-module homomor-
phism
ϕ
:
A M
by
ϕ
(
a
) =
am
. Then the image is a non-trivial submodule,
and hence must be
M
. Then by the first isomorphism theorem, we have
M
=
A/ ker ϕ.
Before we start doing anything, we note the following convenient lemma:
Lemma.
Let
M
be a finitely-generated
A
module. Then
M
has a maximal
proper submodule M
0
.
Proof.
Let
m
1
, · · · , m
k
M
be a minimal generating set. Then in particular
N
=
hm
1
, · · · , m
k1
i
is a proper submodule of
M
. Moreover, a submodule
of
M
containing
N
is proper iff it does not contain
m
k
, and this property is
preserved under increasing unions. So by Zorn’s lemma, there is a maximal
proper submodule.
Definition
(Jacobson radical)
.
The
J
(
A
) of
A
is the intersection of all maximal
left ideals.
This is in fact an ideal, and not just a left one, because
J(A) =
\
{maximal left ideals} =
\
mM,M simple
Ann(m) =
\
M simple
Ann(M),
which we have established is an ideal. Yet, it is still not clear that this is
independent of us saying “left” instead of “right”. It, in fact, does not, and this
follows from the Nakayama lemma:
Lemma
(Nakayama lemma)
.
The following are equivalent for a left ideal
I
of
A.
(i) I J(A).
(ii)
For any finitely-generated left
A
-module
M
, if
IM
=
M
, then
M
= 0,
where
IM
is the module generated by elements of the form
am
, with
a I
and m M.
(iii) G = {1 + a : a I} = 1 + I is a subgroup of the unit group of A.
In particular, this shows that the Jacobson radical is the largest ideal satisfying
(iii), which is something that does not depend on handedness.
Proof.
(i)
(ii): Suppose
I J
(
A
) and
M 6
= 0 is a finitely-generated
A
-module,
and we’ll see that IM M.
Let
N
be a maximal submodule of
M
. Then
M/N
is a simple module,
so for any
¯m M/N
, we know
Ann
(
¯m
) is a maximal left ideal. So
J(A) Ann(M/N ). So IM J(A)M N M.
(ii)
(iii): Assume (ii). We let
x I
and set
y
= 1 +
x
. Hence
1 =
y x Ay
+
I
. Since
Ay
+
I
is a left ideal, we know
Ay
+
I
=
A
. In
other words, we know
I
A
Ay
=
A
Ay
.
Now using (ii) on the finitely-generated module
A/Ay
(it is in fact generated
by 1), we know that
A/Ay
= 0. So
A
=
Ay
. So there exists
z A
such
that 1 =
zy
=
z
(1 +
x
). So (1 +
x
) has a left inverse, and this left inverse
z
lies in
G
, since we can write
z
= 1
zx
. So
G
is a subgroup of the unit
group of A.
(iii)
(i): Suppose
I
1
is a maximal left ideal of
A
. Let
x I
. If
x 6∈ I
1
,
then
I
1
+
Ax
=
A
by maximality of
I
. So 1 =
y
+
zx
for some
y I
1
and
z A
. So
y
= 1
zx G
. So
y
is invertible. But
y I
1
. So
I
1
=
A
.
This is a contradiction. So we found that
I < I
1
, and this is true for all
maximal left ideals I
1
. Hence I J(A).
We now come to the important definition:
Definition (Semisimple algebra). An algebra is semisimple if J(A) = 0.
We will very soon see that for Artinian algebras, being semi-simple is equiva-
lent to a few other very nice properties, such as being completely reducible. For
now, we shall content ourselves with some examples.
Example. For any A, we know A/J(A) is always semisimple.
We can also define
Definition (Simple algebra). An algebra is simple if the only ideals are 0 and
A.
It is trivially true that any simple algebra is semi-simple the Jacobson
radical is an ideal, and it is not
A
. A particularly important example is the
following:
Example.
Consider
M
n
(
k
). We let
e
i
be the matrix with 1 in the (
i, i
)th
entry and zero everywhere else. This is idempotent, i.e.
e
2
i
=
e
i
. It is also
straightforward to check that
Ae
i
=
0 · · · 0 a
1
0 · · · 0
0 · · · 0 a
2
0 · · · 0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 · · · 0 a
n
0 · · · 0
The non-zero column is, of course, the
i
th column. Similarly,
e
i
A
is the matrices
that are zero apart from in the
i
th row. These are in fact all left and all right
ideals respectively. So the only ideals are 0 and A.
As a left A-module, we can decompose A as
A
A =
n
M
i=1
Ae
i
,
which is a decomposition into simple modules.
Definition
(Completely reducible)
.
A module
M
of
A
is completely reducible
iff it is a sum of simple modules.
Here in the definition, we said “sum” instead of “direct sum”, but the
following proposition shows it doesn’t matter:
Proposition. Let M be an A-module. Then the following are equivalent:
(i) M is completely reducible.
(ii) M is the direct sum of simple modules.
(iii)
Every submodule of
M
has a complement, i.e. for any submodule
N
of
M
,
there is a complement N
0
such that M = N N
0
.
Often, the last condition is the most useful in practice.
Proof.
(i) (ii): Let M be completely reducible, and consider the set
n
{S
α
M} : S
α
are simple,
X
S
α
is a direct sum
o
.
Notice this set is closed under increasing unions, since the property of
being a direct sum is only checked on finitely many elements. So by Zorn’s
lemma, it has a maximal element, and let N be the sum of the elements.
Suppose this were not all of
M
. Then there is some
S M
such that
S 6⊆ N
. Then
S N ( S
. By simplicity, they intersect trivially. So
S
+
N
is a direct sum, which is a contradiction. So we must have
N
=
M
, and
M is the direct sum of simple modules.
(ii) (i) is trivial.
(i) (iii): Let N M be a submodule, and consider
n
{S
α
M} : S
α
are simple, N +
X
S
α
is a direct sum
o
.
Again this set has a maximal element, and let
P
be the direct sum of those
S
α
. Again if
P N
is not all of
M
, then pick an
S M
simple such that
S
is not contained in
P N
. Then again
S P N
is a direct sum, which
is a contradiction.
(iii)
(i): It suffices to show that if
N < M
is a proper submodule,
then there exists a simple module that intersects
N
trivially. Indeed, we
can take
N
to be the sum of all simple submodules of
M
, and this forces
N = M .
To do so, pick an
x 6∈ N
, and let
P
be submodule of
M
maximal among
those satisfying
P N
= 0 and
x 6∈ N P
. Then
N P
is a proper
submodule of M. Let S be a complement. We claim S is simple.
If not, we can find a proper submodule
S
0
of
S
. Let
Q
be a complement of
N P S
0
. Then we can write
M = N P S
0
Q
x = n + p + s + q
.
By assumption,
s
and
q
are not both zero. We wlog assume
s
is non-
zero. Then
P Q
is a larger submodule satisfying (
P Q
)
N
= 0 and
x 6∈ N
(
P Q
). This is a contradiction. So
S
is simple, and we are
done.
Using these different characterizations, we can prove that completely reducible
modules are closed under the familiar ex operations.
Proposition.
Sums, submodules and quotients of completely reducible modules
are completely reducible.
Proof.
It is clear by definition that sums of completely reducible modules are
completely reducible.
To see that submodules of completely reducible modules are completely
reducible, let
M
be completely reducible, and
N M
. Then for each
x N
,
there is some simple submodule
S M
containing
x
. Since
S N S
and
contains x, it must be S, i.e. S N. So N is the sum of simple modules.
Finally, to see quotients are completely reducible, if
M
is completely reducible
and N is a submodule, then we can write
M = N P
for some P . Then M/N
=
P , and P is completely reducible.
We will show that every left Artinian algebra is completely reducible over
itself iff it is semi-simple. We can in fact prove a more general fact for
A
-modules.
To do so, we need a generalization of the Jacobson radical.
Definition
(Radical)
.
For a module
M
, we write
Rad
(
M
) for the intersection
of maximal submodules of M, and call it the radical of M.
Thus, we have Rad(
A
A) = J(A) = Rad(A
A
).
Proposition.
Let
M
be an
A
-module satisfying the descending chain condition
on submodules. Then M is completely reducible iff Rad(M) = 0.
Proof.
It is clear that if
M
is completely reducible, then
Rad
(
M
) = 0. Indeed,
we can write
M =
M
αA
S
α
,
where each S
α
is simple. Then
J(A)
\
αA
M
βA\{α}
S
β
= {0}.
Conversely, if
Rad
(
M
) = 0, we note that since
M
satisfies the descending
chain condition on submodules, there must be a finite collection
M
1
, · · · , M
n
of
maximal submodules whose intersection vanish. Then consider the map
M
n
M
i=1
M
M
i
x (x + M
1
, x + M
2
, · · · , x + M
n
)
The kernel of this map is the intersection of the
M
i
, which is trivial. So this
embeds
M
as a submodule of
L
M
M
i
. But each
M
M
i
is simple, so
M
is a submodule
of a completely reducible module, hence completely reducible.
Corollary.
If
A
is a semi-simple left Artinian algebra, then
A
A
is completely
reducible.
Corollary.
If
A
is a semi-simple left Artinian algebra, then every left
A
-module
is completely reducible.
Proof.
Every
A
-module
M
is a quotient of sums of
A
A
. Explicitly, we have a
map
M
mM
A
A M
(a
m
)
P
a
m
m
Then this map is clearly surjective, and thus M is a quotient of
L
M
A
A.
If
A
is not semi-simple, then it turns out it is rather easy to figure out radical
of M , at least if M is finitely-generated.
Lemma.
Let
A
be left Artinian, and
M
a finitely generated left
A
-module, then
J(A)M = Rad(M ).
Proof.
Let
M
0
be a maximal submodule of
M
. Then
M/M
0
is simple, and is in
fact A/I for some maximal left ideal I. Then we have
J(A)
M
M
0
= 0,
since J(A) < I. Therefore J(A)M M
0
. So J(A)M Rad(M).
Conversely, we know
M
J(A)M
is an
A/J
(
A
)-module, and is hence completely
reducible as
A/J
(
A
) is semi-simple (and left Artinian). Since an
A
-submodule
of
M
J(A)M
is the same as an
A/J
(
A
)-submodule, it follows that it is completely
reducible as an A-module as well. So
Rad
M
J(A)M
= 0,
and hence Rad(M ) J(A)M.
Proposition. Let A be left Artinian. Then
(i) J(A) is nilpotent, i.e. there exists some r such that J(A)
r
= 0.
(ii)
If
M
is a finitely-generated left
A
-module, then it is both left Artinian and
left Noetherian.
(iii) A is left Noetherian.
Proof.
(i)
Since
A
is left-Artinian, and
{J
(
A
)
r
:
r N}
is a descending chain of
ideals, it must eventually be constant. So
J
(
A
)
r
=
J
(
A
)
r+1
for some
r
. If
this is non-zero, then again using the descending chain condition, we see
there is a left ideal
I
with
J
(
A
)
r
I 6
= 0 that is minimal with this property
(one such ideal exists, say J(A) itself).
Now pick
x I
with
J
(
A
)
r
x 6
= 0. Since
J
(
A
)
2r
=
J
(
A
)
r
, it follows
that
J
(
A
)
r
(
J
(
A
)
r
x
)
6
= 0. So by minimality,
J
(
A
)
r
x I
. But the other
inclusion clearly holds. So they are equal. So there exists some
a J
(
A
)
r
with x = ax. So
(1 a)x = 0.
But 1 a is a unit. So x = 0. This is a contradiction. So J(A)
r
= 0.
(ii)
Let
M
i
=
J
(
A
)
i
M
. Then
M
i
/M
i+1
is annihilated by
J
(
A
), and hence
completely reducible (it is a module over semi-simple
A/J
(
A
)). Since
M
is a finitely generated left
A
-module for a left Artinian algebra, it satisfies
the descending chain condition for submodules (exercise), and hence so
does M
i
/M
i+1
.
So we know
M
i
/M
i+1
is a finite sum of simple modules, and therefore
satisfies the ascending chain condition. So
M
i
/M
i+1
is left Noetherian,
and hence M is (exercise).
(iii) Follows from (ii) since A is a finitely-generated left A-module.
1.2 Artin–Wedderburn theorem
We are going to state the Artin–Wedderburn theorem for right (as opposed to
left) things, because this makes the notation easier for us.
Theorem
(Artin–Wedderburn theorem)
.
Let
A
be a semisimple right Artinian
algebra. Then
A =
r
M
i=1
M
n
i
(D
i
),
for some division algebra D
i
, and these factors are uniquely determined.
A has exactly r isomorphism classes of simple (right) modules S
i
, and
End
A
(S
i
) = {A-module homomorphisms S
i
S
i
}
=
D
i
,
and
dim
D
i
(S
i
) = n
i
.
If A is simple, then r = 1.
If we had the left version instead, then we need to insert op’s somewhere.
Artin–Wedderburn is an easy consequence of two trivial lemma. The key idea
that leads to Artin–Wedderburn is the observation that the map
A
A
End
A
(
A
A
)
sending
a
to left-multiplication by
a
is an isomorphism of algebras. So we need
to first understand endomorphism algebras, starting with Schur’s lemma.
Lemma
(Schur’s lemma)
.
Let
M
1
, M
2
be simple right
A
-modules. Then either
M
1
=
M
2
, or
Hom
A
(
M
1
, M
2
) = 0. If
M
is a simple
A
-module, then
End
A
(
M
)
is a division algebra.
Proof.
A non-zero
A
-module homomorphism
M
1
M
2
must be injective, as
the kernel is submodule. Similarly, the image has to be the whole thing since
the image is a submodule. So this must be an isomorphism, and in particular
has an inverse. So the last part follows as well.
As mentioned, we are going to exploit the isomorphism
A
A
=
End
A
(
A
A
).
This is easy to see directly, but we can prove a slightly more general result, for
the sake of it:
Lemma.
(i)
If
M
is a right
A
-module and
e
is an idempotent in
A
, i.e.
e
2
=
e
, then
Me
=
Hom
A
(eA, M ).
(ii) We have
eAe
=
End
A
(eA).
In particular, we can take e = 1, and recover End
A
(A
A
)
=
A.
Proof.
(i) We define maps
me (ex 7→ mex)
Me Hom(eA, M )
α(e) α
f
1
f
2
We note that
α
(
e
) =
α
(
e
2
) =
α
(
e
)
e Me
. So this is well-defined. By
inspection, these maps are inverse to each other. So we are done.
Note that we might worry that we have to pick representatives
me
and
ex
for the map
f
1
, but in fact we can also write it as
f
(
a
)(
y
) =
ay
, since
e
is
idempotent. So we are safe.
(ii) Immediate from above by putting M = eA.
Lemma. Let M be a completely reducible right A-module. We write
M =
M
S
n
i
i
,
where
{S
i
}
are distinct simple
A
-modules. Write
D
i
=
End
A
(
S
i
), which we
already know is a division algebra. Then
End
A
(S
n
i
i
)
=
M
n
i
(D
i
),
and
End
A
(M) =
M
M
n
i
(D
i
)
Proof.
The result for
End
A
(
S
n
i
i
) is just the familiar fact that a homomorphism
S
n
S
m
is given by an
m × n
matrix of maps
S S
(in the case of vector
spaces over a field
k
, we have
End
(
k
)
=
k
, so they are matrices with entries in
k). Then by Schur’s lemma, we have
End
A
(M) =
M
i
End
A
(M
i
)
=
M
n
i
(D
i
).
We now prove Artin–Wedderburn.
Proof of Artin–Wedderburn.
If
A
is semi-simple, then it is completely reducible
as a right A-module. So we have
A
=
End(A
A
)
=
M
M
n
i
(D
i
).
We now decompose each
M
n
i
(
D
i
) into a sum of simple modules. We know each
M
n
i
(
D
i
) is a non-trivial
M
n
i
(
D
i
) module in the usual way, and the action of
the other summands is trivial. We can simply decompose each
M
n
i
(
D
i
) as the
sum of submodules of the form
0 0 · · · 0 0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 0 · · · 0 0
a
1
a
2
· · · a
n
i
1
a
n
i
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 0 · · · 0 0
and there are
n
i
components. We immediately see that if we write
S
i
for this
submodule, then we have
dim
D
i
(S
i
) = n
i
.
Finally, we have to show that every simple module S of A is one of the S
i
. We
simply have to note that if
S
is a simple
A
-module, then there is a non-trivial
map
f
:
A S
(say by picking
x S
and defining
f
(
a
) =
xa
). Then in the
decomposition of
A
into a direct sum of simple submodules, there must be one
factor
S
i
such that
f|
S
i
is non-trivial. Then by Schur’s lemma, this is in fact an
isomorphism S
i
=
S.
This was for semi-simple algebras. For a general right Artinian algebra, we
know that
A/J
(
A
) is semi-simple and inherits the Artinian property. Then
Artin–Wedderburn applies to A/J(A).
Some of us might be scared of division algebras. Sometimes, we can get away
with not talking about them. If
A
is not just Artinian, but finite-dimensional,
then so are the D
i
.
Now pick an arbitrary
x D
i
. Then the sub-algebra of
D
i
generated by
x
is
be commutative. So it is in fact a subfield, and finite dimensionality means it is
algebraic over
k
. Now if we assume that
k
is algebraically closed, then
x
must
live in k. So we’ve shown that these D
i
must be k itself. Thus we get
Corollary.
If
k
is algebraically closed and
A
is a finite-dimensional semi-simple
k-algebra, then
A
=
M
M
n
i
(k).
This is true, for example, when k = C.
We shall end this section by applying our results to group algebras. Recall
the following definition:
Definition
(Group algebra)
.
Let
G
be a group and
k
a field. The group algebra
of G over k is
kG =
n
X
λ
g
g : g G, λ
g
k
o
.
This has a bilinear multiplication given by the obvious formula
(λ
g
g)(µ
h
h) = λ
g
µ
h
(gh).
The first thing to note is that group algebras are almost always semi-simple.
Theorem
(Maschke’s theorem)
.
Let
G
be a finite group and
p - |G|
, where
p = char k, so that |G| is invertible in k, then kG is semi-simple.
Proof.
We show that any submodule
V
of a
kG
-module
U
has a complement.
Let π : U V be any k-vector space projection, and define a new map
π
0
=
1
|G|
X
gG
gπg
1
: U V.
It is easy to see that this is a
kG
-module homomorphism
U V
, and is a
projection. So we have
U = V ker π
0
,
and this gives a kG-module complement.
There is a converse to Maschke’s theorem:
Theorem. Let G be finite and kG semi-simple. Then char k - |G|.
Proof.
We note that there is a simple
kG
-module
S
, given by the trivial module.
This is a one-dimensional k vector space. We have
D = End
kG
(S) = k.
Now suppose
kG
is semi-simple. Then by Artin–Wedderburn, there must be
only one summand of S in kG.
Consider the following two ideals of kG: we let
I
1
=
n
X
λ
g
g kG :
X
λ
g
= 0
o
.
This is in fact a two-sided ideal of
kG
. We also have the center of the algebra,
given by
I
2
=
n
λ
X
g kG : λ k
o
.
Now if char k | |G|, then I
2
I
1
. So we can write
kG =
kG
I
1
I
1
=
kG
I
1
I
2
· · · .
But we know
G
acts trivially on
kG
I
1
and
I
2
, and they both have dimension 1.
This gives a contradiction. So we must have char k - |G|.
We can do a bit more of representation theory. Recall that when
k
is
algebraically closed and has characteristic zero, then the number of simple
kG
-
modules is the number of conjugacy classes of
G
. There is a more general result
for a general characteristic p field:
Theorem.
Let
k
be algebraically closed of characteristic
p
, and
G
be finite.
Then the number of simple
kG
modules (up to isomorphism) is equal to the
number of conjugacy classes of elements of order not divisible by
p
. These are
known as the p-regular elements.
We immediately deduce that
Corollary.
If
|G|
=
p
r
for some
r
and
p
is prime, then the trivial module is the
only simple kG module, when char k = p.
Note that we can prove this directly rather than using the theorem, by
showing that
I
=
ker
(
kG k
) is a nilpotent ideal, and annihilates all simple
modules.
Proof sketch of theorem.
The number of simple
kG
modules is just the number
of simple
kG/J
(
kG
) module, as
J
(
kG
) acts trivially on every simple module.
There is a useful trick to figure out the number of simple
A
-modules for a given
semi-simple A. Suppose we have a decomposition
A
=
r
M
i=1
M
n
i
(k).
Then we know
r
is the number of simple
A
-modules. We now consider [
A, A
],
the k-subspace generated by elements of the form xy yx. Then we see that
A
[A, A]
=
r
M
i=1
M
n
i
(k)
[M
n
i
(k), M
n
i
(k)]
.
Now by linear algebra, we know [
M
n
i
(
k
)
, M
n
i
(
k
)] is the trace zero matrices, and
so we know
dim
k
M
n
i
(k)
[M
n
i
(k), M
n
i
(k)]
= 1.
Hence we know
dim
A
[A, A]
= r.
Thus we need to compute
dim
k
kG/J(kG)
[kG/J(kG), kG/J(kG)]
We then note the following facts:
(i) For a general algebra A, we have
A/J(A)
[A/J(A), A/J(A)]
=
A
[A, A] + J(A)
.
(ii) Let g
1
, · · · , g
m
be conjugacy class representatives of G. Then
{g
i
+ [kG, kG]}
forms a k-vector space basis of kG/[kG, kG].
(iii)
If
g
1
, · · · , g
r
is a set of representatives of
p
-regular conjugacy classes, then
n
g
i
+
[kG, kG] + J(kG)
o
form a basis of kG/([kG, kG] + J(kG)).
Hence the result follows.
One may find it useful to note that [
kG, kG
] +
J
(
kG
) consists of the elements
in kG such that x
p
s
[kG, kG] for some s.
In this proof, we look at
A/
[
A, A
]. However, in the usual proof of the result
in the characteristic zero looks at the center
Z
(
A
). The relation between these
two objects is that the first is the 0th Hochschild homology group of
A
, while
the second is the 0th Hochschild cohomology group of A.
1.3 Crossed products
Number theorists are often interested in representations of Galois groups and
kG
-modules where
k
is an algebraic number field, e.g.
Q
. In this case, the
D
i
’s
appearing in Artin–Wedderburn may be non-commutative.
We have already met one case of a non-commutative division ring, namely
the quaternions H. This is in fact an example of a general construction.
Definition
(Crossed product)
.
The crossed product of a
k
-algebra
B
and a
group G is specified by the following data:
A group homomorphism φ : G Aut
k
(B), written
φ
g
(λ) = λ
g
;
A function
Ψ(g, h) : G × G B.
The crossed product algebra has underlying set
X
λ
g
g : λ
g
B.
with operation defined by
λg · µh = λµ
g
Ψ(g, h)(gh).
The function Ψ is required to be such that the resulting product is associative.
We should think of the
µ
g
as specifying what happens when we conjugate
g
pass µ, and then Ψ(g, h)(gh) is the product of g and h in the crossed product.
Usually, we take
B
=
K
, a Galois extension of
k
, and
G
=
Gal
(
K/k
). Then
the action
φ
g
is the natural action of
G
on the elements of
K
, and we restrict to
maps Ψ : G × G K
×
only.
Example.
Consider
B
=
K
=
C
, and
k
=
R
. Then
G
=
Gal
(
C/R
)
=
Z/
2
Z
=
{e, g}, where g is complex conjugation. The elements of H are of the form
λ
e
e + λ
g
g,
where λ
e
, λ
g
C, and we will write
1 · g = g, i · g = k, 1 · e = 1, i · e = i.
Now we want to impose
1 = j
2
= 1g · 1g = ψ(g, g)e.
So we set Ψ(
g, g
) =
1. We can similarly work out what we want the other
values of Ψ to be.
Note that in general, crossed products need not be division algebras.
The crossed product is not just a
k
-algebra. It has a natural structure of a
G-graded algebra, in the sense that we can write it as a direct sum
BG =
M
gG
Bg,
and we have Bg
1
· Bg
2
Bg
1
g
2
.
Focusing on the case where
K/k
is a Galois extension, we use the notation
(
K, G,
Ψ), where Ψ :
G × G K
×
. Associativity of these crossed products is
equivalent to a 2-cocycle condition Ψ, which you will be asked to make precise
on the first example sheet.
Two crossed products (
K, G,
Ψ
1
) and (
K, G,
Ψ
2
) are isomorphic iff the map
G × G K
×
(g, h) Ψ
1
(g, h)(Ψ
2
(g, h))
1
satisfies a 2-coboundary condition, which is again left for the first example sheet.
Therefore the second (group) cohomology
{2-cocycles : G × G K
×
}
{2-coboundaries : G × G K
×
}
determines the isomorphism classes of (associative) crossed products (K, G, Ψ).
Definition
(Central simple algebra)
.
A central simple
k
-algebra is a finite-
dimensional k-algebra which is a simple algebra, and with a center Z(A) = k.
Note that any simple algebra is a division algebra, say by Schur’s lemma.
So the center must be a field. Hence any simple
k
-algebra can be made into a
central simple algebra simply by enlarging the base field.
Example. M
n
(k) is a central simple algebra.
The point of talking about these is the following result:
Fact.
Any central simple
k
-algebra is of the form
M
n
(
D
) for some division
algebra
D
which is also a central simple
k
-algebra, and is a crossed product
(K, G, Ψ).
Note that when
K
=
C
and
k
=
R
, then the second cohomology group has
2 elements, and we get that the only central simple
R
-algebras are
M
n
(
R
) or
M
n
(H).
For amusement, we also note the following theorem:
Fact (Wedderburn). Every finite division algebra is a field.
1.4 Projectives and blocks
In general, if
A
is not semi-simple, then it is not possible to decompose
A
as a
direct sum of simple modules. However, what we can do is to decompose it as a
direct sum of indecomposable projectives.
We begin with the definition of a projective module.
Definition
(Projective module)
.
An
A
-module is projective
P
if given modules
M and M
0
and maps
P
M
0
M 0
α
θ
,
then there exists a map
β
:
P M
0
such that the following diagram commutes:
P
M
0
M 0
α
β
θ
.
Equivalently, if we have a short exact sequence
0 N M
0
M 0,
then the sequence
0 Hom(P, N) Hom(P, M
0
) Hom(P, M) 0
is exact.
Note that we get exactness at
Hom
(
P, N
) and
Hom
(
P, M
0
) for any
P
at all.
Projective means it is also exact at Hom(P, M ).
Example. Free modules are always projective.
In general, projective modules are “like” free modules. We all know that free
modules are nice, and most of the time, when we want to prove things about
free modules, we are just using the property that they are projective. It is also
possible to understand projective modules in an algebro-geometric way they
are “locally free” modules.
It is convenient to characterize projective modules as follows:
Lemma. The following are equivalent:
(i) P is projective.
(ii) Every surjective map φ : M P splits, i.e.
M
=
ker φ N
where N
=
P .
(iii) P is a direct summand of a free module.
Proof.
(i) (ii): Consider the following lifting problem:
P
M P 0
φ
,
The lifting gives an embedding of
P
into
M
that complements
ker φ
(by
the splitting lemma, or by checking it directly).
(ii)
(iii): Every module admits a surjection from a free module (e.g. the
free module generated by the elements of P )
(iii)
(i): It suffices to show that direct summands of projectives are
projective. Suppose P is is projective, and
P
=
A B.
Then any diagram
A
M
0
M 0
α
θ
,
can be extended to a diagram
A B
M
0
M 0
˜α
θ
,
by sending
B
to 0. Then since
A B
=
P
is projective, we obtain a lifting
A B M
0
, and restricting to A gives the desired lifting.
Our objective is to understand the direct summands of a general Artinian
k
-algebra
A
, not necessarily semi-simple. Since
A
is itself a free
A
-module, we
know these direct summands are always projective.
Since
A
is not necessarily semi-simple, it is in general impossible to decompose
it as a direct sum of simples. What we can do, though, is to decompose it as a
direct sum of indecomposable modules.
Definition
(Indecomposable)
.
A non-zero module
M
is indecomposable if
M
cannot be expressed as the direct sum of two non-zero submodules.
Note that since
A
is (left) Artinian, it can always be decomposed as a finite
sum of indecomposable (left) submodules. Sometimes, we are also interested in
decomposing A as a sum of (two-sided) ideals. These are called blocks.
Definition
(Block)
.
The blocks are the direct summands of
A
that are inde-
composable as ideals.
Example. If A is semi-simple Artinian, then Artin-Wedderburn tells us
A =
M
M
n
i
(D
i
),
and the M
n
i
(D
i
) are the blocks.
We already know that every Artinian module can be decomposed as a direct
sum of indecomposables. The first question to ask is whether this is unique. We
note the following definitions:
Definition
(Local algebra)
.
An algebra is local if it has a unique maximal left
ideal, which is J(A), which is the unique maximal right ideal.
If so, then
A/J
(
A
) is a division algebra. This name, of course, comes from
algebraic geometry (cf. local rings).
Definition
(Unique decomposition property)
.
A module
M
has the unique
decomposition property if
M
is a finite direct sum of indecomposable modules,
and if
M =
m
M
i=1
M
i
=
n
M
i=1
M
0
i
,
then n = m, and, after reordering, M
i
= M
0
i
.
We want to prove that
A
as an
A
-module always has the unique decomposi-
tion property. The first step is the following criterion for determining unique
decomposition property.
Theorem
(Krull–Schmidt theorem)
.
Suppose
M
is a finite sum of indecom-
posable
A
-modules
M
i
, with each
End
(
M
i
) local. Then
M
has the unique
decomposition property.
Proof. Let
M =
m
M
i=1
M
i
=
n
M
i=1
M
0
i
.
We prove by induction on
m
. If
m
= 1, then
M
is indecomposable. Then we
must have n = 1 as well, and the result follows.
For m > 1, we consider the maps
α
i
: M
0
i
M M
1
β
i
: M
1
M M
0
i
We observe that
id
M
1
=
n
X
i=1
α
i
β
i
: M
1
M
1
.
Since
End
A
(
M
1
) is local, we know some
α
i
β
i
must be invertible, i.e. a unit, as
they cannot all lie in the Jacobson radical. We may wlog assume
α
1
β
1
is a
unit. If this is the case, then both
α
1
and
β
1
have to be invertible. So
M
1
=
M
0
1
.
Consider the map id θ = φ, where
θ : M M
1
M
0
1
M
L
m
i=2
M
i
M.
α
1
1
Then φ(M
0
1
) = M
1
. So φ|
M
0
1
looks like α
1
. Also
φ
m
M
i=2
M
i
!
=
m
M
i=2
M
i
,
So
φ|
L
m
i=2
M
i
looks like the identity map. So in particular, we see that
φ
is
surjective. However, if φ(x) = 0, this says x = θ(x), So
x
m
M
i=2
M
i
.
But then
θ
(
x
) = 0. Thus
x
= 0. Thus
φ
is an automorphism of
m
with
φ(M
0
1
) = φ(M
1
). So this gives an isomorphism between
m
M
i=2
M
i
=
M
M
1
=
M
M
1
=
n
M
i=2
M
0
i
,
and so we are done by induction.
Now it remains to prove that the endomorphism rings are local. Recall the
following result from linear algebra.
Lemma
(Fitting)
.
Suppose
M
is a module with both the ACC and DCC on
submodules, and let f End
A
(M). Then for large enough n, we have
M = im f
n
ker f
n
.
Proof. By ACC and DCC, we may choose n large enough so that
f
n
: f
n
(M) f
2n
(M)
is an isomorphism, as if we keep iterating
f
, the image is a descending chain and
the kernel is an ascending chain, and these have to terminate.
If m M, then we can write
f
n
(m) = f
2n
(m
1
)
for some m
1
. Then
m = f
n
(m
1
) + (m f
n
(m
1
)) im f
n
+ ker f
n
,
and also
im f
n
ker f
n
= ker(f
n
: f
n
(M) f
2n
(M)) = 0.
So done.
Lemma.
Suppose
M
is an indecomposable module satisfying ACC and DCC
on submodules. Then B = End
A
(M) is local.
Proof.
Choose a maximal left ideal of
B
, say
I
. It’s enough to show that if
x 6∈ I
,
then x is left invertible. By maximality of I, we know B = Bx + I. We write
1 = λx + y,
for some
λ B
and
y I
. Since
y I
, it has no left inverse. So it is not an
isomorphism. By Fitting’s lemma and the indecomposability of
M
, we see that
y
m
= 0 for some m. Thus
(1 + y + y
2
+ · · · + y
m1
)λx = (1 + y + · · · + y
m1
)(1 y) = 1.
So x is left invertible.
Corollary.
Let
A
be a left Artinian algebra. Then
A
has the unique decompo-
sition property.
Proof.
We know
A
satisfies the ACC and DCC condition. So
A
A
is a finite
direct sum of indecomposables.
So if
A
is an Artinian algebra, we know
A
can be uniquely decomposed as a
direct sum of indecomposable projectives,
A =
M
P
j
.
For convenience, we will work with right Artinian algebras and right modules
instead of left ones. It turns out that instead of studying projectives in
A
, we
can study idempotent elements instead.
Recall that
End
(
A
A
)
=
A
. The projection onto
P
j
is achieved by left
multiplication by an idempotent e
j
,
P
j
= e
j
A.
The fact that the
A
decomposes as a direct sum of the
P
j
translates to the
condition
X
e
j
= 1, e
i
e
j
= 0
for i 6= j.
Definition
(Orthogonal idempotent)
.
A collection of idempotents
{e
i
}
is or-
thogonal if e
i
e
j
= 0 for i 6= j.
The indecomposability of P
j
is equivalent to e
j
being primitive:
Definition
(Primitive idempotent)
.
An idempotent is primitive if it cannot be
expressed as a sum
e = e
0
+ e
00
,
where e
0
, e
00
are orthogonal idempotents, both non-zero.
We see that giving a direct sum decomposition of
A
is equivalent to finding
an orthogonal collection of primitive idempotents that sum to 1. This is rather
useful, because idempotents are easier to move around that projectives.
Our current plan is as follows — given an Artinian algebra
A
, we can quotient
out by
J
(
A
), and obtain a semi-simple algebra
A/J
(
A
). By Artin–Wedderburn,
we know how we can decompose
A/J
(
A
), and we hope to be able to lift this
decomposition to one of
A
. The point of talking about idempotents instead is
that we know what it means to lift elements.
Proposition.
Let
N
be a nilpotent ideal in
A
, and let
f
be an idempotent of
A/N
¯
A. Then there is an idempotent e A with f = ¯e.
In particular, we know
J
(
A
) is nilpotent, and this proposition applies. The
proof involves a bit of magic.
Proof.
We consider the quotients
A/N
i
for
i
1. We will lift the idempotents
successively as we increase
i
, and since
N
is nilpotent, repeating this process
will eventually land us in A.
Suppose we have found an idempotent
f
i1
A/N
i1
with
¯
f
i1
=
f
. We
want to find f
i
A/N
i
such that
¯
f
i
= f.
For
i >
1, we let
x
be an element of
A/N
i
with image
f
i1
in
A/N
i1
.
Then since
x
2
x
vansishes in
A/N
i1
, we know
x
2
x N
i1
/N
i
. Then in
particular,
(x
2
x)
2
= 0 A/N
i
. ()
We let
f
i
= 3x
2
2x
3
.
Then by a direct computation using (
), we find
f
2
i
=
f
i
, and
f
i
has image
3
f
i1
2
f
i1
=
f
i1
in
A/N
i1
(alternatively, in characteristic
p
, we can use
f
i
= x
p
). Since N
k
= 0 for some k, this process gives us what we want.
Just being able to lift idempotents is not good enough. We want to lift
decompositions as projective indecomposables. So we need to do better.
Corollary. Let N be a nilpotent ideal of A. Let
¯
1 = f
1
+ · · · + f
r
with {f
i
} orthogonal primitive idempotents in A/N. Then we can write
1 = e
1
+ · · · + e
r
,
with {e
i
} orthogonal primitive idempotents in A, and ¯e
i
= f
i
.
Proof. We define a sequence e
0
i
A inductively. We set
e
0
1
= 1.
Then for each
i >
1, we pick
e
0
i
a lift of
f
i
+
· · ·
+
f
t
e
0
i1
Ae
0
i1
, since by
inductive hypothesis we know that f
i
+ · · · + f
t
e
0
i1
Ae
0
i1
/N . Then
e
0
i
e
0
i+1
= e
0
i+1
= e
0
i+1
e
0
i
.
We let
e
i
= e
0
i
e
0
i+1
.
Then
¯e
i
= f
i
.
Also, if j > i, then
e
j
= e
0
i+1
e
j
e
0
i+1
,
and so
e
i
e
j
= (e
0
i
e
0
i+1
)e
0
i+1
e
j
e
0
i+1
= 0.
Similarly e
j
e
i
= 0.
We now apply this lifting of idempotents to
N
=
J
(
A
), which we know is
nilpotent. We know
A/N
is the direct sum of simple modules, and thus the
decomposition corresponds to
¯
1 = f
1
+ · · · + f
t
A/J(A),
and these
f
i
are orthogonal primitive idempotents. Idempotent lifting then gives
1 = e
1
+ · · · + e
t
A,
and these are orthogonal primitive idempotents. So we can write
A =
M
e
j
A =
M
P
i
,
where
P
i
=
e
i
A
are indecomposable projectives, and
P
i
/P
i
J
(
A
) =
S
i
is simple.
By Krull–Schmidt, any indecomposable projective isomorphic to one of these
P
j
.
The final piece of the picture is to figure out when two indecomposable
projectives lie in the same block. Recall that if
M
is a right
A
-module and
e
is
idempotent, then
Me
=
Hom
A
(eA, M ).
In particular, if M = fA for some idempotent f, then
Hom(eA, f A)
=
fAe.
However, if e and f are in different blocks, say B
1
and B
2
, then
fAe B
1
B
2
= 0,
since B
1
and B
2
are (two-sided!) ideals. So we know
Hom(eA, f A) = 0.
So if
Hom
(
eA, f A
)
6
= 0, then they are in the same block. The existence of a
homomorphism can alternatively be expressed in terms of composition factors.
We have seen that each indecomposable projective P has a simple “top”
P/P J(A)
=
S.
Definition
(Composition factor)
.
A simple module
S
is a composition factor
of a module M if there are submodules M
1
M
2
with
M
2
/M
1
=
S.
Suppose
S
is a composition factor of a module
M
. Then we have a diagram
P
M
2
S 0
So by definition of projectivity, we obtain a non-zero diagonal map
P M
2
M
as shown.
Lemma.
Let
P
be an indecomposable projective, and
M
an
A
-module. Then
Hom(P, M) 6= 0 iff P/P J(A) is a composition factor of M .
Proof.
We have proven
. Conversely, suppose there is a non-zero map
f
:
P
M. Then it factors as
S =
P
P J(A)
im f
(im f )J(A)
.
Now we cannot have
im f
= (
im f
)
J
(
A
), or else we have
im f
= (
im f
)
J
(
A
)
n
= 0
for sufficiently large
n
since
J
(
A
) is nilpotent. So this map must be injective,
hence an isomorphism. So this exhibits S as a composition factor of M .
We define a (directed) graph whose vertices are labelled by indecomposable
projectives, and there is an edge
P
1
P
2
if the top
S
1
of
P
1
is a composition
factor of P
2
.
Theorem.
Indecomposable projectives
P
1
and
P
2
are in the same block if and
only if they lie in the same connected component of the graph.
Proof.
It is clear that
P
1
and
P
2
are in the same connected component, then
they are in the same block.
Conversely, consider a connected component X, and consider
I =
M
P X
P.
We show that this is in fact a left ideal, hence an ideal. Consider any
x A
. Then
for each P X, left-multiplication gives a map P A, and if we decompose
A =
M
P
i
,
then this can be expressed as a sum of maps
f
i
:
P P
i
. Now such a map can
be non-zero only if
P
is a composition factor of
P
i
. So if
f
i
6
= 0, then
P
i
X
.
So left-multiplication by
x
maps
I
to itself, and it follows that
I
is an ideal.
1.5 K
0
We now briefly talk about the notion of K
0
.
Definition
(
K
0
)
.
For any associative
k
-algebra
A
, consider the free abelian
group with basis labelled by the isomorphism classes [
P
] of finitely-generated
projective A-modules. Then introduce relations
[P
1
] + [P
2
] = [P
1
P
2
],
This yields an abelian group which is the quotient of the free abelian group by
the subgroup generated by
[P
1
] + [P
2
] [P
1
P
2
].
The abelian group is K
0
(A).
Example.
If
A
is an Artinian algebra, then we know that any finitely-generated
projective is a direct sum of indecomposable projectives, and this decomposition
is unique by Krull-Schmidt. So
K
0
(A) =
abelian group generated by the isomorphism
classes of indecomposable projectives
.
So
K
0
(
A
)
=
Z
r
, where
r
is the number of isomorphism classes of indecomposable
projectives, which is the number of isomorphism classes of simple modules.
Here we’re using the fact that two indecomposable projectives are isomorphic
iff their simple tops are isomorphic.
It turns out there is a canonical map
K
0
(
A
)
A/
[
A, A
]. Recall we have
met
A/
[
A, A
] when we were talking about the number of simple modules. We
remarked that it was the 0th Hochschild homology group, and when
A
=
kG
,
there is a
k
-basis of
A/
[
A, A
] given by
g
i
+ [
A, A
], where
g
i
are conjugacy class
representatives.
To construct this canonical map, we first look at the trace map
M
n
(A) A/[A, A].
This is a
k
-linear map, invariant under conjugation. We also note that the
canonical inclusion
M
n
(A) M
n+1
(A)
X 7→
X 0
0 0
is compatible with the trace map. We observe that the trace induces an isomor-
phism
M
n
(A)
[M
n
(A), M
n
(A)]
A
[A, A]
,
by linear algebra.
Now if
P
is finitely generated projective. It is a direct summand of some
A
n
.
Thus we can write
A
n
= P Q,
for
P, Q
projective. Moreover, projection onto
P
corresponds to an idempotent
e in M
n
(A) = End
A
(A
n
), and that
P = e(A
n
).
and we have
End
A
(P ) = eM
n
(A)e.
Any other choice of idempotent yields an idempotent
e
1
conjugate to
e
in
M
2n
(
A
).
Therefore the trace of an endomorphism of
P
is well-defined in
A/
[
A, A
],
independent of the choice of e. Thus we have a trace map
End
A
(P ) A/[A, A].
In particular, the trace of the identity map on
P
is the trace of
e
. We call this
the trace of P .
Note that if we have finitely generated projectives
P
1
and
P
2
, then we have
P
1
Q
1
= A
n
P
2
Q
2
= A
m
Then we have
(P
1
P
2
) (Q
1
Q
2
) = A
m+n
.
So we deduce that
tr(P
1
P
2
) = tr P
1
+ tr P
2
.
Definition
(Hattori-Stallings trace map)
.
The map
K
0
(
A
)
A/
[
A, A
] induced
by the trace is the Hattori–Stallings trace map.
Example.
Let
A
=
kG
, and
G
be finite. Then
A/
[
A, A
] is a
k
-vector space
with basis labelled by a set of conjugacy class representatives
{g
i
}
. Then we
know, for a finitely generated projective P , we can write
tr P =
X
r
P
(g
i
)g
i
,
where the
r
p
(
g
i
) may be regarded as class functions. However,
P
may be regarded
as a k-vector space.
So the there is a trace map
End
K
(P ) k,
and also the “character”
χ
p
:
G k
, where
χ
P
(
g
) =
tr g
. Hattori proved that if
C
G
(g) is the centralizer of g G, then
χ
p
(g) = |C
G
(G)|r
p
(g
1
). ()
If
char k
= 0 and
k
is is algebraically closed, then we know
kG
is semi-simple.
So every finitely generated projective is a direct sum of simples, and
K
0
(kG)
=
Z
r
with r the number of simples, and () implies that the trace map
Z
r
=
K
0
(kG)
kG
[kG, kG]
=
k
r
is the natural inclusion.
This is the start of the theory of algebraic
K
-theory, which is a homology
theory telling us about the endomorphisms of free
A
-modules. We can define
K
1
(A) to be the abelianization of
GL(A) = lim
n→∞
GL
n
(A).
K
2
(
A
) tells us something about the relations required if you express
GL
(
A
) in
terms of generators and relations. We’re being deliberately vague. These groups
are very hard to compute.
Just as we saw in the i = 0 case, there are canonical maps
K
i
(A) HH
i
(A),
where
HH
is the Hochschild homology. The
i
= 1 case is called the Dennis
trace map. These are analogous to the Chern maps in topology.
2 Noetherian algebras
2.1 Noetherian algebras
In the introduction, we met the definition of Noetherian algebras.
Definition
(Noetherian algebra)
.
An algebra is left Noetherian if it satisfies
the ascending chain condition (ACC ) on left ideals, i.e. if
I
1
I
2
I
3
· · ·
is an ascending chain of left ideals, then there is some
N
such that
I
N+m
=
I
N
for all m 0.
Similarly, we say an algebra is Noetherian if it is both left and right Noethe-
rian.
We’ve also met a few examples. Here we are going to meet lots more. In
fact, most of this first section is about establishing tools to show that certain
algebras are Noetherian.
One source of Noetherian algebras is via constructing polynomial and power
series rings. Recall that in IB Groups, Rings and Modules, we proved the Hilbert
basis theorem:
Theorem
(Hilbert basis theorem)
.
If
A
is Noetherian, then
A
[
X
] is Noetherian.
Note that our proof did not depend on
A
being commutative. The same
proof works for non-commutative rings. In particular, this tells us
k
[
X
1
, · · · , X
n
]
is Noetherian.
It is also true that power series rings of Noetherian algebras are also Noethe-
rian. The proof is very similar, but for completeness, we will spell it out
completely.
Theorem. Let A be left Noetherian. Then A[[X]] is Noetherian.
Proof.
Let
I
be a left ideal of
A
[[
X
]]. We’ll show that if
A
is left Noetherian,
then I is finitely generated. Let
J
r
= {a : there exists an element of I of the form aX
r
+ higher degree terms}.
We note that J
r
is a left ideal of A, and also note that
J
0
J
1
J
2
J
3
· · · ,
as we can always multiply by
X
. Since
A
is left Noetherian, this chain terminates
at
J
N
for some
N
. Also,
J
0
, J
1
, J
2
, · · · , J
N
are all finitely generated left ideals.
We suppose
a
i1
, · · · , a
is
i
generates
J
i
for
i
= 1
, · · · , N
. These correspond to
elements
f
ij
(X) = a
ij
X
j
+ higher odder terms I.
We show that this finite collection generates
I
as a left ideal. Take
f
(
X
)
I
,
and suppose it looks like
b
n
X
n
+ higher terms,
with b
n
6= 0.
Suppose n < N. Then b
n
J
n
, and so we can write
b
n
=
X
c
nj
a
nj
.
So
f(X)
X
c
nj
f
nj
(X) I
has zero coefficient for X
n
, and all other terms are of higher degree.
Repeating the process, we may thus wlog
n N
. We get
f
(
X
) of the form
d
N
X
N
+ higher degree terms. The same process gives
f(X)
X
c
Nj
f
Nj
(X)
with terms of degree
N
+ 1 or higher. We can repeat this yet again, using the
fact J
N
= J
N+1
, so we obtain
f(X)
X
c
Nj
f
Nj
(x)
X
d
N+1,j
Xf
Nj
(X) + · · · .
So we find
f(X) =
X
e
j
(X)f
Nj
(X)
for some
e
j
(
X
). So
f
is in the left ideal generated by our list, and hence so is
f.
Example.
It is straightforward to see that quotients of Noetherian algebras are
Noetherian. Thus, algebra images of the algebras
A
[
x
] and
A
[[
x
]] would also be
Noetherian.
For example, finitely-generated commutative
k
-algebras are always Noethe-
rian. Indeed, if we have a generating set
x
i
of
A
as a
k
-algebra, then there is an
algebra homomorphism
k[X
1
, · · · , X
n
] A
X
i
x
i
We also saw previously that
Example. Any Artinian algebra is Noetherian.
The next two examples we are going to see are less obviously Noetherian,
and proving that they are Noetherian takes some work.
Definition
(
n
th Weyl algebra)
.
The
n
th Weyl algebra
A
n
(
k
) is the algebra
generated by X
1
, · · · , X
n
, Y
1
, · · · , Y
n
with relations
Y
i
X
i
X
i
Y
i
= 1,
for all i, and everything else commutes.
This algebra acts on the polynomial algebra
k
[
X
1
, · · · , X
n
] with
X
i
acting
by left multiplication and
Y
i
=
X
i
. Thus
k
[
X
1
, · · · , X
n
] is a left
A
n
(
k
) module.
This is the prototype for thinking about differential algebras, and
D
-modules in
general (which we will not talk about).
The other example we have is the universal enveloping algebra of a Lie
algebra.
Definition
(Universal enveloping algebra)
.
Let
g
be a Lie algebra over
k
, and
take a
k
-vector space basis
x
1
, · · · , x
n
. We form an associative algebra with
generators x
1
, · · · , x
n
with relations
x
i
x
j
x
j
x
i
= [x
i
, x
j
],
and this is the universal enveloping algebra U(g).
Example.
If
g
is abelian, i.e. [
x
i
, x
j
] = 0 in
g
, then the enveloping algebra is
the polynomial algebra in x
1
, · · · , x
n
.
Example. If g = sl
2
(k), then we have a basis
0 1
0 0
,
0 0
1 0
,
1 0
0 1
.
They satisfy
[e, f ] = h, [h, e] = 2e, [h, f ] = 2f,
To prove that
A
n
(
k
) and
U
(
g
) are Noetherian, we need some machinery, that
involves some “deformation theory”. The main strategy is to make use of a
natural filtration of the algebra.
Definition
(Filtered algebra)
.
A (
Z
-)filtered algebra
A
is a collection of
k
-vector
spaces
· · · A
1
A
0
A
1
A
2
· · ·
such that A
i
· A
j
A
i+j
for all i, j Z, and 1 A
0
.
For example a polynomial ring is naturally filtered by the degree of the
polynomial.
The definition above was rather general, and often, we prefer to talk about
more well-behaved filtrations.
Definition (Exhaustive filtration). A filtration is exhaustive if
S
A
i
= A.
Definition (Separated filtration). A filtration is separated if
T
A
i
= {0}.
Unless otherwise specified, our filtrations are exhaustive and separated.
For the moment, we will mostly be interested in positive filtrations.
Definition (Positive filtration). A filtration is positive if A
i
= 0 for i < 0.
Our canonical source of filtrations is the following construction:
Example. If A is an algebra generated by x
1
, · · · , x
n
, say, we can set
A
0
is the k-span of 1
A
1
is the k-span of 1, x
1
, · · · , x
n
A
1
is the k-span of 1, x
1
, · · · , x
n
, x
i
x
j
for i, j {1, · · · , n}.
In general,
A
r
is elements that are of the form of a (non-commutative) polynomial
expression of degree r.
Of course, the filtration depends on the choice of the generating set.
Often, to understand a filtered algebra, we consider a nicer object, known as
the associated graded algebra.
Definition
(Associated graded algebra)
.
Given a filtration of
A
, the associated
graded algebra is the vector space direct sum
gr A =
M
A
i
A
i1
.
This is given the structure of an algebra by defining multiplication by
(a + A
i1
)(b + A
j1
) = ab + A
i+j1
A
i+j
A
i+j1
.
In our example of a finitely-generated algebra, the graded algebra is generated
by x
1
+ A
0
, · · · , x
n
+ A
0
A
1
/A
0
.
The associated graded algebra has the natural structure of a graded algebra:
Definition
(Graded algebra)
.
A (
Z
-)graded algebra is an algebra
B
that is of
the form
B =
M
iZ
B
i
,
where
B
i
are
k
-subspaces, and
B
i
B
j
B
i+j
. The
B
i
’s are called the homogeneous
components.
A graded ideal is an ideal of the form
M
J
i
,
where J
i
is a subspace of B
i
, and similarly for left and right ideals.
There is an intermediate object between a filtered algebra and its associated
graded algebra, known as the Rees algebra.
Definition
(Rees algebra)
.
Let
A
be a filtered algebra with filtration
{A
i
}
. Then
the Rees algebra
Rees
(
A
) is the subalgebra
L
A
i
T
i
of the Laurent polynomial
algebra A[T, T
1
] (where T commutes with A).
Since 1 A
0
A
1
, we know T Rees(A). The key observation is that
Rees(A)/(T )
=
gr A.
Rees(A)/(1 T )
=
A.
Since
A
n
(
k
) and
U
(
g
) are finitely-generated algebras, they come with a
natural filtering induced by the generating set. It turns out, in both cases, the
associated graded algebras are pretty simple.
Example.
Let
A
=
A
n
(
k
), with generating set
X
1
, · · · , X
n
and
Y
1
, · · · , Y
n
. We
take the filtration as for a finitely-generated algebra. Now observe that if
a
i
A
i
,
and a
j
A
j
, then
a
i
a
j
a
j
a
i
A
i+j2
.
So we see that gr A is commutative, and in fact
gr A
n
(k)
=
k[
¯
X
1
, · · · ,
¯
X
n
,
¯
Y
1
, · · · ,
¯
Y
n
],
where
¯
X
i
,
¯
Y
i
are the images of
X
i
and
Y
i
in
A
1
/A
0
respectively. This is not
hard to prove, but is rather messy. It requires a careful induction.
Example.
Let
g
be a Lie algebra, and consider
A
=
U
(
g
). This has generating
set
x
1
, · · · , x
n
, which is a vector space basis for
g
. Again using the filtration for
finitely-generated algebras, we get that if a
i
A
i
and a
j
A
j
, then
a
i
a
j
a
j
a
i
A
i+j1
.
So again gr A is commutative. In fact, we have
gr A
=
k[¯x
1
, · · · , ¯x
n
].
The fact that this is a polynomial algebra amounts to the same as the Poincar´e-
Birkhoff-Witt theorem. This gives a k-vector space basis for U(g).
In both cases, we find that
gr A
are finitely-generated and commutative, and
therefore Noetherian. We want to use this fact to deduce something about
A
itself.
Lemma.
Let
A
be a positively filtered algebra. If
gr A
is Noetherian, then
A
is
left Noetherian.
By duality, we know that A is also right Noetherian.
Proof. Given a left ideal I of A, we can form
gr I =
M
I A
i
I A
i1
,
where I is filtered by {I A
i
}. By the isomorphism theorem, we know
I A
i
I A
i1
=
I A
i
+ A
i1
A
i1
A
i
A
i1
.
Then gr I is a left graded ideal of gr A.
Now suppose we have a strictly ascending chain
I
1
< I
2
< · · ·
of left ideals. Since we have a positive filtration, for some
A
i
, we have
I
1
A
i
(
I
2
A
i
and I
1
A
i1
= I
2
A
i1
. Thus
gr I
1
( gr I
2
( gr I
3
( · · · .
This is a contradiction since
gr A
is Noetherian. So
A
must be Noetherian.
Where we need positivity is the existence of that transition from equality to
non-equality. If we have a
Z
-filtered algebra instead, then we need to impose
some completeness assumption, but we will not go into that.
Corollary. A
n
(k) and U(g) are left/right Noetherian.
Proof. gr A
n
(
k
) and
gr U
(
g
) are commutative and finitely generated algebras.
Note that there is an alternative filtration for
A
n
(
k
) yielding a commutative
associated graded algebra, by setting A
0
= k[X
1
, · · · , X
n
] and
A
1
= k[X
1
, · · · , X
n
] +
n
X
j=1
k[X
1
, · · · , X
n
]Y
j
,
i.e. linear terms in the
Y
, and then keep on going. Essentially, we are filtering on
the degrees of the
Y
i
only. This also gives a polynomial algebra as an associative
graded algebra. The main difference is that when we take the commutator,
we don’t go down by two degrees, but one only. Later, we will see this is
advantageous when we want to get a Poisson bracket on the associated graded
algebra.
We can look at further examples of Noetherian algebras.
Example.
The quantum plane
k
q
[
X, Y
] has generators
X
and
Y
, with relation
XY = qY X
for some
q k
×
. This thing behaves differently depending on whether
q
is a
root of unity or not.
This quantum plane first appeared in mathematical physics.
Example.
The quantum torus
k
q
[
X, X
1
, Y, Y
1
] has generators
X
,
X
1
,
Y
,
Y
1
with relations
XX
1
= Y Y
1
= 1, XY = qY X.
The word “quantum” in this context is usually thrown around a lot, and
doesn’t really mean much apart from non-commutativity, and there is very little
connection with actual physics.
These algebras are both left and right Noetherian. We cannot prove these
by filtering, as we just did. We will need a version of Hilbert’s basis theorem
which allows twisting of the coefficients. This is left as an exercise on the second
example sheet.
In the examples of
A
n
(
k
) and
U
(
g
), the associated graded algebras are
commutative. However, it turns out we can still capture the non-commutativity
of the original algebra by some extra structure on the associated graded algebra.
So suppose
A
is a (positively) filtered algebra whose associated graded algebra
gr A
is commutative. Recall that the filtration has a corresponding Rees algebra,
and we saw that Rees A/(T )
=
gr A. Since gr A is commutative, this means
[Rees A, Rees A] (T ).
This induces a map
Rees
(
A
)
× Rees
(
A
)
(
T
)
/
(
T
2
), sending (
r, s
)
7→ T
2
+ [
r, s
].
Quotienting out by (T ) on the left, this gives a map
gr A × gr A
(T )
(T
2
)
.
We can in fact identify the right hand side with gr A as well. Indeed, the map
gr A
=
Rees(A)
(T )
(T )
(T
2
)
mult. by T
,
is an isomorphism of gr A
=
Rees A/(T )-modules. We then have a bracket
{ · , · } : gr A × gr A gr A
(¯r, ¯s) {r, s}
.
Note that in our original filtration of the Weyl algebra
A
n
(
k
), since the commu-
tator brings us down by two degrees, this bracket vanishes identically, but using
the alternative filtration does not give a non-zero { · , · }.
This { · , · } is an example of a Poisson bracket.
Definition
(Poisson algebra)
.
An associative algebra
B
is a Poisson algebra if
there is a k-bilinear bracket { · , · } : B × B B such that
B is a Lie algebra under { · , · }, i.e.
{r, s} = −{s, r}
and
{{r, s}, t} + {{s, t}, r} + {{t, r}, s} = 0.
We have the Leibnitz rule
{r, st} = s{r, t} + {r, s}t.
The second condition says {r, · } : B B is a derivation.
2.2 More on A
n
(k) and U(g)
Our goal now is to study modules of
A
n
(
k
). The first result tells us we must
focus on infinite dimensional modules.
Lemma.
Suppose
char k
= 0. Then
A
n
(
k
) has no non-zero modules that are
finite-dimensional k-vector spaces.
Proof.
Suppose
M
is a finite-dimensional module. Then we’ve got an algebra
homomorphism θ : A
n
(k) End
k
(M)
=
M
m
(k), where m = dim
k
M.
In A
n
(k), we have
Y
1
X
1
X
1
Y
1
= 1.
Applying the trace map, we know
tr(θ(Y
1
)θ(X
1
) θ(X
1
)θ(Y
1
)) = tr I = m.
But since the trace is cyclic, the left hand side vanishes. So
m
= 0. So
M
is
trivial.
A similar argument works for the quantum torus, but using determinants
instead.
We’re going to make use of our associated graded algebras from last time,
which are isomorphic to polynomial algebras. Given a filtration
{A
i
}
of
A
, we
may filter a module with generating set S by setting
M
i
= A
i
S.
Note that
A
j
M
i
M
i+j
,
which allows us to form an associated graded module
gr M =
M
M
i
M
i+1
.
This is a graded
gr A
-module, which is finitely-generated if
M
is. So we’ve got a
finitely-generated graded module over a graded commutative algebra.
To understand this further, we prove some results about graded modules
over commutative algebras, which is going to apply to our gr A and gr M.
Definition
(Poincar´e series)
.
Let
V
be a graded module over a graded algebra
S, say
V =
M
i=0
V
i
.
Then the Poincar´e series is
P (V, t) =
X
i=0
(dim V
i
)t
i
.
Theorem
(Hilbert-Serre theorem)
.
The Poincar´e series
P
(
V, t
) of a finitely-
generated graded module
V =
M
i=0
V
i
over a finitely-generated generated commutative algebra
S =
M
i=0
S
i
with homogeneous generating set x
1
, · · · , x
m
is a rational function of the form
f(t)
Q
(1 t
k
i
)
,
where f (t) Z[t] and k
i
is the degree of the generator x
i
.
Proof.
We induct on the number
m
of generators. If
m
= 0, then
S
=
S
0
=
k
,
and
V
is therefore a finite-dimensional
k
-vector space. So
P
(
V, t
) is a polynomial.
Now suppose
m >
0. We assume the theorem is true for
< m
generators.
Consider multiplication by x
m
. This gives a map
V
i
V
i+k
m
x
m
,
and we have an exact sequence
0 K
i
V
i
V
i+k
m
L
i+k
m
0,
x
m
()
where
K =
M
K
i
= ker(x
m
: V V )
and
L =
M
L
i+k
m
= coker(x
m
: V V ).
Then
K
is a graded submodule of
V
and hence is a finitely-generated
S
-module,
using the fact that
S
is Noetherian. Also,
L
=
V/x
m
V
is a quotient of
V
, and it
is thus also finitely-generated.
Now both
K
and
L
are annihilated by
x
m
. So they may be regarded as
S
0
[x
1
, · · · , x
m1
]-modules. Applying dim
k
to (), we know
dim
k
(K
i
) dim
k
(V
i
) + dim(V
i+k
m
) dim(L
i+k
m
) = 0.
We multiply by t
i+k
m
, and sum over i to get
t
k
m
P (K, t) t
k
m
P (V, t) + P (V, t) P (L, t) = g(t),
where
g
(
t
) is a polynomial with integral coefficients arising from consideration
of the first few terms.
We now apply the induction hypothesis to K and L, and we are done.
Corollary.
If each
k
1
, · · · , k
m
= 1, i.e.
S
is generated by
S
0
=
k
and homoge-
neous elements
x
1
, · · · , x
m
of degree 1, then for large enough
i
, then
dim V
i
=
φ
(
i
)
for some polynomial
φ
(
t
)
Q
[
t
] of
d
1, where
d
is the order of the pole of
P (V, t) at t = 1. Moreover,
i
X
j=0
dim V
j
= χ(i),
where χ(t) Q[t] of degree d.
Proof. From the theorem, we know that
P (V, t) =
f(t)
(1 t)
d
,
for some d with f (1) 6= 0, f Z[t]. But
(1 t)
1
= 1 + t + t
2
+ · · ·
By differentiating, we get an expression
(1 t)
d
=
X
d + i 1
d 1
t
i
.
If
f(t) = a
0
+ a
1
t + · · · + a
s
t
s
,
then we get
dim V
i
= a
0
d + i 1
d 1
+ a
1
d + i 2
d 1
+ · · · + a
s
d + i s 1
d 1
,
where we set
r
d1
= 0 if
r < d
1, and this expression can be rearranged to
give φ(i) for a polynomial φ(t) Q[t], valid for i s > 0. In fact, we have
φ(t) =
f(1)
(d 1)!
t
d1
+ lower degree term.
Since f (1) 6= 0, this has degree d 1.
This implies that
i
X
j=0
dim V
j
is a polynomial in Q[t] of degree d.
This
φ
(
t
) is the Hilbert polynomial, and
χ
(
t
) the Samuel polynomial. Some
people call χ(t) the Hilbert polynomial instead, though.
We now want to apply this to our cases of
gr A
, where
A
=
A
n
(
k
) or
U
(
g
),
filtered as before. Then we deduce that
i
X
0
dim
M
j
M
j1
= χ(i),
for a polynomial χ(t) Q[t]. But we also know
i
X
j=0
dim
M
j
M
j1
= dim M
i
.
We are now in a position to make a definition.
Definition
(Gelfand-Kirillov dimension)
.
Let
A
=
A
n
(
k
) or
U
(
g
) and
M
a
finitely-generated
A
-module, filtered as before. Then the Gelfand-Kirillov di-
mension
d
(
M
) of
M
is the degree of the Samuel polynomial of
gr M
as a
gr A-module.
This makes sense because
gr A
is a commutative algebra in this case. A
priori, it seems like this depends on our choice of filtering on
M
, but actually, it
doesn’t. For a more general algebra, we can define the dimension as below:
Definition
(Gelfand-Kirillov dimension)
.
Let
A
be a finitely-generated
k
-
algebra, which is filtered as before, and a finitely-generated
A
-module
M
, filtered
as before. Then the GK-dimension of M is
d(M) = lim sup
n→∞
log(dim M
n
)
log n
.
In the case of
A
=
A
n
(
k
) or
U
(
g
), this matches the previous definition. Again,
this does not actually depend on the choice of generating sets.
Recall we showed that no non-zero
A
n
(
k
)-module
M
can have finite dimension
as a
k
-vector space. So we know
d
(
M
)
>
0. Also, we know that
d
(
M
) is an
integer for cases
A
=
A
n
or
U
(
g
), since it is the degree of a polynomial. However,
for general
M
=
A
, we can get non-integral values. In fact, the values we can
get are 0
,
1
,
2, and then any real number
2. We can also have
if the
lim sup
doesn’t exist.
Example.
If
A
=
kG
, then we have
GK
-
dim
(
kG
)
<
iff
G
has a subgroup
H
of finite index with
H
embedding into the strictly upper triangular integral
matrices, i.e. matrices of the form
1 ∗ · · · ∗
0 1 · · · ∗
.
.
.
.
.
.
.
.
.
.
.
.
0 0 · · · 1
.
This is a theorem of Gromoll, and is quite hard to prove.
Example.
We have
GK
-
dim
(
A
) = 0 iff
A
is finite-dimensional as a
k
-vector
space.
We have
GK-dim(k[X]) = 1,
and in general
GK-dim(k[X
1
, · · · , X
n
]) = n.
Indeed, we have
dim
k
(mth homogeneous component) =
m + n
n
.
So we have
χ(t) =
t + n
n
This is of degree n, with leading coefficient
1
n!
.
We can make the following definition, which we will not use again:
Definition
(Multiplicity)
.
Let
A
be a commutative algebra, and
M
an
A
-module.
The multiplicity of M with d(M) = d is
d! × leading coefficient of χ(t).
On the second example sheet, we will see that the multiplicity is integral.
We continue looking at more examples.
Example.
We have
d
(
A
n
(
k
)) = 2
n
, and
d
(
U
(
g
)) =
dim
k
g
. Here we are using
the fact that the associated graded algebras are polynomial algebras.
Example.
We met
k
[
X
1
, · · · , X
n
] as the “canonical”
A
n
(
k
)-module. The filtra-
tion of the module matches the one we used when thinking about the polynomial
algebra as a module over itself. So we get
d(k[X
1
, · · · , X
n
]) = n.
Lemma. Let M be a finitely-generated A
n
-module. Then d(M ) 2n.
Proof.
Take generators
m
1
, · · · , m
s
of
M
. Then there is a surjective filtered
module homomorphism
A
n
· · · A
n
M
(a
1
, · · · , a
s
)
P
a
i
m
i
It is easy to see that quotients can only reduce dimension, so
GK-dim(M) d(A
n
· · · A
n
).
But
χ
A
n
⊕···⊕A
n
=
A
n
has degree 2n.
More interestingly, we have the following result:
Theorem
(Bernstein’s inequality)
.
Let
M
be a non-zero finitely-generated
A
n
(k)-module, and char k = 0. Then
d(M) n.
Definition
(Holonomic module)
.
An
A
n
(
k
) module
M
is holonomic iff
d
(
M
) =
n.
If we have a holonomic module, then we can quotient by a maximal submodule,
and get a simple holonomic module. For a long time, people thought all simple
modules are holonomic, until someone discovered a simple module that is not
holonomic. In fact, most simple modules are not holonomic, but we something
managed to believe otherwise.
Proof.
Take a generating set and form the canonical filtrations
{A
i
}
of
A
n
(
k
)
and
{M
i
}
of
M
. We let
χ
(
t
) be the Samuel polynomial. Then for large enough
i, we have
χ(i) = dim M
i
.
We claim that
dim A
i
dim Hom
k
(M
i
, M
2i
) = dim M
i
× dim M
2i
.
Assuming this, for large enough i, we have
dim A
i
χ(i)χ(2i).
But we know
dim A
i
=
i + 2
2n
,
which is a polynomial of degree 2
n
. But
χ
(
t
)
χ
(2
t
) is a polynomial of degree
2d(M). So we get that
n d(M ).
So it remains to prove the claim. It suffices to prove that the natural map
A
i
Hom
k
(M
i
, M
2i
),
given by multiplication is injective.
So we want to show that if
a A
i
6
= 0, then
aM
i
6
= 0. We prove this by
induction on
i
. When
i
= 0, then
A
0
=
k
, and
M
0
is a finite-dimensional
k
-vector
space. Then the result is obvious.
If
i >
0, we suppose the result is true for smaller
i
. We let
a A
i
is non-zero.
If aM
i
= 0, then certainly a 6∈ k. We express
a =
X
c
αβ
X
α
1
1
X
α
2
2
· · · X
α
n
n
Y
β
1
1
· · · Y
β
n
n
,
where α = (α
1
, · · · , α
n
), β = (β
1
, · · · , β
n
), and c
α,β
k.
If possible, pick a
j
such that
c
α,α
6
= 0 for some
α
with
α
j
6
= 0 (this happens
when there is an X involved). Then
[Y
j
, a] =
X
α
j
c
α,β
X
α
1
1
· · · X
α
j
1
j
· · · X
α
n
n
Y
β
1
1
· · · Y
β
n
n
,
and this is non-zero, and lives in A
i1
.
If aM
i
= 0, then certainly aM
i1
= 0. Hence
[Y
j
, a]M
i1
= (Y
j
a aY
j
)M
i1
= 0,
using the fact that Y
j
M
i1
M
i
. This is a contradiction.
If a only has Y ’s involved, then we do something similar using [X
j
, a].
There is also a geometric way of doing this.
We take k = C. We know gr A
n
is a polynomial algebra
gr A
n
= k[
¯
X
1
, · · · ,
¯
X
n
,
¯
Y
1
, · · · ,
¯
Y
n
],
which may be viewed as the coordinate algebra of the cotangent bundle on affine
n
-space
C
n
. The points of this correspond to the maximal ideals of
gr A
n
. If
I
is a left ideal of
A
n
(
C
), then we can form
gr I
and we can consider the set of
maximal ideals containing it. This gives us the characteristic variety
Ch
(
A
n
/I
).
We saw that there was a Poisson bracket on
gr A
n
, and this may be used to
define a skew-symmetric form on the tangent space at any point of the cotangent
bundle. In this case, this is non-degenerate skew-symmetric form.
We can consider the tangent space
U
of
Ch
(
A
n
/I
) at a non-singular point,
and there’s a theorem of Gabber (1981) which says that
U U
, where
is
with respect to the skew-symmetric form. By non-degeneracy, we must have
dim U n, and we also know that
dim Ch(A
n
/I) = d(A
n
/I).
So we find that d(A
n
/I) n.
In the case of
A
=
U
(
g
), we can think of
gr A
as the coordinate algebra on
g
, the vector space dual of
g
. The Poisson bracket leads to a skew-symmetric
form on tangent spaces at points of
g
. In this case, we don’t necessarily get
non-degeneracy. However, on
g
, we have the adjoint action of the corresponding
Lie group
G
, and this induces a co-adjoint action on
g
. Thus
g
is a disjoint
union of orbits. If we consider the induced skew-symmetric form on tangent
spaces of orbits (at non-singular points), then it is non-degenerate.
2.3 Injective modules and Goldie’s theorem
The goal of this section is to prove Goldie’s theorem.
Theorem
(Goldie’s theorem)
.
Let
A
be a right Noetherian algebra with no
non-zero ideals all of whose elements are nilpotent. Then
A
embeds in a finite
direct sum of matrix algebras over division algebras.
The outline of the proof is as follows given any
A
, we embed
A
in an
“injective hull”
E
(
A
). We will then find that similar to what we did in Artin–
Wedderburn, we can decompose
End
(
E
(
A
)) into a direct sum of matrix algebras
over division algebras. But we actually cannot. We will have to first quotient
End(E(A)) by some ideal I.
On the other hand, we do not actually have an embedding of
A
=
End
A
(
A
)
into
End
(
E
(
A
)). Instead, what we have is only a homomorphism
End
A
(
A
)
End
(
E
(
A
))
/I
, where we quotient out by the same ideal
I
. So actually the two
of our problems happen to cancel each other out.
We will then prove that the kernel of this map contains only nilpotent
elements, and then our hypothesis implies this homomorphism is indeed an
embedding.
We begin by first constructing the injective hull. This is going to involve
talking about injective modules, which are dual to the notion of projective
modules.
Definition
(Injective module)
.
An
A
-module
E
is injective if for every diagram
of A-module maps
0 M N
E
θ
φ
ψ
,
such that
θ
is injective, there exists a map
ψ
that makes the diagram commute.
Equivalently, Hom( · , E) is an exact functor.
Example. Take A = k. Then all k-vector spaces are injective k-modules.
Example. Take A = k[X]. Then k(X) is an injective k[X]-module.
Lemma. Every direct summand of an injective module is injective, and direct
products of injectives is injective.
Proof. Same as proof for projective modules.
Lemma. Every A-module may be embedded in an injective module.
We say the category of
A
-modules has enough injectives. The dual result for
projectives was immediate, as free modules are projective.
Proof.
Let
M
be a right
A
-module. Then
Hom
k
(
A, M
) is a right
A
-module via
(fa)(x) = f(ax).
We claim that Hom
k
(A, M ) is an injective module. Suppose we have
0 M
1
N
1
Hom
k
(A, M )
θ
φ
We consider the k-module diagram
0 M
1
N
1
M
θ
α
β
where
α
(
m
1
) =
φ
(
m
1
)(1). Since
M
is injective as a
k
-module, we can find the
β
such that α = βθ. We define ψ : N
1
Hom
k
(A, M ) by
ψ(n
1
)(x) = β(n
1
x).
It is straightforward to check that this does the trick. Also, we have an embedding
M Hom
k
(A, M ) by m 7→ (φ
n
: x 7→ mx).
The category theorist will write the proof in a line as
Hom
A
( · , Hom
k
(A, M ))
=
Hom
k
( ·
A
A, M )
=
Hom
k
( · , M ),
which is exact since M is injective as a k-module.
Note that neither the construction of
Hom
k
(
A, M
), nor the proof that it is
injective requires the right
A
-modules structure of
M
. All we need is that
M
is
an injective k-module.
Lemma.
An
A
-module is injective iff it is a direct summand of every extension
of itself.
Proof.
Suppose
E
is injective and
E
0
is an extension of
E
. Then we can form
the diagram
0 E E
0
E
id
ψ
,
and then by injectivity, we can find ψ. So
E
0
= E ker ψ.
Conversely, suppose
E
is a direct summand of every extension. But by the
previous lemma, we can embed
E
in an injective
E
0
. This implies that
E
is a
direct summand of E
0
, and hence injective.
There is some sort of “smallest” injective a module embeds into, and this
is called the injective hull, or injective envelope. This is why our injectives are
called
E
. The “smallness” will be captured by the fact that it is an essential
extension.
Definition
(Essential submodule)
.
An essential submodule
M
of an
A
-module
N
is one where
M V 6
=
{
0
}
for every non-zero submodule
V
of
N
. We say
N
is an essential extension of M.
Lemma. An essential extension of an essential extension is essential.
Proof.
Suppose
M < E < F
are essential extensions. Then given
N F
, we
know
N E 6
=
{
0
}
, and this is a submodule of
E
. So (
N E
)
M
=
N M 6
= 0.
So F is an essential extension of M.
Lemma. A maximal essential extension is an injective module.
Such maximal things exist by Zorn’s lemma.
Proof.
Let
E
be a maximal essential extension of
M
, and consider any embedding
E F
. We shall show that
E
is a direct summand of
F
. Let
S
be the set of
all non-zero submodules
V
of
F
with
V E
=
{
0
}
. We apply Zorn’s lemma to
get a maximal such module, say V
1
.
Then
E
embeds into
F/V
1
as an essential submodule. By transitivity of
essential extensions,
F/V
1
is an essential extension of
M
, but
E
is maximal. So
E
=
F/V
1
. In other words,
F = E V
1
.
We can now make the following definition:
Definition
(Injective hull)
.
A maximal essential extension of
M
is the injective
hull (or injective envelope) of M, written E(M ).
Proposition.
Let
M
be an
A
-module, with an inclusion
M I
into an injective
module. Then this extends to an inclusion E(M ) I.
Proof. By injectivity, we can fill in the diagram
0 M E(M)
I
ψ
.
We know
ψ
restricts to the identity on
M
. So
ker ψ M
=
{
0
}
. By Since
E
(
M
)
is essential, we must have ker ψ = 0. So E(M ) embeds into I.
Proposition.
Suppose
E
is an injective essential extension of
M
. Then
E
=
E(M). In particular, any two injective hulls are isomorphic.
Proof.
By the previous lemma,
E
(
M
) embeds into
E
. But
E
(
M
) is a maximal
essential extension. So this forces E = E(M).
Using what we have got, it is not hard to see that
Proposition.
E(M
1
M
2
) = E(M
1
) E(M
2
).
Proof.
We know that
E
(
M
1
)
E
(
M
2
) is also injective (since finite direct sums
are the same as direct products), and also
M
1
M
2
embeds in
E
(
M
1
)
E
(
M
2
).
So it suffices to prove this extension is essential.
Let V E(M
1
) E(M
2
). Then either V/E(M
1
) 6= 0 or V /E(M
2
) 6= 0.
We wlog it is the latter. Note that we can naturally view
V
E(M
2
)
E(M
1
) E(M
2
)
E(M
2
)
=
E(M
1
).
Since M
1
E(M
1
) is essential, we know
M
1
(V/E(M
2
)) 6= 0.
So there is some
m
1
+
m
2
V
such that
m
2
E
(
M
2
) and
m
1
M
1
. Now
consider
{m E(M
2
) : am
1
+ m V for some a A}.
This is a non-empty submodule of
E
(
M
2
), and so contains an element of
M
2
,
say n. Then we know am
1
+ n V (M
1
M
2
), and we are done.
The next two examples of injective hulls will be stated without proof:
Example. Take A = k[X], and M = k[X]. Then E(M) = k(X).
Example.
Let
A
=
k
[
X
] and
V
=
k
be the trivial module, where
X
acts by 0.
Then
E(V ) =
k[X, X
1
]
Xk[X]
,
which is a quotient of A-modules. We note V embeds in this as
V
=
k[X]
Xk[X]
k[X, X
1
]
Xk[X]
.
Definition
(Uniform module)
.
A non-zero module
V
is uniform if given non-zero
submodules V
1
, V
2
, then V
1
V
2
6= {0}.
Lemma. V is uniform iff E(V ) is indecomposable.
Proof.
Suppose
E
(
V
) =
A B
, with
A, B
non-zero. Then
V A 6
=
{
0
}
and
V B 6
=
{
0
}
since the extension is essential. So we have two non-zero submodules
of V that intersect trivially.
Conversely, suppose
V
is not uniform, and let
V
1
, V
2
be non-zero submod-
ules that intersect trivially. By Zorn’s lemma, we suppose these are maximal
submodules that intersect trivially. We claim
E(V
1
) E(V
2
) = E(V
1
V
2
) = E(V )
To prove this, it suffices to show that
V
is an essential extension of
V
1
V
2
, so
that E(V ) is an injective hull of V
1
V
2
.
Let
W V
be non-zero. If
W
(
V
1
V
2
) = 0, then
V
1
(
V
2
W
) is a larger
pair of submodules with trivial intersection, which is not possible. So we are
done.
Definition
(Domain)
.
An algebra is a domain if
xy
= 0 implies
x
= 0 or
y
= 0.
This is just the definition of an integral domain, but when we have non-
commutative algebras, we usually leave out the word “integral”.
To show that the algebras we know and love are indeed domains, we again
do some deformation.
Lemma.
Let
A
be a filtered algebra, which is exhaustive and separated. Then
if gr A is a domain, then so is A.
Proof.
Let
x A
i
\ A
i1
, and
y A
j
\ A
j1
. We can find such
i, j
for any
elements
x, y A
because the filtration is exhaustive and separated. Then we
have
¯x = x + A
i1
6= 0 A
i
/A
i1
¯y = y + A
j1
6= 0 A
j
/A
j1
.
If
gr A
is a domain, then we deduce
¯x¯y 6
= 0. So we deduce that
xy 6∈ A
i+j1
. In
particular, xy 6= 0.
Corollary. A
n
(k) and U(g) are domains.
Lemma.
Let
A
be a right Noetherian domain. Then
A
A
is uniform, i.e.
E
(
A
A
)
is indecomposable.
Proof.
Suppose not, and so there are
xA
and
yA
non-zero such that
xA yA
=
{0}. So xA yA is a direct sum.
But
A
is a domain and so
yA
=
A
as a right
A
-module. Thus
yxA yyA
is
a direct sum living inside yA. Further decomposing yyA, we find that
xA yxA y
2
xA · · · y
n
xA
is a direct sum of non-zero submodules. But this is an infinite strictly ascending
chain as n , which is a contradiction.
Recall that when we proved Artin–Wedderburn, we needed to use Krull–
Schmidt, which told us the decomposition is unique up to re-ordering. That
relied on the endomorphism algebra being local. We need something similar
here.
Lemma.
Let
E
be an indecomposable injective right module. Then
End
A
(
E
)
is a local algebra, with the unique maximal ideal given by
I = {f End(E) : ker f is essential}.
Note that since
E
is indecomposable injective, given any non-zero
V E
, we
know
E
(
V
) embeds into, and hence is a direct summand of
E
. Hence
E
(
V
) =
E
.
So
ker f
being essential is the same as saying
ker f
being non-zero. However,
this description of the ideal will be useful later on.
Proof.
Let
f
:
E E
and
ker f
=
{
0
}
. Then
f
(
E
) is an injective module, and
so is a direct summand of
E
. But
E
is indecomposable. So
f
is surjective. So it
is an isomorphism, and hence invertible. So it remains to show that
I = {f End(E) : ker f is essential}
is an ideal.
If
ker f
and
ker g
are essential, then
ker
(
f
+
g
)
ker f ker g
, and the
intersection of essential submodules is essential. So ker(f + g) is also essential.
Also, if
ker g
is essential, and
f
is arbitrary, then
ker
(
f g
)
ker g
, and is
hence also essential. So I is a maximal left ideal.
The point of this lemma is to allow us to use Krull–Schmidt.
Lemma.
Let
M
be a non-zero Noetherian module. Then
M
is an essential
extension of a direct sum of uniform submodules N
1
, · · · , N
r
. Thus
E(M)
=
E(N
1
) · · · E(N
r
)
is a direct sum of finitely many indecomposables.
This decomposition is unique up to re-ordering (and isomorphism).
Proof.
We first show any non-zero Noetherian module contains a uniform one.
Suppose not, and
M
is in particular not uniform. So it contains non-zero
V
1
, V
0
2
with
V
1
V
0
2
= 0. But
V
0
2
is not uniform by assumption. So it contains non-zero
V
2
and V
0
3
with zero intersection. We keep on repeating. Then we get
V
1
V
2
· · · V
n
is a strictly ascending chain of submodules of M , which is a contradiction.
Now for non-zero Noetherian
M
, pick
N
1
uniform in
M
. Either
N
1
is essential
in
M
, and we’re done, or there is some
N
0
2
non-zero with
N
1
N
0
2
= 0. We pick
N
2
uniform in N
0
2
. Then either N
1
N
2
is essential, or. . .
And we are done since M is Noetherian. Taking injective hulls, we get
E(M) = E(N
1
) · · · E(N
r
),
and we are done by Krull–Schmidt and the previous lemma.
This is the crucial lemma, which isn’t really hard. This allows us to define
yet another dimension for Noetherian algebras.
Definition
(Uniform dimension)
.
The uniform dimension, or Goldie rank of
M is the number of indecomposable direct summands of E(M).
This is analogous to vector space dimensions in some ways.
Example.
The Goldie rank of domains is 1, as we showed
A
A
is uniform. This
is true for A
n
(k) and U(g).
Lemma.
Let
E
1
, · · · , E
r
be indecomposable injectives. Put
E
=
E
1
· · · E
r
.
Let I = {f End
A
(E) : ker f is essential}. This is an ideal, and then
End
A
(E)/I
=
M
n
1
(D
1
) · · · M
n
s
(D
s
)
for some division algebras D
i
.
Proof. We write the decomposition instead as
E = E
n
1
1
· · · E
n
r
r
.
Then as in basic linear algebra, we know elements of
End
(
E
) can be written as
an r × r matrix whose (i, j)th entry is an element of Hom(E
n
i
i
, E
n
j
j
).
Now note that if
E
i
6
=
E
j
, then the kernel of a map
E
i
E
j
is essential in
E
i
. So quotienting out by I kills all of these “off-diagonal” entries.
Also
Hom
(
E
n
i
i
, E
n
i
i
) =
M
n
i
(
End
(
E
i
)), and so quotienting out by
I
gives
M
n
i
(End(E
i
)/{essential kernel})
=
M
n
i
(D
i
), where
D
i
=
End(E
i
)
essential kernel
,
which we know is a division algebra since I is a maximal ideal.
The final piece to proving Goldie’s theorem is the following piece
Lemma.
If
A
is a right Noetherian algebra, then any
f
:
A
A
A
A
with
ker f
essential in A
A
is nilpotent.
Proof. Consider
0 < ker f ker f
2
· · · .
Suppose
f
is not nilpotent. We claim that this is a strictly increasing chain.
Indeed, for all n, we have f
n
(A
A
) 6= 0. Since ker f is essential, we know
f
n
(A
A
) ker f 6= {0}.
This forces ker f
n+1
> ker f
n
, which is a contradiction.
We can now prove Goldie’s theorem.
Theorem
(Goldie’s theorem)
.
Let
A
be a right Noetherian algebra with no
non-zero ideals all of whose elements are nilpotent. Then
A
embeds in a finite
direct sum of matrix algebras over division algebras.
Proof. As usual, we have a map
A End
A
(A
A
)
x left multiplication by x
For a map A
A
A
A
, it lifts to a map E(A
A
) E(A
A
) by injectivity:
0 A
A
E(A
A
)
A
A
E(A
A
)
f
θ
f
0
θ
We can complete the diagram to give a map
f
0
:
E
(
A
A
)
E
(
A
A
), which
restricts to
f
on
A
A
. This is not necessarily unique. However, if we have two
lifts
f
0
and
f
00
, then the difference
f
0
f
00
has
A
A
in the kernel, and hence has
an essential kernel. So it lies in I. Thus, if we compose maps
A
A
End
A
(A
A
) End(E(A
A
))/I .
The kernel of this consists of
A
which when multiplying on the left has essential
kernel. This is an ideal all of whose elements is nilpotent. By assumption, any
such ideal vanishes. So we have an embedding of
A
in
End
(
E
(
A
A
))
/I
, which we
know to be a direct sum of matrix algebras over division rings.
Goldie didn’t present it like this. This work in injective modules is due to
Matlis.
We saw that (right Noetherian) domains had Goldie rank 1. So we get that
End
(
E
(
A
))
/I
=
D
for some division algebra
D
. So by Goldie’s theorem, a right
Noetherian algebra embeds in a division algebra. In particular, this is true for
A
n
(k) and U(g).
3 Hochschild homology and cohomology
3.1 Introduction
We now move on to talk about Hochschild (co)homology. We will mostly talk
about Hochschild cohomology, as that is the one that is interesting. Roughly
speaking, given a
k
-algebra
A
and an
A
-
A
-bimodule
M
, Hochschild cohomology is
an infinite sequence of
k
-vector spaces
HH
n
(
A, M
) indexed by
n N
associated
to the data. While there is in theory an infinite number of such vector spaces,
we are mostly going to focus on the cases of
n
= 0
,
1
,
2, and we will see that
these groups can be interpreted as things we are already familiar with.
The construction of these Hochschild cohomology groups might seem a bit
arbitrary. It is possible to justify these a priori using the general theory of
homological algebra and/or model categories. On the other hand, Hochschild
cohomology is sometimes used as motivation for the general theory of homological
algebra and/or model categories. Either way, we are not going to develop these
general frameworks, but are going to justify Hochschild cohomology in terms of
its practical utility.
Unsurprisingly, Hochschild (co)homology was first developed by Hochschild
in 1945, albeit only working with algebras of finite (vector space) dimension. It
was introduced to give a cohomological interpretation and generalization of some
results of Wedderburn. Later in 1962/1963, Gerstenhaber saw how Hochschild
cohomology was relevant to the deformations of algebras. More recently, it’s
been realized that that the Hochschild cochain complex has additional algebraic
structure, which allows yet more information about deformation.
As mentioned, we will work with
A
-
A
-bimodules over an algebra
A
. If our
algebra has an augmentation, i.e. a ring map to
k
, then we can have a decent
theory that works with left or right modules. However, for the sake of simplicity,
we shall just work with bimodules to make our lives easier.
Recall that a
A
-
A
-bimodule is an algebra with both left and right
A
actions
that are compatible. For example,
A
is an
A
-
A
-bimodule, and we sometimes write
it as
A
A
A
to emphasize this. More generally, we can view
A
(n+2)
=
A
k
· · ·
k
A
as an A-A-bimodule by
x(a
0
a
1
· · · a
n+1
)y = (xa
0
) a
1
· · · (a
n+1
y).
The crucial property of this is that for any
n
0, the bimodule
A
(n+2)
is a
free
A
-
A
-bimodule. For example,
A
k
A
is free on a single generator 1
k
1,
whereas if {x
i
} is a k-basis of A, then A
k
A
k
A is free on {1
k
x
i
k
1}.
The general theory of homological algebra says we should be interested in
such free things.
Definition
(Free resolution)
.
Let
A
be an algebra and
M
an
A
-
A
-bimodule. A
free resolution of M is an exact sequence of the form
· · · F
2
F
1
F
0
M
d
2
d
1
d
0
,
where each F
n
is a free A-A-bimodule.
More generally, we can consider a projective resolution instead, where we
allow the bimodules to be projective. In this course, we are only interested in
one particular free resolution:
Definition
(Hochschild chain complex)
.
Let
A
be a
k
-algebra with multiplication
map µ : A A. The Hochschild chain complex is
· · · A
k
A
k
A A
k
A A 0.
d
1
d
0
µ
We refer to
A
k
(n+2)
as the degree
n
term. The differential
d
:
A
k
(n+3)
A
k
(n+2)
is given by
d(a
0
k
· · ·
k
a
n+1
) =
n+1
X
i=0
(1)
i
a
0
k
· · ·
k
a
i
a
i+1
k
· · ·
k
a
n+2
.
This is a free resolution of
A
A
A
(the exactness is merely a computation, and
we shall leave that as an exercise to the reader). In a nutshell, given an
A
-
A
-
bimodule
M
, its Hochschild homology and cohomology is obtained by applying
·
A-A
M
and
Hom
A-A
(
· , M
) to the Hochschild chain complex, and then taking
the homology and cohomology of the resulting chain complex. We shall explore
in more detail what this means.
It is a general theorem that we could have applied the functors
·
A-A
M
and
Hom
A-A
(
· , M
) to any projective resolution of
A
A
A
and take the (co)homology,
and the resulting vector spaces will be the same. However, we will not prove
that, and will just always stick to this standard free resolution all the time.
3.2 Cohomology
As mentioned, the construction of Hochschild cohomology involves applying
Hom
A-A
(
· , M
) to the Hochschild chain complex, and looking at the terms
Hom
A-A
(
A
(n+2)
, M
). This is usually not very convenient to manipulate, as
it involves talking about bimodule homomorphisms. However, we note that
A
(n+2)
is a free
A
-
A
-bimodule generated by a basis of
A
n
. Thus, there is a
canonical isomorphism
Hom
A-A
(A
(n+2)
, M )
=
Hom
k
(A
n
, M ),
and k-linear maps are much simpler to work with.
Definition
(Hochschild cochain complex)
.
The Hochschild cochain complex
of an
A
-
A
-bimodule
M
is what we obtain by applying
Hom
A-A
(
· , M
) to the
Hochschild chain complex of A. Explicitly, we can write it as
Hom
k
(k, M) Hom
k
(A, M ) Hom
k
(A A, M ) · · · ,
δ
0
δ
1
where
(δ
0
f)(a) = af(1) f(1)a
(δ
1
f)(a
1
a
2
) = a
1
f(a
2
) f(a
1
a
2
) + f(a
1
)a
2
(δ
2
f)(a
1
a
2
a
3
) = a
1
f(a
2
a
3
) f(a
1
a
2
a
3
)
+ f(a
1
a
2
a
3
) f(a
1
a
2
)a
3
(δ
n1
f)(a
1
· · · a
n
) = a
1
f(a
2
· · · a
n
)
+
n
X
i=1
(1)
i
f(a
1
· · · a
i
a
i+1
· · · a
n
)
+ (1)
n+1
f(a
1
· · · a
n1
)a
n
The reason the end ones look different is that we replaced
Hom
A-A
(
A
(n+2)
, M
)
with Hom
k
(A
n
, M ).
The crucial observation is that the exactness of the Hochschild chain complex,
and in particular the fact that d
2
= 0, implies im δ
n1
ker δ
n
.
Definition (Cocycles). The cocycles are the elements in ker δ
n
.
Definition (Coboundaries). The coboundaries are the elements in im δ
n
.
These names came from algebraic topology.
Definition (Hochschild cohomology groups). We define
HH
0
(A, M ) = ker δ
0
HH
n
(A, N ) =
ker δ
n
im δ
n1
These are k-vector spaces.
If we do not want to single out
HH
0
, we can extend the Hochschild cochain
complex to the left with 0, and setting
δ
n
= 0 for
n <
0 (or equivalently extending
the Hochschild chain complex to the right similarly), Then
HH
0
(A, M ) =
ker δ
0
im δ
1
= ker δ
0
.
The first thing we should ask ourselves is when the cohomology groups van-
ish. There are two scenarios where we can immediately tell that the (higher)
cohomology groups vanish.
Lemma.
Let
M
be an injective bimodule. Then
HH
n
(
A, M
) = 0 for all
n
1.
Proof. Hom
A-A
( · , M ) is exact.
Lemma.
If
A
A
A
is a projective bimodule, then
HH
n
(
A, M
) = 0 for all
M
and
all n 1.
If we believed the previous remark that we could compute Hochschild coho-
mology with any projective resolution, then this result is immediate indeed,
we can use
· · ·
0
0
A A
0 as the projective resolution. However,
since we don’t want to prove such general results, we shall provide an explicit
computation.
Proof.
If
A
A
A
is projective, then all
A
n
are projective. At each degree
n
, we
can split up the Hochschild chain complex as the short exact sequence
0
A
(n+3)
ker d
n
A
(n+2)
im d
n1
0
d
n
d
n1
The im d is a submodule of A
(n+1)
, and is hence projective. So we have
A
(n+2)
=
A
(n+3)
ker d
n
im d
n1
,
and we can write the Hochschild chain complex at n as
ker d
n
A
(n+3)
ker d
n
A
(n+3)
ker d
n
im d
n1
A
(n+1)
im d
n1
im d
n1
(a, b) (b, 0)
(c, d) (0, d)
d
n
d
n1
Now
Hom
(
· , M
) certainly preserves the exactness of this, and so the Hochschild
cochain complex is also exact. So we have HH
n
(A, M ) = 0 for n 1.
This is a rather strong result. By knowing something about
A
itself, we
deduce that the Hochschild cohomology of any bimodule whatsoever must vanish.
Of course, it is not true that
HH
n
(
A, M
) vanishes in general for
n
1, or
else we would have a rather boring theory. In general, we define
Definition (Dimension). The dimension of an algebra A is
Dim A = sup{n : HH
n
(A, M ) 6= 0 for some A-A-bimodule M }.
This can be infinite if such a sup does not exist.
Thus, we showed that
A
A
A
embeds as a direct summand in
A A
, then
Dim A = 0.
Definition
(
k
-separable)
.
An algebra
A
is
k
-separable if
A
A
A
embeds as a
direct summand of A A.
Since A A is a free A-A-bimodule, this condition is equivalent to A being
projective. However, there is some point in writing the definition like this.
Note that an
A
-
A
-bimodule is equivalently a left
A A
op
-module. Then
A
A
A
is a direct summand of
A A
if and only if there is a separating idempotent
e A A
op
so that
A
A
A
viewed as A A
op
-bimodule is (A A
op
)e.
This is technically convenient, because it is often easier to write down a
separating idempotent than to prove directly A is projective.
Note that whether we write
AA
op
or
AA
is merely a matter of convention.
They have the same underlying set. The notation
A A
is more convenient
when we take higher powers, but we can think of
A A
op
as taking
A
as a
left-
A
right-
k
module and
A
op
as a left-
k
right-
A
, and tensoring them gives a
A-A-bimodule.
We just proved that separable algebras have dimension 0. Conversely, we
have
Lemma. If Dim A = 0, then A is separable.
Proof. Note that there is a short exact sequence
0 ker µ A A A 0
µ
If we can show this splits, then
A
is a direct summand of
A A
. To do so, we
need to find a map A A ker µ that restricts to the identity on ker µ.
To do so, we look at the first few terms of the Hochschild chain complex
· · · im d ker µ A A A 0
d
µ
.
By assumption, for any
M
, applying
Hom
A-A
(
· , M
) to the chain complex gives
an exact sequence. Omitting the annoying
A-A
subscript, this sequence looks like
0 Hom(A, M)
µ
Hom(A A, M)
()
Hom(ker µ, M) Hom(im d, M)
d
· · ·
Now
d
sends
Hom
(
ker µ, M
) to zero. So
Hom
(
ker µ, M
) must be in the image
of (). So the map
Hom(A A, M ) Hom(ker µ, M)
must be surjective. This is true for any
M
. In particular, we can pick
M
=
ker µ
.
Then the identity map
id
ker µ
lifts to a map
A A ker µ
whose restriction to
ker µ is the identity. So we are done.
Example. M
n
(
k
) is separable. It suffices to write down the separating idem-
potent. We let
E
ij
be the elementary matrix with 1 in the (
i, j
)th slot and 0
otherwise. We fix j, and then
X
i
E
ij
E
ji
A A
op
is a separating idempotent.
Example.
Let
A
=
kG
with
char k - |G|
. Then
A A
op
=
kG
(
kG
)
op
=
kG kG. But this is just isomorphic to k(G × G), which is again semi-simple.
Thus, as a bimodule,
A A
is completely reducible. So the quotient of
A
A
A
is a direct summand (of bimodules). So we know that whenever
char k - |G|
,
then kG is k-separable.
The obvious question is — is this notion actually a generalization of separable
field extensions? This is indeed the case.
Fact.
Let
L/K
be a finite field extension. Then
L
is separable as a
K
-algebra
iff L/K is a separable field extension.
However
k
-separable algebras must be finite-dimensional
k
-vector spaces. So
this doesn’t pass on to the infinite case.
In the remaining of the chapter, we will study what Hochschild cohomology
in the low dimensions mean. We start with
HH
0
. The next two propositions
follow from just unwrapping the definitions:
Proposition. We have
HH
0
(A, M ) = {m M : am ma = 0 for all a A}.
In particular, HH
0
(A, A) is the center of A.
Proposition.
ker δ
1
= {f Hom
k
(A, M ) : f (a
1
a
2
) = a
1
f(a
2
) + f(a
1
)a
2
}.
These are the derivations from A to M . We write this as Der(A, M).
On the other hand,
im δ
0
= {f Hom
k
(A, M ) : f (a) = am ma for some m M}.
These are called the inner derivations from A to M. So
HH
1
(A, M ) =
Der(A, M )
InnDer(A, M )
.
Setting A = M , we get the derivations and inner derivations of A.
Example. If A = k[x], and char k = 0, then
DerA =
p(X)
d
dx
: p(x) k[X]
,
and there are no (non-trivial) inner derivations because
A
is commutative. So
we find
HH
1
(k[X], k[X])
=
k[X].
In general, Der(A) form a Lie algebra, since
D
1
D
2
D
2
D
1
End
k
(A)
is in fact a derivation if D
1
and D
2
are.
There is another way we can think about derivations, which is via semi-direct
products.
Definition
(Semi-direct product)
.
Let
M
be an
A
-
A
-bimodule. We can form
the semi-direct product of
A
and
M
, written
A n M
, which is an algebra with
elements (a, m) A × M , and multiplication given by
(a
1
, m
1
) · (a
2
, m
2
) = (a
1
a
2
, a
1
m
2
+ m
1
a
2
).
Addition is given by the obvious one.
Alternatively, we can write
A n M
=
A + Mε,
where
ε
commutes with everything and
ε
2
= 0. Then
Mε
forms an ideal with
(Mε)
2
= 0.
In particular, we can look at
A n A
=
A
+
. This is often written as
A
[
ε
].
Previously, we saw first cohomology can be understood in terms of derivations.
We can formulate derivations in terms of this semi-direct product.
Lemma. We have
Der
k
(A, M )
=
{algebra complements to M in A n M isomorphic to A}.
Proof. A complement to M is an embedded copy of A in A n M,
A A n M
a (a, D
a
)
The function
A M
given by
a 7→ D
a
is a derivation, since under the embedding,
we have
ab 7→ (ab, aD
b
+ D
a
b).
Conversely, a derivation
f
:
A M
gives an embedding of
A
in
A n M
given by
a 7→ (a, f(a)).
We can further rewrite this characterization in terms of automorphisms of
the semi-direct product. This allows us to identify inner derivations as well.
Lemma. We have
Der(A, M )
=
automorphisms of A n M of the form
a 7→ a + f (a)ε, 7→
,
where we view A n M
=
A + Mε.
Moreover, the inner derivations correspond to automorphisms achieved by
conjugation by 1 + , which is a unit with inverse 1 .
The proof is a direct check.
This applies in particular when we pick
M
=
A
A
A
, and the Lie algebra of
derivations of A may be thought of as the set of infinitesimal automorphisms.
Let’s now consider
HH
2
(
A, M
). This is to be understood in terms of exten-
sions, of which the “trivial” example is the semi-direct product.
Definition
(Extension)
.
Let
A
be an algebra and
M
and
A
-
A
-bimodule. An
extension of
A
by
M
. An extension of
A
by
M
is a
k
-algebra
B
containing a
2-sided ideal I such that
I
2
= 0;
B/I
=
A; and
M
=
I as an A-A-bimodule.
Note that since
I
2
= 0, the left and right multiplication in
B
induces an
A-A-bimodule structure on I, rather than just a B-B-bimodule.
We let
π
:
B A
be the canonical quotient map. Then we have a short
exact sequence
0 I B A 0 .
Then two extensions
B
1
and
B
2
are isomorphic if there is a
k
-algebra isomorphism
θ : B
1
B
2
such that the following diagram commutes:
B
1
0 I A 0
B
2
θ
.
Note that the semi-direct product is such an extension, called the split extension.
Proposition.
There is a bijection between
HH
2
(
A, M
) with the isomorphism
classes of extensions of A by M.
This is something that should be familiar if we know about group cohomology.
Proof.
Let
B
be an extension with, as usual,
π
:
B A
,
I
=
M
=
ker π
,
I
2
= 0.
We now try to produce a cocycle from this.
Let
ρ
be any
k
-linear map
A B
such that
π
(
ρ
(
a
)) =
a
. This is possible
since
π
is surjective. Equivalently,
ρ
(
π
(
b
)) =
b mod I
. We define a
k
-linear map
f
ρ
: A A I
=
M
by
a
1
a
2
7→ ρ(a
1
)ρ(a
2
) ρ(a
1
a
2
).
Note that the image lies in I since
ρ(a
1
)ρ(a
2
) ρ(a
1
a
2
) (mod I).
It is a routine check that f
ρ
is a 2-cocycle, i.e. it lies in ker δ
2
.
If we replace ρ by any other ρ
0
, we get f
ρ
0
, and we have
f
ρ
(a
1
a
2
) f
ρ
0
(a
1
a
2
)
= ρ(a
1
)(ρ(a
2
) ρ
0
(a
2
)) (ρ(a
1
a
2
) ρ
0
(a
1
a
2
)) + (ρ(a
1
) ρ
0
(a
1
))ρ
0
(a
2
)
= a
1
· (ρ(a
2
) ρ
0
(a
2
)) (ρ(a
1
a
2
) ρ
0
(a
1
a
2
)) + (ρ(a
1
) ρ
0
(a
1
)) · a
2
,
where · denotes the A-A-bimodule action in I. Thus, we find
f
ρ
f
ρ
0
= δ
1
(ρ ρ
0
),
noting that ρ ρ
0
actually maps to I.
So we obtain a map from the isomorphism classes of extensions to the second
cohomology group.
Conversely, given an
A
-
A
-bimodule
M
and a 2-cocycle
f
:
A A M
, we
let
B
f
= A M
as k-vector spaces. We define the multiplication map
(a
1
, m
1
)(a
2
, m
2
) = (a
1
a
2
, a
1
m
2
+ m
1
a
2
+ f(a
1
a
2
)).
This is associative precisely because of the 2-cocycle condition. The map (
a, m
)
a
yields a homomorphism
π
:
B A
, with kernel
I
being a two-sided ideal of
B
which has
I
2
= 0. Moreover,
I
=
M
as an
A
-
A
-bimodule. Taking
ρ
:
A B
by
ρ(a) = (a, 0) yields the 2-cocycle we started with.
Finally, let
f
0
be another 2 co-cycle cohomologous to
f
. Then there is a
linear map τ : A M with
f f
0
= δ
1
τ.
That is,
f(a
1
A
2
) = f
0
(a
1
a
2
) + a
1
· τ(a
2
) τ(a
1
a
2
) + τ(a
1
) · a
2
.
Then consider the map B
f
B
0
f
given by
(a, m) 7→ (a, m + τ(a)).
One then checks this is an isomorphism of extensions. And then we are done.
In the proof, we see 0 corresponds to the semi-direct product.
Corollary. If HH
2
(A, M ) = 0, then all extensions are split.
We now prove some theorems that appear to be trivialities. However, they
are trivial only because we now have the machinery of Hochschild cohomology.
When they were first proved, such machinery did not exist, and they were written
in a form that seemed less trivial.
Theorem (Wedderburn, Malcev). Let B be a k-algebra satisfying
Dim(B/J(B)) 1.
J(B)
2
= 0
Then there is an subalgebra A
=
B/J(B) of B such that
B = A n J(B).
Furthermore, if
Dim
(
B/J
(
B
)) = 0, then any two such subalgebras
A, A
0
are
conjugate, i.e. there is some x J(B) such that
A
0
= (1 + x)A(1 + x)
1
.
Notice that 1 + x is a unit in B.
In fact, the same result holds if we only require
J
(
B
) to be nilpotent. This
follows from an induction argument using this as a base case, which is messy
and not really interesting.
Proof. We have J(B)
2
= 0. Since we know Dim(B/J(B)) 1, we must have
HH
2
(A, J(B)) = 0
where
A
=
B
J(B)
.
Note that we regard
J
(
B
) as an
A
-
A
-bimodule here. So we know that all
extension of A by J(B) are semi-direct, as required.
Furthermore, if
Dim
(
B/J
(
B
)) = 0, then we know
HH
1
(
A, J
(
A
)) = 0. So by
our older lemmas, we see that complements are all conjugate, as required.
Corollary.
If
k
is algebraically closed and
dim
k
B <
, then there is a subal-
gebra A of B such that
A
=
B
J(B)
,
and
B = A n J(B).
Moreover,
A
is unique up to conjugation by units of the form 1+
x
with
x J
(
B
).
Proof.
We need to show that
Dim
(
A
) = 0. But we know
B/J
(
B
) is a semi-
simple
k
-algebra of finite dimension, and in particular is Artinian. So by
Artin–Wedderburn, we know
B/J
(
B
) is a direct sum of matrix algebras over
k
(since k is algebraically closed and dim
k
(B/J(B))).
We have previously observed that
M
n
(
k
) is
k
-separable. Since
k
-separability
behaves well under direct sums, we know
B/J
(
B
) is
k
-separable, hence has
dimension zero.
It is a general fact that J(B) is nilpotent.
3.3 Star products
We are now going to do study some deformation theory. Suppose
A
is a
k
-algebra.
We write
V
for the underlying vector space of
A
. Then there is a natural algebra
structure on
V
k
k
[[
t
]] =
V
[[
t
]], which we may write as
A
[[
t
]]. However, we
might we to consider more interesting algebra structures on this vector space.
Of course, we don’t want to completely forget the algebra structure on
A
. So we
make the following definition:
Definition
(Star product)
.
Let
A
be a
k
-algebra, and let
V
be the underlying
vector space. A star product is an associative
k
[[
t
]]-bilinear product on
V
[[
t
]]
that reduces to the multiplication on A when we set t = 0.
Can we produce non-trivial star products? It seems difficult, because when
we write down an attempt, we need to make sure it is in fact associative, and
that might take quite a bit of work. One example we have already seen is the
following:
Example.
Given a filtered
k
-algebra
A
0
, we formed the Rees algebra associated
with the filtration, and it embeds as a vector space in (
gr A
0
)[[
t
]]. Thus we get a
product on (gr A
0
)[[t]].
There are two cases where we are most interested in when
A
0
=
A
n
(
k
) or
A
0
=
U
(
g
). We saw that
gr A
0
was actually a (commutative) polynomial algebra.
However, the product on the Rees algebra is non-commutative. So the
-product
will be non-commutative.
In general, the availability of star products is largely controlled by the
Hochschild cohomology of
A
. To understand this, let’s see what we actually
need to specify to get a star product. Since we required the product to be a
k[[t]]-bilinear map
f : V [[t]] × V [[t]] V [[t]],
all we need to do is to specify what elements of
V
=
A
are sent to. Let
a, b V = A. We write
f(a, b) = ab + tF
1
(a, b) + t
2
F
2
(a, b) + · · · .
Because of bilinearity, we know
F
i
are
k
-bilinear maps, and so correspond to
k-linear maps V V V . For convenience, we will write
F
0
(a, b) = ab.
The only non-trivial requirement f has to satisfy is associativity:
f(f(a, b), c) = f(a, f (b, c)).
What condition does this force on our
F
i
? By looking at coefficients of
t
, this
implies that for all λ = 0, 1, 2, · · · , we have
X
m+n=λ
m,n0
F
m
(F
n
(a, b), c) F
m
(a, F
n
(b, c))
= 0. ()
For
λ
= 0, we are just getting the associativity of the original multiplication on
A. When λ = 1, then this says
aF
1
(b, c) F
1
(ab, c) + F
1
(a, bc) F
1
(a, b)c = 0.
All this says is that
F
1
is a 2-cocycle! This is not surprising. Indeed, we’ve
just seen (a while ago) that working mod
t
2
, the extensions of
A
by
A
A
A
are
governed by
HH
2
. Thus, we will refer to 2-cocycles as infinitesimal deformations
in this context.
Note that given an arbitrary 2 co-cycle
A A A
, it may not be possible
to produce a star product with the given 2-cocycle as F
1
.
Definition
(Integrable 2-cocycle)
.
Let
f
:
A A A
be a 2-cocycle. Then it
is integrable if it is the F
1
of a star product.
We would like to know when a 2-cocycle is integrable. Let’s rewrite (
) as
(
λ
):
X
m+n=λ
m,n>0
F
m
(F
n
(a, b), c) F
m
(a, F
n
(b, c))
= (δ
2
F
λ
)(a, b, c). (
λ
)
Here we are identifying F
λ
with the corresponding k-linear map A A A.
For λ = 2, this says
F
1
(F
1
(ab), c) F
1
(a, F
1
(b, c)) = (δ
2
F
2
)(a, b, c).
If
F
1
is a 2-cocycle, then one can check that the LHS gives a 3-cocycle. If
F
1
is integrable, then the LHS has to be equal to the RHS, and so must be a
coboundary, and thus has cohomology class zero in HH
3
(A, A).
In fact, if
F
1
, · · · , F
λ1
satisfy (
1
)
, · · · ,
(
λ1
), then the LHS of (
λ
) is also
a 3-cocycle. If it is a coboundary, and we have defined
F
1
, · · · , F
λ
1
, then we can
define
F
λ
such that (
λ
) holds. However, if it is not a coboundary, then we get
stuck, and we see that our choice of
F
1
, · · · , F
λ1
does not lead to a
-product.
The 3-cocycle appearing on the LHS of (
λ
) is an obstruction to integrability.
If, however, they are always coboundaries, then we can inductively define
F
1
, F
2
, · · · to give a -product. Thus we have proved
Theorem
(Gerstenhaber)
.
If
HH
3
(
A, A
) = 0, then all infinitesimal deformations
are integrable.
Of course, even if
HH
3
(
A, A
)
6
= 0, we can still get
-products, but we need
to pick our F
1
more carefully.
Now after producing star products, we want to know if they are equivalent.
Definition
(Equivalence of star proeducts)
.
Two star products
f
and
g
are
equivalent on
V k
[[
t
]] if there is a
k
[[
t
]]-linear automorphism Φ of
V
[[
t
]] of the
form
Φ(a) = a +
1
(a) + t
2
φ
2
(a) + · · ·
sch that
f(a, b) = Φ
1
g(Φ(a), Φ(b)).
Equivalently, the following diagram has to commute:
V [[t]] V [[t]] V [[t]]
V [[t]] V [[t]] V [[t]]
f
ΦΦ
Φ
g
Star products equivalent to the usual product on A k[[t]] are called trivial.
Theorem
(Gerstenhaber)
.
Any non-trivial star product
f
is equivalent to one
of the form
g(a, b) = ab + t
n
G
n
(a, b) + t
n+1
G
n+1
(a, b) + · · · ,
where
G
n
is a 2-cocycle and not a coboundary. In particular, if
HH
2
(
A, A
) = 0,
then any star product is trivial.
Proof. Suppose as usual
f(a, b) = ab + tF
1
(a, b) + t
2
F
2
(a, b) + · · · ,
and suppose F
1
, · · · , F
n1
= 0. Then it follows from () that
δ
2
F
n
= 0.
If F
n
is a coboundary, then we can write
F
n
= δφ
n
for some φ
n
: A A. We set
Φ
n
(a) = a + t
n
φ
n
(a).
Then we can compute that
Φ
1
n
(f
n
(a), Φ
n
(b)))
is of the form
ab + t
n+1
G
n+1
(a, b) + · · · .
So we have managed to get rid of a further term, and we can keep going until
we get the first non-zero term not a coboundary.
Suppose this never stops. Then
f
is trivial we are using that
· · ·
Φ
n+2
Φ
n+1
Φ
n
converges in the automorphism ring, since we are adding terms of
higher and higher degree.
We saw that derivations can be thought of as infinitesimal automorphisms.
One can similarly consider k[[t]]-linear maps of the form
Φ(a) = a +
1
(a) + t
2
φ
2
(a) + · · ·
and consider whether they define automorphisms of
A k
[[
t
]]. Working modulo
t
2
, we have already done this problem we are just considering automorphisms
of A[ε], and we saw that these automorphisms correspond to derivations.
Definition
(Integrable derivation)
.
We say a derivation is integrable if there is
an automorphism of A k[[t]] that gives the derivation when we mod t
2
.
In this case, the obstructions are 2-cocycles which are not coboundaries.
Theorem
(Gerstenhaber)
.
Suppose
HH
2
(
A, A
) = 0. Then all derivations are
integrable.
The proof is an exercise on the third example sheet.
We haven’t had many examples so far, because Hochschild cohomology is
difficult to compute. But we can indeed do some examples.
Example. Let A = k[x]. Since A is commutative, we have
HH
0
(A, A) = A.
Since A is commutative, A has no inner derivations. So we have
HH
1
(A, A) = DerA =
f(x)
d
dx
: f(x) k[x]
.
For any i > 1, we have
HH
i
(A, A) = 0.
So we have
Dim(A) = 1.
We can do this by explicit calculation. If we look at our Hochschild chain
complex, we had a short exact sequence
0 ker µ A A A 0 ()
and thus we have a map
A A A A A
d
whose image is ker µ .
The point is that
ker µ
is a projective
A
-
A
-bimodule. This will mean that
HH
i
(
A, M
) = 0 for
i
2 in the same way we used to show that when
A
A
A
is a
projective A-A-bimodule for i 1. In particular, HH
i
(A, A) = 0 for i 2.
To show that ker µ is projective, we notice that
A A = k[X]
k
k[X]
=
k[X, Y ].
So the short exact sequence () becomes
0 (X Y )k[X, Y ] k[X, Y ] k[X] 0 .
So (X Y )k[X, Y ] is a free k[X, Y ] module, and hence projective.
We can therefore use our theorems to see that any extension of
k
[
X
] by
k
[
X
] is split, and any
-product is trivial. We also get that any derivation is
integrable.
Example. If we take A = k[X
1
, X
2
], then again this is commutative, and
HH
0
(A, A) = A
HH
1
(A, A) = DerA.
We will talk about HH
2
later, and similarly
HH
i
(A, A) = 0
for i 3.
From this, we see that we may have star products other than the trivial ones,
and in fact we know we have, because we have one arising from the Rees algebra
of
A
1
(
k
). But we know that any infinitesimal deformation yields a star product.
So there are much more.
3.4 Gerstenhaber algebra
We now want to understand the equations (
) better. To do so, we consider the
graded vector space
HH
·
(A, A) =
M
A=0
HH
n
(A, A),
as a whole. It turns out this has the structure of a Gerstenhaber algebra
The first structure to introduce is the cup product. They are a standard tool
in cohomology theories. We will write
S
n
(A, A) = Hom
k
(A
n
, A) = Hom
A-A
(A
(n+2)
,
A
A
A
).
The Hochschild chain complex is then the graded chain complex S
·
(A, A).
Definition (Cup product). The cup product
: S
m
(A, A) S
n
(A, A) S
m+n
(A, A)
is defined by
(f g)(a
1
· · · a
m
b
1
· · · b
n
) = f(a
1
· · · a
m
) · g(b
1
· · · b
n
),
where a
i
, b
j
A.
Under this product, S
·
(A, A) becomes an associative graded algebra.
Observe that
δ(f g) = δf g + (1)
mn
f δg.
So we say
δ
is a (left) graded derivation of the graded algebra
S
·
(
A, A
). In
homological (graded) algebra, we often use the same terminology but with
suitable sign changes which depends on the degree.
Note that the cocycles are closed under
. So cup product induces a product
on
HH
·
(
A, A
). If
f S
m
(
A, A
) and
g S
n
(
A, A
), and both are cocycles, then
(1)
m
(g f (1)
mn
(f g)) = δ(f g),
where f g is defined as follows: we set
f
i
g(a
1
· · · a
i1
b
1
· · · b
n
a
i+1
· · · a
m
)
= f(a
1
· · · a
i1
g(b
1
· · · b
n
) a
i+1
· · · a
m
).
Then we define
f g =
m
X
i=1
(1)
(n1)(i1)
f
i
g.
This product
is not an associative product, but is giving a pre-Lie structure to
S
·
(A, A).
Definition (Gerstenhaber bracket). The Gerstenhaber bracket is
[f, g] = f g (1)
(n+1)(m+1)
g f
This defines a graded Lie algebra structure on the Hochschild chain complex,
but notice that we have a degree shift by 1. It is a grade Lie algebra on
S
·
+1
(A, A).
Of course, we really should define what a graded Lie algebra is.
Definition (Graded Lie algebra). A graded Lie algebra is a vector space
L =
M
L
i
with a bilinear bracket [ · , · ] : L × L L such that
[L
i
, L
j
] L
i+j
;
[f, g] (1)
mn
[g, f]; and
The graded Jacobi identity holds:
(1)
mp
[[f, g], h] + (1)
mn
[[g, h], f] + (1)
np
[[h, f ], g] = 0
where f L
m
, g L
n
, h L
p
.
In fact,
S
·
+1
(
A, A
) is a differential graded Lie algebra under the Gerstenhaber
bracket.
Lemma. The cup product on HH
·
(A, A) is graded commutative, i.e.
f g = (1)
mn
(g f).
when f HH
m
(A, A) and g HH
n
(A, A).
Proof. We previously “noticed” that
(1)
m
(g f (1)
mn
(f g)) = δ(f g),
Definition
(Gerstenhaber algebra)
.
A Gerstenhaber algebra is a graded vector
space
H =
M
H
i
with
H
·
+1
a graded Lie algebra with respect to a bracket [
· , ·
] :
H
m
× H
n
H
m+n1
, and an associative product
:
H
m
× H
n
H
m+n
which is graded
commutative, such that if
f H
m
, then [
f, ·
] acts as a degree
m
1 graded
derivation of :
[f, g h] = [f, g] h + (1)
(m1)
ng [f, h]
if g H
n
.
This is analogous to the definition of a Poisson algebra. We’ve seen that
HH
·
(A, A) is an example of a Gerstenhaber algebra.
We can look at what happens in low degrees. We know that
H
0
is a
commutative k-algebra, and : H
0
× H
1
H
1
is a module action.
Also,
H
1
is a Lie algebra, and [
· , ·
] :
H
1
× H
0
H
0
is a Lie module
action, i.e.
H
0
gives us a Lie algebra representation of
H
1
. In other words,
the corresponding map [
· , ·
] :
H
1
End
k
(
H
0
) gives us a map of Lie algebras
H
1
Der(H
0
).
The prototype Gerstenhaber algebra is the exterior algebra
V
DerA
for a
commutative algebra A (with A in degree 0).
Explicitly, to define the exterior product over
A
, we first consider the tensor
product over A of two A-modules V and W , defined by
V
A
W =
V
k
W
hav w v awi
The exterior product is then
V
A
V =
V
A
V
hv v : v V i
.
The product is given by the wedge, and the Schouten bracket is given by
[λ
1
· · · λ
m
, λ
0
1
· · · λ
0
n
]
= (1)
(m1)(n1)
X
i,j
(1)
i+j
[λ
i
, λ
j
] λ
1
· · · λ
m
| {z }
ith missing
λ
0
1
· · · λ
0
n
| {z }
jth missing
.
For any Gerstenhaber algebra
H
=
L
H
i
, there is a canonical homomorphism
of Gerstenhaber algebras
^
H
0
H
1
H.
Theorem
(Hochschild–Kostant–Ronsenberg (HKR) theorem)
.
If
A
is a “smooth”
commutative k-algebra, and char k = 0, then the canonical map
^
A
(DerA) HH
(A, A)
is an isomorphism of Gerstenhaber algebras.
We will not say what “smooth” means, but this applies if
A
=
k
[
X
1
, · · · , X
n
],
or if
k
=
C
or
R
and
A
is appropriate functions on smooth manifolds or algebraic
varieties.
In the 1960’s, this was stated just for the algebra structure, and didn’t think
about the Lie algebra.
Example. Let A = k[X, Y ], with char k = 0. Then HH
0
(A, A) = A and
HH
1
(A, A) = DerA
=
p(X, Y )
y
+ q(X, Y )
Y
: p, q A
.
So we have
HH
2
(A, A) = DerA
A
DerA,
which is generated as an A-modules by
X
Y
. Then
HH
i
(A, A) = 0 for all i 3
We can now go back to talk about star products. Recall when we considered
possible star products on
V
k
k
[[
t
]], where
V
is the underlying vector space of
the algebra
A
. We found that associativity of the star product was encapsulated
by some equations (
λ
). Collectively, these are equivalent to the statement
Definition (Maurer–Cartan equation). The Maurer–Cartan equation is
δf +
1
2
[f, f]
Gerst
= 0
for the element
f =
X
t
λ
F
λ
,
where F
0
(a, b) = ab.
When we write [
· , ·
]
Gerst
, we really mean the
k
[[
t
]]-linear extension of the
Gerstenhaber bracket.
If we want to think of things in cohomology instead, then we are looking
at things modulo coboundaries. For the graded Lie algebra
V
·
+1
(
DerA
), the
Maurer–Cartan elements, i.e. solutions of the Maurer–Cartan equation, are the
formal Poisson structures. They are formal power series of the form
Π =
X
t
i
π
i
for π
i
DerA DerA, satisfying
, Π] = 0.
There is a deep theorem of Kontsevich from the early 2000’s which implies
Theorem (Kontsevich). There is a bijection
equivalence classes
of star products
n
classes of formal
Poisson structures
o
This applies for smooth algebras in
char
0, and in particular for polynomial
algebras A = k[X
1
, · · · , X
n
].
This is a difficult theorem, and the first proof appeared in 2002.
An unnamed lecturer once tried to give a Part III course with this theorem as
the punchline, but the course ended up lasting 2 terms, and they never reached
the punchline.
3.5 Hochschild homology
We don’t really have much to say about Hochschild homology, but we are morally
obliged to at least write down the definition.
To do Hochschild homology, we apply
·
A-A
M
for an
A
-
A
-bimodule
M
to
the Hochschild chain complex.
· · · A
k
A
k
A A
k
A A 0.
d d
µ
,
We will ignore the
A
0 bit. We need to consider what
·
A-A
·
means. If
we have bimodules
V
and
W
, we can regard
V
as a right
A A
op
-module. We
can also think of W as a left A A
op
module. We let
B = A A
op
,
and then we just consider
V
B
W =
V
k
W
hvx w v xw : w Bi
=
V
k
W
hava
0
w v a
0
wai
.
Thus we have
· · · (A
k
A
k
A)
A-A
M (A
k
A)
A-A
M
=
M
b
1
b
0
,
Definition (Hochschild homology). The Hochschild homology groups are
HH
0
(A, M ) =
M
im b
0
HH
i
(A, M ) =
ker b
i1
im b
i
for i > 0.
A long time ago, we counted the number of simple
kG
-modules for
k
alge-
braically closed of characteristic
p
when
G
is finite. In the proof, we used
A
[A,A]
,
and we pointed out this is HH
0
(A, A).
Lemma.
HH
0
(A, M ) =
M
hxm mx : m M, x Ai
.
In particular,
HH
0
(A, A) =
A
[A, A]
.
Proof. Exercise.
4 Coalgebras, bialgebras and Hopf algebras
We are almost at the end of the course. So let’s define an algebra.
Definition (Algebra). A k-algebra is a k-vector space A and k-linear maps
µ : A A A u : k A
x y 7→ xy λ 7→ λI
called the multiplication/product and unit such that the following two diagrams
commute:
A A A A A
A A A
µid
id µ
µ
µ
k A A A A k
A
=
uid
µ
id u
=
These encode associativity and identity respectively.
Of course, the point wasn’t to actually define an algebra. The point is to
define a coalgebra, whose definition is entirely dual.
Definition (Coalgebra). A coalgebra is a k-vector space C and k-linear maps
∆ : C C C ε : C k
called comultiplication/coproduct and counit respectively, such that the following
diagrams commute:
C C C C C
C C C
id
id
k C C C C k
C
εid id ε
µ
=
=
These encode coassociativity and coidentity
A morphism of coalgebras
f
:
C D
is a
k
-linear map such that the following
diagrams commute:
C D
C C D D
f
ff
C D
k k
ε
f
ε
A subspace
I
of
C
is a co-ideal if ∆(
I
)
C I
+
I C
, and
ε
(
I
) = 0. In this
case, C/I inherits a coproduct and counit.
A cocommutative coalgebra is one for which
τ
∆ = ∆, where
τ
:
V W
W V given by the v w 7→ w v is the “twist map”.
It might be slightly difficult to get one’s head around what a coalgebra actually
is. It, of course, helps to look at some examples, and we will shortly do so. It
also helps to know that for our purposes, we don’t really care about coalgebras
per se, but things that are both algebras and coalgebras, in a compatible way.
There is a very natural reason to be interested in such things. Recall that
when doing representation theory of groups, we can take the tensor product of
two representations and get a new representation. Similarly, we can take the
dual of a representation and get a new representation.
If we try to do this for representations (ie. modules) of general algebras, we
see that this is not possible. What is missing is that in fact, the algebras
kG
and
U
(
g
) also have the structure of coalgebras. In fact, they are Hopf algebras,
which we will define soon.
We shall now write down some coalgebra structures on kG and U(g).
Example. If G is a group, then kG is a co-algebra, with
∆(g) = g g
ε
λ
g
X
(g)
=
X
λ
g
.
We should think of the specification ∆(
g
) =
g g
as saying that our groups act
diagonally on the tensor products of representations. More precisely, if
V, W
are
representations and v V, w W , then g acts on v w by
∆(g) · (v w) = (g g) · (v w) = (gv) (gw).
Example.
For a Lie algebra
g
over
k
, the universal enveloping algebra
U
(
g
) is
a co-algebra with
∆(x) = x 1 + 1 x
for x g, and we extend this by making it an algebra homomorphism.
To define ε, we note that elements of U(g) are uniquely of the form
λ +
X
λ
i
1
,...,i
n
x
i
1
1
· · · x
i
n
n
,
where {x
i
} is a basis of g (the PBW theorem). Then we define
ε
λ +
X
λ
i
1
,...,i
n
x
i
1
1
· · · x
i
n
n
= λ.
This time, the specification of is telling us that if
X g
and
v, w
are elements
of a representation of g, then X acts on the tensor product by
∆(X) · (v w) = Xv w + v Xw.
Example. Consider
O(M
n
(k)) = k[X
ij
: 1 i, j n],
the polynomial functions on
n × n
matrices, where
X
ij
denotes the
ij
th entry.
Then we define
∆(X
ij
) =
n
X
i=1
X
i`
X
`j
,
and
ε(X
ij
) = δ
ij
.
These are again algebra maps.
We can also talk about
O
(
GL
n
(
k
)) and
O
(
SL
n
(
k
)). The formula of the
determinant gives an element
D O
(
M
n
(
k
)). Then
O
(
GL
n
(
k
)) is given by
adding a formal inverse to
D
in
O
(
GL
n
(
k
)), and
O
(
SL
n
(
k
)) is obtained by
quotienting out O(GL
n
(k)) by the bi-ideal hD 1i.
From an algebraic geometry point of view, these are the coordinate algebra
of the varieties M
n
(k), GL
n
(k) and SL
n
(k).
This is dual to matrix multiplication.
We have seen that we like things that are both algebras and coalgebras,
compatibly. These are known as bialgebras.
Definition
(Bialgebra)
.
A bialgebra is a
k
-vector space
B
and maps
µ, υ,
, ε
such that
(i) (B, µ, u) is an algebra.
(ii) (B, , ε) is a coalgebra.
(iii) and ε are algebra morphisms.
(iv) µ and u are coalgebra morphisms.
Being a bialgebra means we can take tensor products of modules and still
get modules. If we want to take duals as well, then it turns out the right notion
is that of a Hopf algebra:
Definition
(Hopf algebra)
.
A bialgebra (
H, µ, u,
, ε
) is a Hopf algebra if there
is an antipode S : H H that is a k-linear map such that
µ (S id) ∆ = µ (id S) ∆ = u ε.
Example. kG is a Hopf algebra with S(g) = g
1
.
Example. U(g) is a Hopf algebra with S(x) = x for x U(g).
Note that our examples are all commutative or co-commutative. The term
quantum groups usually refers to a non-commutative non-co-commutative Hopf
algebras. These are neither quantum nor groups.
As usual, we write
V
for
Hom
k
(
V, k
), and we note that if we have
α
:
V W
,
then this induces a dual map α
: W
V
.
Lemma.
If
C
is a coalgebra, then
C
is an algebra with multiplication ∆
(that
is,
|
C
C
) and unit ε
. If C is co-commutative, then C
is commutative.
However, if an algebra
A
is infinite dimensional as a
k
-vector space, then
A
may not be a coalgebra. The problem is that (
A
A
) is a proper subspace of
(
A A
)
, and
µ
of an infinite dimensional
A
need not take values in
A
A
.
However, all is fine for finite dimensional
A
, or if
A
is graded with finite
dimensional components, where we can form a graded dual.
In general, for a Hopf algebra H, one can define the Hopf dual,
H
0
= {f H
: ker f contains an ideal of finite codimension}.
Example.
Let
G
be a finite group. Then (
kG
)
is a commutative non-co-
commutative Hopf algebra if G is non-abelian.
Let
{g}
be the canonical basis for
kG
, and
{φ
g
}
be the dual basis of (
kG
)
.
Then
∆(φ
g
) =
X
h
1
h
2
=g
φ
h
1
φ
h
2
.
There is an easy way of producing non-commutative non-co-commutative Hopf
algebras we take a non-commutative Hopf algebra and a non-co-commutative
Hopf algebra, and take the tensor product of them, but this is silly.
The easiest non-trivial example of a non-commutative non-co-commutative
Hopf algebra is the Drinfeld double, or quantum double, which is a general
construction from a finite dimensional hopf algebra.
Definition (Drinfeld double). Let G be a finite group. We define
D(G) = (kG)
k
kG
as a vector space, and the algebra structure is given by the crossed product
(kG)
o G, where G acts on (kG)
by
f
g
(x) = f(gxg
1
).
Then the product is given by
(f
1
g
1
)(f
2
g
2
) = f
1
f
g
1
1
2
g
1
g
2
.
The coalgebra structure is the tensor of the two coalgebras (
kG
)
and
kG
, with
∆(φ
g
h) =
X
g
1
g
2
=g
φ
g
1
h φ
g
2
h.
D
(
G
) is quasitriangular, i.e. there is an invertible element
R
of
D
(
G
)
D
(
G
)
such that
R∆(x)R
1
= τ(∆(x)),
where τ is the twist map. This is given by
R =
X
g
(φ
g
1) (1 g)
R
1
=
X
g
(φ
g
1) (1 g
1
).
The equation
R
R
1
=
τ
∆ results in an isomorphism between
U V
and
V U
for D(G)-bimodules U and V , given by flip follows by the action of R.
If
G
is non-abelian, then this is non-commutative and non-co-commutative.
The point of defining this is that the representations of
D
(
G
) correspond to the
G-equivariant k-vector bundles on G.
As we said, this is a general construction.
Theorem
(Mastnak, Witherspoon (2008))
.
The bialgebra cohomology
H
·
bi
(
H, H
) for a finite-dimensional Hopf algebra is equal to
HH
·
(
D
(
H
)
, k
),
where k is the trivial module, and D(H) is the Drinfeld double.
In 1990, Gerstenhaber and Schack defined bialgebra cohomology, and proved
results about deformations of bialgebras analogous to our results from the
previous chapter for algebras. In particular, one can consider infinitesimal
deformations, and up to equivalence, these correspond to elements of the 2nd
cohomology group.
There is also the question as to whether an infinitesimal deformation is
integrable to give a bialgebra structure on
V k
[[
t
]], where
V
is the underlying
vector space of the bialgebra.
Theorem
(Gerstenhaber–Schack)
.
Every deformation is equivalent to one where
the unit and counit are unchnaged. Also, deformation preserves the existence of
an antipode, though it might change.
Theorem
(Gerstenhaber–Schack)
.
All deformations of
O
(
M
n
(
k
)) or
O
(
SL
n
(
k
))
are equivalent to one in which the comultiplication is unchanged.
We nwo try to deform
O
(
M
2
(
k
)). By the previous theorems, we only have
to change the multiplication. Consider O
q
(M
2
(k)) defined by
X
12
X
11
= qX
11
X
12
X
22
X
12
= qX
12
X
22
X
21
X
11
= qX
11
X
21
X
22
X
21
= qX
21
X
22
X
21
X
12
= X
12
X
21
X
11
X
22
X
22
X
11
= (q
1
q)X
12
X
21
.
We define the quantum determinant
det
q
= X
11
X
22
q
1
X
12
X
21
= X
22
X
11
qX
12
X
21
.
Then
∆(det
q
) = det
q
det
q
, ε(det
q
) = 1.
Then we define
O(SL
2
(k)) =
O(M
2
(k))
(det
q
1)
,
where we are quotienting by the 2-sided ideal. It is possible to define an antipode,
given by
S(X
11
) = X
22
S(X
12
) = qX
12
S(X
21
) = q
1
X
21
S(X
22
) = X
11
,
and this gives a non-commutative and non-co-commutative Hopf algebra. This
is an example that we pulled out of a hat. But there is a general construction
due to Faddeev–Reshetikhin–Takhtajan (1988) via
R
-matrices, which are a way
of producing a k-linear map
V V V V,
where V is a fintie-dimesnional vector space.
We take a basis
e
1
, · · · , e
n
of
V
, and thus a basis
e
1
e
j
of
V V
. We write
R
`m
ij
for the matrix of R, defined by
R(e
i
e
j
) =
X
`,m
R
`m
ij
e
`
e
m
.
The rows are indexed by pairs (
, m
), and the columns by pairs (
i, j
), which are
put in lexicographic order.
The action of
R
on
V V
induces 3 different actions on
V V V
. For
s, t {
1
,
2
,
3
}
, we let
R
st
be the invertible map
V V V V V V
which
acts like
R
on the
s
th and
t
th components, and identity on the other. So for
example,
R
12
(e
1
e
2
v) =
`m
X
i,j
e
`
e
m
v.
Definition
(Yang–Baxter equation)
. R
satisfies the quantum Yang–Baxter
equation (QYBE ) if
R
12
R
13
R
23
= R
23
R
13
R
12
and the braided form of QYBE (braid equation) if
R
12
R
23
R
12
= R
23
R
12
R
23
.
Note that
R
satisfies QYBE iff
satisfies the braid equation. Solutions to
either case are R-matrices.
Example. The identity map and the twist map τ satisfies both.
Take V to be 2-dimensional, and R to be the map
R
`m
ij
=
q 0 0 0
0 1 0 0
0 q q
1
1 0
0 0 0 q
,
where q 6= 0 K. Thus, we have
R(e
1
e
1
) = qe
1
e
2
R(e
2
e
1
) = e
2
e
1
R(e
1
e
2
) = e
1
e
2
+ (q q
1
)e
2
e
1
R(e
2
e
2
) = qe
2
e
2
,
and this satisfies QYBE. Similarly,
()
`m
ij
=
q 0 0 0
0 0 1 0
0 1 q q
1
0
0 0 0 q
satisfies the braid equation.
We now define the general construction.
Definition (R-symmetric algebra). Given the tensor algebra
T (V ) =
M
n=0
V
n
,
we form the R-symmetric algebra
S
R
(V ) =
T (V )
hz R(z) : z V V i
.
Example. If R is the identity, then S
R
(V ) = T (V ).
Example. If R = τ, then S
R
(V ) is the usual symmetric algebra.
Example. The quantum plane O
q
(k
2
) can be written as S
R
(V ) with
R(e
1
e
2
) = qe
2
e
1
R(e
1
e
1
) = e
1
e
1
R(e
2
e
1
) = q
1
e
1
e
2
R(e
2
e
2
) = e
2
e
2
.
Generally, given a
V
which is finite-dimensional as a vector space, we can
identify (V V )
with V
V
.
We set
E
=
V V
=
End
k
(
V
)
=
M
n
(
k
). We define
R
13
and
R
24
:
E E
E E
, where
R
13
acts like
R
on terms 1 and 3 in
E
=
V V
V V
, and
identity on the rest; R
24
acts like R
on terms 2 and 4.
Definition
(Coordinate algebra of quantum matrices)
.
The coordinate algebra
of quantum matrices associated with R is
T (E)
hR
13
(z) R
24
(z) : z E Ei
= S
R
(E),
where
T = R
24
R
1
13
.
The coalgebra structure remains the same as
O
(
M
n
(
k
)), and for the antipode,
we write
E
1
for the image of
e
1
in
S
R
(
V
), and similarly
F
j
for
f
j
. Then we map
E
1
7→
n
X
j=1
X
ij
E
j
F
j
7→
n
X
i=1
F
i
X
ij
.
This is the general construction we are up for.
Example. We have
O
q
(M
2
(k)) = A
(V )
for
R
`m
ij
=
q 0 0 0
0 1 0 0
0 q q
1
1 0
0 0 0 q
,