1991 Mathematics Subject Classification. 22E47. The research of the first named author was partially supported by RGC-CERG grants of Hong Kong SAR and National Nature Science Foundation of China. The research of the second named author was partially supported by a grant from the Ministry of Science and Technology of Republic of Croatia. The second and third named authors were also supported by the joint French-Croatian program COGITO. Parts of this work were done during authors' visits to CNRS, University of Paris VII, Mathematisches Forschungsinstitut Oberwolfach, and Institute of Mathematical Sciences and Department of Mathematics at the National University of Singapore. The authors thank these institutions for their generous support and hospitality.
Dirac operators and Lie algebra cohomology
Jing-Song Huang, Pavle Pandžić,
David Renard
Department of Mathematics, Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong SAR, China E-mail address : mahuang@ust.hk Department of Mathematics, University of Zagreb, Bijenicka 30, 10000 Zagreb, Croatia E-mail address : pandzic@math.hr Centre de mathematiques Laurent Schwartz, Ecole Polytechnique, 91128 Palaiseau Cedex, France E-mail address : renard@math.polytechnique.fr
-
Abstract.
Dirac cohomology is a new tool to study unitary and admissible representations of semisimple Lie groups. It was introduced by Vogan and further studied by Kostant and ourselves [V2] , [HP1] , [K4] . The aim of this paper is to study the Dirac cohomology for the Kostant cubic Dirac operator and its relation to Lie algebra cohomology. We show that the Dirac cohomology coincides with the corresponding nilpotent Lie algebra cohomology in many cases, while in general it has better algebraic behavior and it is more accessible for calculation.
1 Introduction
In 1928 Dirac discovered a matrix valued first-order differential operator as a square root of the Laplacian operator in order to understand elementary particles. Since then this operator and its various analogues are called Dirac operators in the scientific community.
They have played an important role in physics as well as in mathematics; as an example, let us mention the famous Atiyah-Singer Index Theorem. Applications to representation theory of reductive Lie groups were started by Parthasarathy [P] , who defined a Dirac operator on the symmetric space
and used it for geometric realization of most discrete series representations. This approach was further developed by Atiyah and Schmid [AS] , who constructed all discrete series representations as kernels of the Dirac operator acting on the associated spinor bundles.
On the other hand, Lie algebra cohomology (with respect to nilpotent subalgebras) has been a fundamental invariant of representations of Lie algebras and groups starting from Kostant's 1962 paper [K1] , which gives an algebraic proof of the Bott-Borel-Weil theorem.
Lie algebra cohomology is related to characters, asymptotics of matrix coefficients, embeddings into various “standard” modules and various geometric realizations of representations.
It has been clear for decades that the Dirac operators are formally similar to the differentials of the de Rham or Dolbeault cohomology. There have been two problems with using that analogy in representation theory. First, the Dirac operators on non-symmetric homogeneous spaces are not as nicely behaved. Second, the index of the Dirac operator behaves well only for some unitary representations, like the discrete series; so the algebraic tools of representation theory do not work well with it. The first problem was resolved by Kostant [K2] , who introduced the modified cubic Dirac operator that works well also on non-symmetric homogeneous spaces. The second problem was resolved by Vogan [V2] , who introduced the concept of Dirac cohomology which applies both to unitary and nonunitary representations: it is the kernel of the Dirac operator
modulo the intersection of the image and the kernel. He made a conjecture on the infinitesimal character of irreducible representations with nonzero Dirac cohomology, which was proved in [HP1] . Kostant generalized this result to the case of his cubic Dirac operator and applied it to the topology of homogeneous spaces [K4] . Recently, Kumar [Ku] and Alekseev and Meinreken [AM] found further generalizations of the results in [HP1] and [K4] . They put these results into the broader setting of non-commutative equivariant cohomology.
The aim of this paper is to explore the relation between Dirac cohomology and (nilpotent) Lie algebra cohomology, in those situations when both can be defined. Our results show that Dirac cohomology coincides with the corresponding Lie algebra cohomology for a large family of unitary representations, including the discrete series representations, and for all unitary representations in certain special cases. On the other hand, Dirac cohomology seems to have a better algebraic behavior.
We have also been able to use Dirac cohomology and the proved Vogan's conjecture to obtain improvements of some classical results. In [HP2] we describe how to simplify certain parts of Atiyah-Schmid's construction of discrete series representations [AS] and sharpen the Langlands' formula on automorphic forms [L] , [HoP] . The details will appear in a forthcoming book by the first and second named authors. This suggests that Dirac cohomology may be useful also in other situations, as a new tool for tackling problems not accessible to classical cohomology theories.
Let us now describe our main results on Dirac cohomology and its relation to (nilpotent) Lie algebra cohomology more precisely. Let
be a connected semisimple Lie group with finite center and complexified Lie algebra
. Let
be a Cartan involution of
(and
), let
be the corresponding maximal compact subgroup, and let
be the complexified Cartan decomposition. As usual, the corresponding real forms will be denoted by
,
, etc. (One could also work in a more general setting of a reductive group
in the Harish-Chandra class.) If
is of hermitian type, then
decomposes as a sum of two abelian subalgebras,
and
. We prove a Hodge decomposition for
-cohomology and
-homology for all unitary
-modules, and show that they are both isomorphic to the Dirac cohomology up to a twist by a modular character. We note that in this setting Enright [E] found an explicit formula for the
-cohomology of unitary highest weight modules.
For general
, and any two reductive subalgebras
to which the Killing form restricts nondegenerately, we show that Kostant's cubic Dirac operator
can be decomposed as a sum of two anti-commuting Dirac operators
. Here
denotes a certain “diagonal embedding”. Decompositions of this kind also appear in [AM] .
In particular, we consider the case when
is a reductive subalgebra, which is real, i.e.,
is the complexification of
. Then we can show that for any admissible
-module, the Dirac cohomology with respect to
is the same as the kernel of
on the Dirac cohomology with respect to
. Furthermore, the Dirac cohomology with respect to
is also the same as the Dirac cohomology with respect to
of the kernel of
.
In particular, if
and
have equal rank, then
can be a Levi subalgebra
of a
-stable parabolic subalgebra
of
, with
. It follows that the Dirac cohomology with respect to
, which is closely related to the
-cohomology, has the advantage of being much easier to calculate. For instance, let
be a compact Cartan subalgebra, and
be a Borel subalgebra. Then we show how to (easily) explicitly calculate the Dirac cohomology of the discrete series representations with respect to the Dirac operator
. Comparing the obtained result with Schmid's formula for
-cohomology in [S] , we see that they are the same up to an expected modular twist. In case
is of hermitian type, the Dirac cohomology with respect to
coincides with the
-cohomology or
-homology (up to a modular twist) for all unitary representations.
We now describe the organization of our paper. In Sections 2 and 3 we decompose the Kostant cubic Dirac operator as sum of two “half Dirac operators”, which correspond to the differentials of
-cohomology and
-homology. Section 3 also contains a new proof of the Casselman-Osborne theorem on Lie algebra cohomology using an approach similar to that of [HP1] . This is aimed at explaining the formal similarity of the two results. In Section 4 we prove a Hodge decomposition for
-cohomology or
-homology for unitary representations in the Hermitian case, with the Dirac cohomology providing the “harmonic representatives” of both. The same proof applies to any finite-dimensional representation of an arbitrary semisimple group. In Section 5 we define the relative Dirac operators and show that in some cases Dirac cohomology can be calculated in stages. In Section 6 we develop this further and show how to use it to calculate Dirac cohomology explicitly. In particular, we compare the Dirac cohomology for a Levi subalgebra with the
-cohomology in some cases. In Section 7 we obtain a Hodge decomposition and equality of Dirac and
-cohomology for arbitrary unitary modules in the Hermitian case. We conclude the paper by showing that the homological properties of Dirac cohomology are quite different from those of Lie algebra cohomology. In fact, under certain conditions, we show that there is a six-term exact sequence of Dirac cohomology corresponding to a short exact sequence of
-modules. So Dirac cohomology resembles a K-theory, rather than a cohomology theory.
This work was initiated by David Vogan [V3] . We would like to thank him for many interesting and stimulating conversations. We believe that the results in this paper are not the end of the theory of Dirac cohomology in representation theory, but rather the beginning of further investigations and applications. For example, the results of this paper should be related to the results of Connes and Moscovici [CM] in a similar way as the results of [HP1] are related to [AS] .
2 Construction of certain Dirac operators
Let
be a complex semisimple Lie algebra,
a parabolic subalgebra,
the opposite parabolic subalgebra, and
. Then
Furthermore, the restrictions of the Killing form
to
and
are non-degenerate, and the above decomposition is orthogonal. Since
and
are isotropic subspaces in perfect duality under
, we can identify
with
; this identification is
-invariant. Let
be a basis of
, and let
be the dual basis of
.
Let
be the Clifford algebra of
. Unlike in [HP1] , we will use the same defining relations as Kostant, namely
in particular, if
, then
, and not
like in [HP1] . Of course, over
there is no substantial difference between the two conventions.
We are going to make use of the well known principle of constructing invariants by contracting dual indices. The aim is to construct a family of interesting
-invariants in
. These will include Kostant's cubic Dirac operator
, but we will get
as a sum of four members of the family, and we will also be able to combine them in different ways, to get other operators with properties similar to the properties of
. For example, we will have nice expressions for their squares. The form of this principle we need is the statement of the following lemma; the proof is quite easy and essentially reduces to the fact that under the identification
, the identity map corresponds to the sum
.
Lemma 2.1.
Let
be a linear map which is
-equivariant with respect to the adjoint actions.
Then
is independent of the chosen basis
and
-invariant. Here
ranges over all
-tuples of integers in
,
, and
. □
For example,
can be composed of the obvious inclusions
and
, products, commutators in
and the Killing form
. Here are several examples of this kind which we will study in the following:
Examples 2.2.
| |
| |
| |
Note the symmetry obtained by exchanging the roles of
and
. To see how the Dirac operator fits in here, note that one can build an orthonormal basis
of
from
and
, by putting
for
. Then it is easy to check that
Also, we can rewrite
,
and
as follows:
Lemma 2.3.
| |
| |
| |
-
Proof.
Since
, we can write it as
Also, the sum in the definition of
is clearly twice the same sum over only those
for which
. The only issue is thus to pass from the Clifford product to the wedge product. For this, we use (1.6) in [K2] . First, since
is isotropic,
. Next, we calculate
The second two terms here are clearly zero if
is different from
and
. For
, the second term is
while the third is zero. For
, the second term is zero while the third is
. It follows that we will be done if we can show
However, using Lemma 2.1, we see that
is an
-invariant element of
. Since there are no nonzero
-invariants in
, this sum must be
. Analogously,
So we proved the first equality for
. Now in the form with wedge product, we can clearly commute
in front of
and
, and then we obtain the second equality by reversing the above argument.
The formulas for
are obvious from the defining relations of
. □
Consider now the basis
of
, given by
the dual basis is then
Notice that for any
Hence we can write Kostant's cubic element
as
In particular, we obtain Kostant's cubic Dirac operator as
We are also particularly interested in the elements
Note that
. We will use the fact that commuting with
operates on
and
in the following way:
Proposition 2.4.
-
Proof.
The second two relations follow from the first two, and the first two are immediate from the following lemma. □
Lemma 2.5.
Commuting with
in the Clifford algebra
acts as
on
and as
on
.
-
Proof.
Clearly, for
,
. For
, we calculate
Namely, since
is isotropic,
for any
. The first claim now follows, and the second is analogous. □
Kostant [K2] , Theorem 2.16, has calculated
Here
denotes the Casimir element of
. Further,
is the Casimir element for the diagonal copy
of
, embedded into
via
where
is the action map followed by the standard inclusion of
into
using the identification
. Finally,
is the constant
.
Using this result and the above remarks, we can now quickly calculate the squares of
,
and
:
Proposition 2.6.
-
Proof.
From Kostant's expression for
, it is clear that
commutes with all
-invariant elements of
. In particular,
commutes with
, and using Proposition 2.4 we see
Since
and
, it follows that
so
=
. On the other hand, Proposition 2.4 implies that
So we see that
, and also
. □
Now we can completely describe the Lie superalgebra
spanned by
,
,
and
inside the superalgebra
. Here
is an associative superalgebra with
-grading of the Clifford factor; so it is also a Lie superalgebra in the usual way, with the supercommutator
.
is a subalgebra of the Lie superalgebra
, with the commutation relations from Propositions 2.4 and 2.6. Namely,
is central, and
The last relation is obtained as follows:
.
Note that we can regard
as a
-graded Lie superalgebra with,
and
of degrees
,
,
and
respectively. Note also that the subalgebra of
spanned by
,
and
is a Heisenberg superalgebra.
The Lie superalgebra
can be identified with
, i.e., the endomorphisms of the superspace
, with the first
even and the second
odd. In Kac's classification [Kac] it is denoted by
. It was extensively used by physicists under the name supersymmetric algebra. It is a completely solvable Lie superalgebra, and its irreducible finite-dimensional representations are described in [Kac] .
To finish this section, let us note that
is independent not only of the choice of basis
but also of the choice of
. On the other hand,
,
and
do depend on the choice of
.
3
-cohomology and
-homology
We retain the notation from previous sections. Let
be an admissible
-module.
For
, let
be the set of
-cochains of the complex defining the
-cohomology of
. The differential
is given by the usual formula
We have the following identifications :
Like in the previous section, we fix a basis
of
and denote the dual basis of
by
.
Lemma 3.1.
Through the above identifications, the differential
is given by
where
denotes the projection of
on
.
-
Proof.
This is a straigtforward calculation, starting from the fact that the identification
is given via
. □
The space
is also the space of
-chains for the
-homology of
, with differential
given by :
Note that we are tensoring with
from the right and not from the left as usual; this is because we will have an action of
on
which will be more natural in this order. For reasons that will become apparent later, we will instead of
consider the operator
. Of course,
defines the same homology as
.
To get our Dirac operators act, we need to consider the
module
, where
is the spin module for the Clifford algebra
. We will use the identification of
with
, given explicitly in [K4] and [K3] . Namely, one can construct
as the left ideal in
generated by the element
. One then has
, which is isomorphic to
as a vector space, and the action of
is given by left Clifford multiplication. Explicitly,
and
act on
by
Namely, since
and
are isotropic, the Clifford and wedge products coincide on each of them; in particular,
.
The natural action of
on
is the tensor product of the restriction of the
-action on
and the spin action on
. On the other hand, the usual
action on
-cohomology and
-homology is given by the adjoint action on
and
. Thus, our identification of
with
is not an
-isomorphism. However, as was proved in [K3] , Prop.3.6, the two actions differ only by a twist with the one dimensional
-module
of weight
.
This means that, if we consider
and
as operators on
via the above identification, then as an
-module, the cohomology of
gets identified with
, while the homology of
gets identified with
.
Proposition 3.2.
Under the action of
on
, the operators
and
from Section 2 act as
and
respectively.
In particular, the cubic Dirac operator
acts as
.
-
Proof.
We are going to use the explicit formulas for the action of
on
given above.
From these formulas, it is immediate that the action of
coincides with the first (single) sum in the expression for
, while the action of
transforms
into
Since
, we see that this is equal to minus twice the first (single) sum in the expression for the
-homology operator
.
It remains to identify the action of the cubic terms
and
.
We use the expression for
from Lemma 2.3, i.e.,
. This element transforms
into
| |
| |
Now we sum
, and after commuting
into its proper place, we get the second (double) sum in the expression for
.
For
we use its definition from 2.2. To write the calculation nicely, we introduce the following notation: for
, let
If
then we change the sign and define
If
, we set
. This now allows us to write
It follows that
transforms
into
Upon summing up
and
, we get that this is equal to
This is now clearly invariant for exchanging the roles of
and
, hence it is twice the same sum extending just over
, i.e., minus twice the second (double) sum in the expression for
. □
It is now clear why we considered
instead of just
; in this way we have the action of
being equal to
.
Before we go on, let us note how the element
of Section 2 acts on
; it is in fact a degree operator up to a shift. This means
can be used to identify the degrees in which the cohomology (homology) is appearing.
Proposition 3.3.
The element
of
from 2.2 acts on
as multiplication by the scalar
. Consequently, it preserves the kernel and image of both
and
, and hence acts on the
-th cohomology of
and the
-th homology of
, by the same scalar
. In particular,
is the degree operator.
-
Proof.
Using Lemma 2.5 we see that for any
,
. On the other hand, by Lemma 2.3,
, hence
. It now immediately follows that in
we have
, so the action of
on
is indeed multiplication by the scalar
.
It now immediately follows that
preserves the kernel and image of
and
, as these are homogeneous operators (of degree
and
respectively). (Note that this last assertion can also be obtained from the commutation relations of Proposition 2.4.) □
We will now state a result for the operators
and
analogous to the one obtained for
in [HP1] and [K4] . A corollary will be the Casselman-Osborne Theorem. Our goal here is not to give a new proof of the Casselman-Osborne Theorem, the existing ones being completely satisfactory, but to shed some light on the formal similarity between the Casselman-Osborne Theorem and the main result of [HP1] .
Define
by
| |
| |
| |
where
is 1 for even
and
for odd
. In the following we fix a compact group
with complexified Lie algebra
.
Theorem 3.4.
and
are
-equivariant. They induce maps from
into itself and
on
.
Furthermore
(resp.
,
), and one has
, (resp.
,
).
-
Proof.
The result for
is due to Kostant [K4] . The proof is the same as the proof of the main result of [HP1] . We give details for
, the proof for
being entirely similar. As in [HP1] , we use the standard filtration on
, which induces a filtration
on
. This filtration being
-invariant, it induces in turn a filtration on
.
Clearly,
The
-gradation on the Clifford algebra
induces a
-gradation on
. We set
and
. Then
and this
-gradation is compatible with the filtration
If
, then
. If
, then
. Thus
induce
Let
be the image of
in
(notice that the cubic term disappears since it is in
). If
,
and if
,
Therefore
. Note also that unlike
,
is a differential on all of
, because
. Hence
is a differential on
.
Let us compute
for
. We can assume
are different from each other and likewise for
.
If
for all
, then the contribution to the sum is zero. If
, then
So we see
Since
and
, one has
It follows that
is (up to a sign depending on the
-degree) equal to
, where
is the Koszul differential for the vector space
. It is well known that
is exact except at degree zero, where the cohomology is
, embedded as the constants. It follows that
is exact except at degree zero, where the cohomology is
embedded in the obvious way. It remains to pass to the invariants:
Lemma 3.5.
The differential
on
is exact except at degree zero. The zeroth cohomology is
, embedded in the obvious way. More precisely
To prove the lemma, we need to show that
. To see this, we may choose some element
in the center of
, such that
is the centralizer of
in
,
has real eigenvalues, and
is the sum of the eigenspaces of
corresponding to the positive eigenvalues. Making
act on an element in
, we see that this element has to be in
. We can now finish the proof of the theorem. We proceed as in [HP1] by induction on the degree of the filtration. □
Let
. Since
, we can write
for some
and
.
Our goal is now to compute
. Let
be a Cartan subalgebra of
, and let us denote respectively by
and
the Weyl groups of
in
and
.
We have Harish-Chandra isomophisms
and an obvious inclusion
. Set
. With this notation we have:
Lemma 3.6.
For all
,
.
-
Proof.
The proof is similar to the proof of Theorem 4.2 in [K4] , but much simpler. We give only a sketch. Let
be a Borel subalgebra of
containing
, and suppose the Cartan algebra
has been chosen to lie in
.
Let
be the irreducible finite dimensional representation with highest weight
(relative to
), and let
be a non-zero highest weight vector in
. Recall the element
in
used to define the spin module
.
One can see easily that
and that
defines a non-zero cohomology class in
. Since the infinitesimal character of
is given by
, and any
acts by the scalar
on
(see [K4] , Theorem 4.1), we get
for all
.
Since this is true for all dominant weights
, and since these form a Zariski dense set in
, we conclude that indeed
. □
Let
be a representation of
admitting an infinitesimal character
and let
.
Then
acts on
by the scalar
.
Notice that
acts on
as
. Namely,
leaves
and
stable, and induces the zero action on
.
Thus, the induced action of
on
is equal to the scalar multiplication by
. This is exactly the statement of the Casselman-Osborne Theorem (see [CO] , or [V1] , Theorem 3.1.5). Namely, our definition of
differs from the map
from the Casselman-Osborne Theorem exactly by the above
-shift.
4 Hodge decomposition for
cohomology
In this section we study positive definite hermitian forms on
for unitary modules
, such that
and
are minus adjoint to each other. Here unitarity is defined with respect to a fixed real form
of
; we also fix a corresponding Cartan involution
and Cartan decompositions
and
.
Such a form however exists only in very special situations; if
is simple, then the pair
must be hermitian symmetric,
must be equal to
, and
and
must be the abelian subalgebras
of
. In that case, we get a Hodge type decomposition:
-homology,
-cohomology and the Dirac cohomology are all represented by the space of “harmonics”,
.
A parallel result holds much more generally if
is a finite-dimensional module. Since this case is known, we include it only as a remark. However we start by studying a sufficiently general situation which applies to both cases.
Let
be a subalgebra of
to which
restricts non-degenerately and let
be the orthogonal complement to
with respect to
. We choose a maximal isotropic subspace
of
. Since
(with
denoting conjugation with respect to
) defines a positive definite Hermitian form on
and hence also on
, the subspace
intersects
trivially. Let
be a spin module for the Clifford algebra
corresponding to this polarization.
We extend the form
to all of
in the usual way, using the determinant.
Lemma 4.1.
With respect to the form
on
, the adjoint of the operator
is
.
-
Proof.
We need to show that
for any
and
. Assume
. We can assume that
and
, with
,
and
all belonging to a fixed orthonormal basis of
.
Then both sides of the equality to be established are nonzero precisely when
is one of the
's, while the other
's are precisely (all) the
's. In this case both sides are clearly the same.
If
the claim follows by considering
. In case
is odd-dimensional, the extra element is handled easily. □
We now assume that
is real and
-stable, i.e.,
is the complexification of a subalgebra
with
.
Lemma 4.2.
(a) The cubic part
of Kostant's cubic Dirac operator
is self-adjoint with respect to the form
on
.
(b) If
for a parabolic subalgebra
, so that
as in the previous sections, then
is adjoint to
.
-
Proof.
(a) Let us choose bases
of
and
of
orthonormal with respect to
. By Lemma 4.1, the adjoint of
is
and the adjoint of
is
. Moreover, the dual bases of
,
with respect to
are
,
, and so
(the other terms are zero). Since the adjoint of
is
, the adjoint of
is
, and the coefficients are real, we see that
is self-adjoint.
(b) By the formulas from Lemma 2.3, it is enough to show that
is adjoint to
. This however follows immediately from Lemma 4.1: we can normalize the
's so that
, hence the adjoint of
is
□
Let now
be a unitary
-module and let
be a
-stable parabolic subalgebra. Then the adjoint of the operator
on
with respect to the unitary form is
, while the adjoint of
on
is either
, or
, depending on whether
is in
or in
. If we consider the tensor product hermitian form on
, denoted again by
, and as before denote by
and
the
respectively
parts of
, then we see:
Corollary 4.3.
With respect to the form
on
, the adjoint of
is
while the adjoint of
is
. Hence the adjoint of
is
and the adjoint of
is
. □
We will use this corollary in Sections 5 and 7. Now we turn our attention to the case when
and
do not appear in
. This is the already mentioned case, when
, and
is contained in
. Then
is forced to be abelian and we denote as usual
,
. In this case, the Dirac operator
is the “ordinary” Dirac operator corresponding to
, there is no cubic part, and we conclude
Corollary 4.4.
Let
be a hermitian symmetric pair and set
. Let
be a unitary
-module and consider the form
on
. Then the operators
and
are minus adjoints of each other. Hence the Dirac operator
is anti-self-adjoint.
More generally, if
contains
(this can happen when
is not simple), then
is anti-self-adjoint with respect to
. □
For the rest of this section we assume that
. So
is anti-self-adjoint with respect to the positive definite form
on
. In particular, the operators
and
have the same kernel on
.
By [P] , Proposition 3.2 (or more generally, [K2] , Theorem 2.16), we know that
where
and
are the Casimir operators for
and diagonally embedded
, and
is the constant
. It follows that if
acts on
by a constant, then
is up to a constant equal to
on
. Since
acts by a scalar on each
-type in
, the same is true for
. So
is a semisimple operator, i.e.,
is a direct sum of eigenspaces for
. In particular:
Corollary 4.5.
If the
-module
has infinitesimal character, then
-
Proof.
We have seen that
is a direct sum of eigenspaces for
. Clearly, the zero eigenspace is
, and the sum of the nonzero eigenspaces is
. □
It is now easy to obtain a variant of the usual Hodge decomposition. The following arguments are well known; see e.g. [W] , Scholium 9.4.4. We first have
Lemma 4.6.
(a)
; (b)
is orthogonal to
and
is orthogonal to
.
-
Proof.
(a) Since
, it is clear that
is contained in
. On the other hand, if
, then
, hence
. So
, hence
. Now
implies that also
.
(b) is obvious since
and
are minus adjoint to each other. □
Combining Corollary 4.5, Lemma 4.6 and the fact
, we get
Theorem 4.7.
Let
be a hermitian symmetric pair and set
and
. Let
be an irreducible unitary
-module. Then: (a)
; (b)
; (c)
.
In particular, the Dirac cohomology of
is equal to
-cohomology and to
-homology, up to modular twists:
More precisely, (up to modular twists) the Dirac cohomology
is the space of harmonic representatives for both
-cohomology and
-homology.
□
Remark 4.8.
There is a variant of the above results for a finite dimensional module
. In this case one puts a positive definite form on
invariant under the compact form
of
; this is sometimes called an admissible form. In this way the adjoint of
on
will be
if
and
if
. Hence the adjoint of
on
is
for all
.
So this form combines with
on
better than a unitary form. In particular,
and
are now adjoint to each other with respect to
,
is self-adjoint, and the above arguments, including Theorem 4.7, apply without change. There is no need here to assume that
;
can be any parabolic subalgebra of (any)
, with the assumption that
is
-stable and real.
This case was however already known; it is implicit in [
K3]
and it is explicitly mentioned in [
V3]
.
Moreover, we can generalize this further and prove self-adjointness of
for the more general setting when
is not necessarily a Levi subalgebra of a parabolic subalgebra, but any real and
-stable reductive subalgebra (to which
then restricts nondegenerately). Namely, getting back to the setting of Lemma 4.2 (a) and its proof, we see that not only the cubic term
but also the linear term
of
is self adjoint with respect to
on
. Hence
is self-adjoint.
Remark 4.9.
Here is an example which shows that a Hodge decomposition like in Theorem 4.7 does not hold generally if
is not compact, even for unitary modules in the Hermitian symmetric case. Let
be the
-module of the even Weil representation of
, and let
be the
containing the compact Cartan subalgebra and corresponding to the short noncompact root.
Then
is a Levi factor of a
-stable parabolic subalgebra with abelian
. One can explicitly calculate everything and show that the Dirac cohomology is strictly larger than
-homology or
-cohomology (while the latter two are still equal).
5 Relative Dirac operators
In this section we compare various Dirac operators arising from two compatible decompositions
Here both decompositions are like in the Kostant's setting for the cubic Dirac operator:
and
are reductive subalgebras of
such that the Killing form
restricts to a nondegenerate form on each of them, while
respectively
are the orthogonals of
respectively
. Compatibility of the two decompositions means
Clearly, all these decompositions are orthogonal.
Example 5.1.
If
is a
-stable Levi subalgebra of
, then
and
satisfy the above conditions.
Now
is also a reductive subalgebra of
to which
restricts nondegenerately. The corresponding decomposition is
The corresponding Dirac operator will be denoted by
. We note that this setting includes the case in which
is a subalgebra of
, and therefore the setting
from the introduction.
To write down
, take orthonormal bases
for
, and
for
. Identify
|
(1)
|
where
denotes the
-graded tensor product. Now if
is the union of the bases
and
, then by Kostant's definition,
is the element
of
.
Kostant's original definition uses exterior multiplication instead of Clifford multiplication in the second sum (with the Clifford and exterior algebras identified as vector spaces via Chevalley identification). For orthogonal vectors, there is however no difference between exterior and Clifford multiplication, so the above definition is the same as Kostant's. Taking into account the decomposition (1), we see
(2)
D
(
g
,
r
∩
r
′
)
=
∑
i
Z
i
⊗
Z
i
⊗
1
+
∑
j
Z
j
′
⊗
1
⊗
Z
j
′
−
1
2
∑
i
<
j
<
k
B
(
[
Z
i
,
Z
j
]
,
Z
k
)
⊗
Z
i
Z
j
Z
k
⊗
1
−
1
2
∑
i
<
j
∑
k
B
(
[
Z
i
,
Z
j
]
,
Z
k
′
)
⊗
Z
i
Z
j
⊗
Z
k
′
−
1
2
∑
i
<
j
<
k
B
(
[
Z
i
′
,
Z
j
′
]
,
Z
k
′
)
⊗
1
⊗
Z
i
′
Z
j
′
Z
k
′
.
Note that the terms with
,
and
do not appear, because
, as
is orthogonal to
.
We consider
as the subalgebra
of
.
In view of this, we see that the first and third sum in (2) combine to give
, the Kostant's cubic Dirac operator corresponding to
.
The remaining three sums come from the cubic Dirac operator corresponding to
. However, this is an element of the algebra
, and this algebra has to be embedded into
diagonally, not as
. Namely, we use the diagonal embedding
from the setting
; the definition is the same as for
above Proposition 2.6. Thus we embed
We will denote
by
and call it a relative Dirac operator. In case when there are several subalgebras and confusion might arise, we will use more precise notation
instead of
and give up the notation
. This diagonal embedding has already been used by Kostant and Alekseev-Meinrenken; in particular, decompositions like the following one can be founded in [AM] .
Theorem 5.2.
With notation as above, (i)
; (ii) The summands
and
anticommute.
-
Proof.
To prove (i), we need to describe the image under
of
|
(3)
|
and see that it matches the second, fourth and fifth sum in (2). In fact, it is obvious that the image under
of the second sum in (3) equals the fifth sum in (2), and it remains to identify
|
(4)
|
Namely,
, where
is the action map of
on
, followed by the standard inclusion of
into
, given by
where
are the matrix units relative to the basis
. Thus we are left with showing that the second sum in (4) equals the third sum in (2), i.e., that
This will follow if we see that
for any
. But this last equality is clear from the definition of
. Namely, the matrix of
on
in the basis
is
.
To prove (ii), we use the fact that
commutes with
, which is one of the most basic properties of
. It follows that the anticommutator
is zero for any
. Hence
. It remains to see that also
This follows from the definition of
, since all the
-parts of the monomial terms of
, and also all
, are odd. □
Example 5.3.
In the setting of Example 5.1, we obtain
and both decompositions have anticommuting summands. If
is contained in
, which is possible if and only if
and
have equal ranks, then we get only one nontrivial decomposition,
. This case will be of special interest below.
We now want to use Theorem 5.2 to relate the Dirac cohomology of the various Dirac operators involved. In some cases one can apply the following easy fact from linear algebra.
We define the cohomology of any linear operator
on a vector space
to be the vector space
.
Lemma 5.4.
Let
and
be anticommuting linear operators on a vector space
. Assume that
diagonalizes on
, i.e.
, with
on
. Then the cohomology
of
on
is the same as the cohomology of the restriction of
to
.
-
Proof.
Since
commutes with
, its kernel, image and cohomology decompose accordingly to eigenspaces
. We thus have to prove that
has no cohomology on
for
. In other words, we are to prove that
on
.
Let
be such that
, i.e.,
. Then
and hence
is in the image of
. □
Corollary 5.5.
In the setting of Lemma 5.4, assume further that
; so
. Then
is equal to the cohomology of
restricted to the cohomology (i.e., kernel) of
.
-
Proof.
By Lemma 5.4,
is the cohomology of
on
. But on
,
. □
To apply this to Dirac cohomology, denote by
the Dirac cohomology of a
-module
with respect to
; analogous notation will be used for other Dirac operators. The reader should bear in mind that
is in fact the cohomology of the operator
on the space
.
Corollary 5.6.
Let
be a reductive subalgebra of
, with
nondegenerate.
As usual, let
be the orthocomplement of
. Assume that either
is even, or
is even1
. Let
be an irreducible unitary
-module. Then
i.e., the Dirac cohomology can be calculated “in stages”, as the
-cohomology of the
-cohomology.
-
Proof.
Since
is unitary, we can consider the form
on
introduced in Section 4, where
is the spin module for
. We can extend this form to all of
, by choosing any positive definite form on the spin module
for
. Here we identify
, which can be done by the assumption on dimensions. Let
and
.
By Corollary 4.3,
is anti-self-adjoint, and consequently the conditions of Corollary 5.5 are satisfied. So the cohomology with respect to
is the cohomology with respect to
of
.
Now
is a
-module, with Lie algebra
acting through
. The Dirac cohomology of this module with respect to
is thus identified with the cohomology with respect to
. □
6 The case of compact Levi subalgebra
In this section we first consider a reductive subalgebra
of
contained in
. Later on we will specialize to the case when
is a Levi subalgebra of a
-stable parabolic subalgebra of
.
The first thing we will do in this situation is generalize Corollary 5.6 to nonunitary modules. Like there, we assume for simplicity that either
is even or
is even, so that we can write the spin module as
. The idea is to reverse the roles of
and
. Namely, for any admissible
-module
, we can decompose
into a direct sum of finite dimensional (unitary) modules for the spin double cover
of
. Hence, by Remark 4.8, there is a positive definite form
on
, such that
is self-adjoint with respect to
.
It follows that
is a semisimple operator, while for
we still have that
is semisimple. In this situation, we have the following lemma which complements Corollary 5.5:
Lemma 6.1.
Let
and
be anticommuting linear operators on a vector space
, such that
and
are semisimple (i.e., can be diagonalized). Then
is the cohomology (i.e., the kernel) of
acting on
.
-
Proof.
Applying Lemma 5.4, we can replace
by
, i.e., assume
. On the other hand, by Corollary 5.5,
is the cohomology of
acting on
.
Since
is semisimple, we can decompose
into the (discrete) sum of eigenspaces for
. Here if both
and
are eigenvalues, we choose one of them to represent the pair. Since
anticommutes with
, it preserves
, and maps
to
and vice versa. Therefore,
decomposes into a
-part and
-parts. The
-part is equal to
and we will be done if we show that
has no kernel on
. Let
be in
, and assume that
. This implies
is in
, so
. This however can only happen if both
and
are in
, again because
exchanges
and
. But then also
is in
, so
is zero in cohomology and we are done.
□
This now immediately implies the following theorem which says that Dirac cohomology with respect to a subalgebra
as above can be calculated “in stages”.
Theorem 6.2.
Let
be a reductive subalgebra of
contained in
. Let
be an admissible
-module. Then the Dirac cohomology with respect to
can be calculated as the Dirac cohomology with respect to
of the Dirac cohomology with respect to
of
. In other words:
Also, we can reverse the order of taking Dirac cohomology, i.e.,
-
Proof.
The first formula was explained above, and the second is a direct application of Corollary 5.5, with
and
(opposite from Corollary 5.6). □
For the rest of this section we consider a
-stable parabolic subalgebra
of
, with the Levi subalgebra
contained in
. In particular, there is a Cartan subalgebra
of
contained in
; so
and
have equal rank. The opposite parabolic subalgebra is
. As before, we denote
, so
.
We apply the above considerations to
. Since
is a finite dimensional
-module, and
and
have equal rank,
is given by [K4] , Theorem 5.1. (This can also be read off from Remark 4.8, by using Poincaré duality to pass to
-cohomology and then the better known Kostant's formula for
-cohomology; in fact, this is how Kostant proves it.) This gives
very explicitly provided we know
explicitly. For example, one can in this way calculate the Dirac cohomology of the discrete series representations with respect to the (compact) Cartan subalgebra
:
Example 6.3.
Let
be a discrete series representation; here
is a Borel subalgebra of
containing a compact Cartan subalgebra
. The infinitesimal character of
is
. Then the Dirac cohomology of
with respect to
consists of a single
-type
, whose highest weight is
, where
. This is obtained from the highest weight of the lowest
-type of
,
, by shifting by
(the lowest weight of
).
This result is contained in the work of Parthasarathy and Schmid. One can also prove it as follows: it is shown in [
HP1]
, Proposition 5.4, that this
-type is contained in the Dirac cohomology. Since
has a unique lowest
-type, and since
is the lowest weight of the spin module, with multiplicity one, it follows that any other
-type has strictly larger highest weight, and thus can not contribute to the Dirac cohomology.
We now apply the above mentioned Kostant's formula (Theorem 5.1 of [
K4]
) to calculate the Dirac cohomology with respect to
(we again stress that
and
have equal rank):
It follows from
that
Remark 6.4.
Comparing with Schmid's formula in Theorem 4.1 of [
S]
, we have
(note that Schmid's
is our
, and his
is our
.) In other words,
-cohomology of a discrete series representation coincides with the Dirac cohomology up to a
-shift. This fact is however not covered by our results in Sections 4 and 7. This indicates that it should be possible to generalize our results; however, the next example shows they do not hold in full generality. □
Example 6.5.
In the setting of Remark 4.9, one can show that
is strictly larger than
. Here
is the compact Cartan subalgebra, and the last cohomology can be calculated in stages, as
.
We now want to similarly analyze the “half-Dirac” operators
and
. Let
be a basis for
and let
be a basis for
. These can be taken to be the root vectors corresponding to compact, respectively noncompact positive roots, with respect to some
compatible with
. We normalize these bases so that the dual bases for
respectively
with respect to the Killing form are
respectively
.
As before,
,
and
denote the Dirac operator for the pair
and its parts. We can now further decompose these parts and write
and analogously for
and
. Here
The expressions for
,
,
,
and
are obtained by exchanging
with
and
with
.
In the following, we will consider the Clifford algebra
as a subalgebra of
, embedded as
. In particular,
gets identified with
,
with
and so on.
Recall that by Theorem 5.2,
, where
is the image of
under the diagonal embedding
. Here
sends
identically onto
, and for
,
with
given by the action of
on
as before.
Clearly,
, while
is the sum of all other of the above parts. We want to make this more precise; namely, in the obvious notation,
, and we want to identify the images of these summands under
. Denote these images by
and
.
To do this, we use an expression for
in terms of a basis and a dual basis:
if
is a basis of
with dual basis
, then
(This was already used in Section 5, for an orthonormal basis. The proof of this slightly more general version is the same.) Applying this to the basis
and the dual basis
, we get
(we used
).
Since
, we see that
. We need to calculate the middle term,
. Applying the above expression for
, we get four sums over
and
.
We first notice that the first of these three sums is 0, since
is 0 on
. To calculate the second sum, write
, and observe that since
,
. Therefore the second of the three sums is
The third sum is 0, since we can assume
is in
and hence orthogonal to
.
Namely, we can choose
and
(and also
and
) to be root vectors with respect to
.
Finally, the fourth sum is calculated by noting that since
,
. It follows that the fourth sum is
A completely analogous calculation applies to
, so we proved:
Proposition 6.6.
Under the diagonal map
,
and
correspond to
Note the unexpected feature of this result, the mixing of the positive and negative parts under the diagonal embedding. Namely,
and
have opposite positions from the ones one would expect. So we do not have an analogue of Theorem 5.2 for
and
, unless
. This last thing happens precisely when the pair
is hermitian symmetric. This is the reason why we are able to obtain results about
-cohomology only in hermitian case. Maybe this peculiar behavior has something to do with the fact that some of the most concrete results about
-cohomology, like [E] , [C] or [A] , are also obtained in hermitian situation only.
Another difficulty with the non-hermitian case is the fact that while we can write
, the two summands here are not differentials (they are also not
-invariant). So there is no hope to get a Hodge decomposition like the one in Section 4.
Of course, there is also no
-homology or cohomology, since
is not a Lie algebra. Yet there is perfectly well defined Dirac cohomology for
, and one can hope that it will somehow replace the nonexistent
-cohomology.
What we do get without the hermitian assumption, is a copy of
inside
, spanned by
,
,
and
. This follows immediately from the fact that
is (obviously) a morphism of superalgebras.
If the pair
is hermitian symmetric, as we will assume in the following, then there is another copy of
inside
, supercommuting with the first one. It is spanned by
,
,
, and
. The fact that these two copies of
supercommute is completely analogous to Theorem 5.2.(ii). (Without the hermitian assumption, one could try to replace this second
by a smaller superalgebra, spanned just by
and
.) Note that
and
. Clearly, an analogous decomposition holds for
; it also holds for
. Namely, since the two copies of
supercommute, we have
This last equality can also be obtained from Kostant's formula for
:
Here we have used the more precise notation for the various diagonal embeddings to avoid confusion; note the equality
.
In other words, we see that the
corresponding to the pair
sits diagonally in the direct product of the two copies of
described above. Let us summarize the above discussion:
Corollary 6.7.
Assume
is a hermitian symmetric pair, let
be a
-stable parabolic subalgebra of
, and assume that
and
. Then there are two supercommuting copies of
inside
. One is spanned by
,
,
and
and the other is spanned by
,
,
, and
. The diagonal of the product of these two super subalgebras is the copy of
spanned by
,
,
and
from the end of Section 2. □
7 Hodge decomposition for
-cohomology in hermitian case
In this section,
is a hermitian symmetric pair,
is a
-stable parabolic subalgebra of
, and we assume
and
.
Let
be a unitary
-module, and consider the form
on
introduced in Section 4. To apply the results of Section 4, we decompose
and embed
as
. The form
restricts to the analogous definite form on
.
Denote as before by
the Dirac operator for the pair
and by
and
its parts coresponding to
and
. By Corollary 6.7,
, and similarly for
. Moreover, the copy of
corresponding to the pair
supercommutes with the copy of
corresponding to the pair
.
By Corollary 4.3, the adjoints of
and
are respectively
and
. So the adjoint of
is
.
We consider the positive semidefinite operator
. By the above remarks we have
We know from Section 4 that
decomposes into eigenspaces of
for eigenvalues
. Each eigenspace is
-invariant, and each
-isotypic component of
is contained in an eigenspace. We assume
is admissible, so the eigenspaces are finite-dimensional.
Passing from
to
is tensoring with the finite-dimensional
-module
. On this last space, there is no action of
or
. So every eigenspace of
on
just gets tensored with
, and this gives the eigenspace on
for the same eigenvalue.
Since
commutes with
, it preserves these eigenspaces. Moreover, the Levi subgroup
corresponding to
is compact. So is then the double cover
, which acts on
. Since
commutes with
, it also preserves its eigenspaces and hence these eigenspaces decompose into
-irreducibles. Since
is up to a constant equal to the Casimir element of
, it follows that
diagonalizes on each eigenspace of
. To conclude:
Lemma 7.1.
is a direct sum of eigenspaces for
. In particular,
. □
This is an analogue of Corollary 4.5 for
in place of
. Now the arguments proving Lemma 4.6 and Theorem 4.7 work without change, and we obtain
Theorem 7.2.
(a)
; (b)
; (c)
; (d)
.
In other words, we have obtained a Hodge theorem for
-cohomology.
To obtain it also for
-homology, we note that
, and so
So the situation for
is exactly the same as for
and we conclude
Theorem 7.3.
(a)
; (b)
; (c)
; (d)
.
In other words, Hodge decomposition also holds for
-homology. Moreover, we see that
-cohomology and
-homology have the same set of harmonic representatives,
. In particular they are isomorphic.
We now want to relate
-cohomology and
-homology to Dirac cohomology with respect to
. The main observation here is
Lemma 7.4.
.
-
Proof.
The operators
and
are both positive semidefinite and their sum is
. This immediately implies the first equality. The second follows from
(since
is self-adjoint) and
(since
is anti-self-adjoint). □
We can now combine Theorems 7.2 and 7.3 with Lemmas 7.4 and 4.6 to conclude
Corollary 7.5.
-
Proof.
It is obvious that the left hand side contains the right hand side. Conversely, if
, then
by Theorem 7.2,
by Lemma 7.4, so
by Lemma 4.6 and so also
. Analogously,
and
. □
Since
can be thought of as the kernel of
acting on the kernel of
, and similarly for the
-operators, in view of Theorem 4.7 and Remark 4.8 we can reinterprete Corollary 7.5 as follows:
Corollary 7.6.
To calculate the
-cohomology of
, one can first calculate the
-cohomology of
to obtain a
-module, and then calculate the
-cohomology of this module. Analogously, to calculate the
-homology of
, one can first calculate the
-homology of
, and then the
-homology of the resulting
-module. □
Remark 7.7.
Note that this is in fact the Hochschild-Serre spectral sequence for the ideal
of
respectively the ideal
of
. What we have obtained is that these Hochschild-Serre spectral sequences are always degenerate for a unitary
-module
. □
We now turn our attention to the Dirac cohomology of
. In addition to the above considerations, we bring in Corollary 5.6, and note that for both
and
the cohomology is the same as the kernel or the kernel of the square. Thus we obtain:
Theorem 7.8.
The Dirac cohomology
of a unitary
-module
is isomorphic to the
-cohomology of
and the
-homology of
up to appropriate modular twists. Moreover, all three cohomologies have the same set of harmonic representatives,
. □
8 Homological properties of Dirac cohomology
Let us start by showing that although we proved that in some cases Dirac cohomology of a unitary
-module with respect to
can be identified with
-cohomology or
-homology, one should by no means expect that these notions agree for general
-modules. Let us see that this is not the case even for
-modules.
Consider the module
which is a nontrivial extension of the discrete series representation
of highest weight
by the trivial module
:
(In other words,
is a dual Verma module.) The
-weights of
(for the basis element
of
) are
. We are considering the case
,
is spanned by
and
is spanned by
.
For any
-module
we have
with
given by
,
given by
, and
. By an easy direct calculation, we see that
as vector spaces. Here
denotes a vector in
of
-weight
. So we see
On the other hand, for
and
another easy calculation (which can be shortened by using Theorem 4.7, as both
and
are unitary) implies
To explain why Dirac cohomology of
differs from the
-cohomology of
, we will examine their behavior with respect to extensions. Recall the well known long exact sequences for Lie algebra homology and cohomology corresponding to our short exact sequence
They are
for
-homology and
for
-cohomology. Here all arrows are the obvious ones except for the one labelled by
.
For Dirac cohomology, instead of a long exact sequence (which clearly does not makes sense in general, as Dirac cohomology is not
-graded), there is a six-term exact sequence.
In the above example, this sequence is
We see that in this example the six-term sequence agrees with the
-homology long exact sequence only because of the presence of zeros; in general, all three sequences are different.
To define the six-term sequence in a more general situation, let us assume that the Dirac cohomology is
-graded. This happens whenever we are starting from
with
even-dimensional; this is automatic when
is a Levi subalgebra. Let
be a short exact sequence of
-modules. Tensor this sequence by
, and denote the arrows still by
and
(they get tensored by the identity on
). Assuming that
is a semisimple operator for each of the three modules, we can construct a six-term exact sequence
The horizontal arrows are induced by
and
. The vertical arrows are the connecting homomorphisms, defined as follows. Let
represent a Dirac cohomology class, so
. Choose
such that
. Since
is semisimple, we can assume
. Since
, we see that
for some
. Since
, we see that
, so
defines a cohomology class. This class is by definition the image of the class of
under the connecting homomorphism. Clearly, we changed parity when we applied
, and this defines both vertical arrows at once. It is easy to see that this map is well defined, and that the resulting six-term sequence is exact. To conclude:
Theorem 8.1.
Let
be an orthogonal decomposition, with
a reductive subalgebra and
even-dimensional. Let
be a short exact sequence of
-modules and assume that the square of the Dirac operator
is a semisimple operator for
,
and
. Then there is a six-term exact sequence corresponding to this short exact sequence, as described above. □
Finally, let us comment on what can be done when
is odd dimensional, say
. The usual spin modules
and
are not
-graded and thus it seems the above construction does not make sense. Recall that
and
are defined by writing
, considering the spin module for
, and letting the last basis element of
act in two different ways (preserving the even and odd subspace instead of exchanging them).
We can instead consider the unique irreducible graded module
of
. It can be constructed as the restriction of the (unique) spin module for
to
. As a non-graded module,
decomposes as
. If we define Dirac cohomology using
in place of
or
, we double it, but we do get a
-grading.
Then the above construction works also in the odd case. Thus, this is probably a more natural definition of Dirac cohomology in the odd case.
It remains to see what can be done if
is not a semisimple operator. One possibility might be to consider a more general definition of Dirac cohomology in that setting.
References
-
J. Adams, Nilpotent cohomology of the oscillator representation, J. Reine Angew. Math. 449 (1994), 1–7.
-
A. Alekseev, E. Meinrenken, Lie theory and the Chern-Weil homomorphism, math.RT/0308135.
-
M. Atiyah, W. Schmid, A geometric construction of the discrete series for semisimple Lie groups, Invent. Math. 42 (1977), 1–62.
-
A. Borel, N. Wallach, Continuous cohomology, discrete subgroups, and representations of reductive groups, Second edition. Mathematical Surveys and Monographs, 67. American Mathematical Society, Providence, RI, 2000.
-
W. Casselman, M. S. Osborne, The
-cohomology of representations with an infinitesimal character, Comp. Math. 31 (1975), 219–227.
-
C. Chevalley, The algebraic theory of spinors, Columbia University Press, 1954.
-
D. H. Collingwood, The
-homology of Harish-Chandra modules: generalizing a theorem of Kostant, Math. Ann. 272 (1985), no. 2, 161–187.
-
A. Connes, H. Moscovici, The
-index theorem for homogeneous spaces of Lie groups, Ann. of Math. 115 (1982), no. 2, 291–330.
-
T. J. Enright, Analogues of Kostant's
-cohomology formulas for unitary highest weight modules, J. Reine Angew. Math. 392 (1988), 27–36.
-
R. Hotta, R. Parthasarathy, A geometric meaning of the multiplicities of integrable discrete classes in
, Osaka J. Math. 10 (1973), 211–234.
-
J.-S. Huang, P. Pandžić, Dirac cohomology, unitary representations and a proof of a conjecture of Vogan, J. Amer. Math. Soc. 15 (2002), 185–202.
-
J.-S. Huang, P. Pandžić, Dirac operators in representation theory, Representations of Real and P-adic Groups, Lecture Notes Series, Vol. 2, Institute for Mathematical Sciences, National University of Singapore, Singapore University Press and World Scientific, 2004, pp. 163-219.
-
V. Kac, Lie superalgebras, Adv. in Math. 26 (1977), 8–96.
-
B. Kostant, Lie algebra cohomology and the generalized Borel-Weil theorem, Ann. of Math. 74 (1961), 329–387.
-
B. Kostant, A cubic Dirac operator and the emergence of Euler number multiplets of representations for equal rank subgroups, Duke Math. Jour. 100 (1999), 447–501.
-
B. Kostant, A generalization of the Bott-Borel-Weil theorem and Euler number multiplets of representations, Lett. Math. Phys. 52 (2000), 61–78.
-
B. Kostant, Dirac cohomology for the cubic Dirac operator, Studies in memory of I. Schur, in Progress of Math. vol. 210 (2003), 69–93.
-
S. Kumar, Induction functor in non-commutative equivariant cohomology and Dirac cohomology, preprint, University of North Carolina, 2003.
-
R. Langlands, The dimension of spaces of automorphic forms, Amer. J. of Math. 85 (1963), 99–125.
-
R. Parthasarathy Dirac operator and the discrete series, Ann. of Math. 96 (1972), 1–30.
-
W. Schmid,
-cohomology and the discrete series, Ann. of Math. 103 (1976), 375–394.
-
D.A. Vogan, Representations of real reductive Lie groups, Birkhäuser, Boston, 1981.
-
D.A. Vogan, Dirac operators and unitary representations, 3 talks at MIT Lie groups seminar, Fall 1997.
-
D.A. Vogan,
-cohomology in representation theory, a talk at “Functional Analysis VII”, Dubrovnik, Croatia, September 2001.
-
N. R. Wallach, Real Reductive Groups, Volume I, Academic Press, 1988.
Department of Mathematics, Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong SAR, China E-mail address : mahuang@ust.hk Department of Mathematics, University of Zagreb, Bijenicka 30, 10000 Zagreb, Croatia E-mail address : pandzic@math.hr Centre de mathematiques Laurent Schwartz, Ecole Polytechnique, 91128 Palaiseau Cedex, France E-mail address : renard@math.polytechnique.fr