• Aucun résultat trouvé

COHOMOLOGY OF LIE ALGEBRAS

N/A
N/A
Protected

Academic year: 2022

Partager "COHOMOLOGY OF LIE ALGEBRAS"

Copied!
10
0
0

Texte intégral

(1)

3(b) Cohomology of Lie algebras

Let G be any connected semisimple Lie group, K ⊂ G( R ) a maximal compact subgroup, (ρ, V ) a finite-dimensional complex representation of G, Γ ⊂ G(R) a torsion-free discrete arithmetic subgroup. We let X = G(R)/K be the symmetric space attached to G and define the local system

V ˜ = Γ\(X × V )→X Γ = Γ\X.

Here Γ acts diagonally on X × V . Note that Γ = π 1 (X Γ ) and ˜ V is a local system.

Define

C ( ˜ V ) = {f ∈ C (G(R), V ) | f (γg) = ρ(γ )f (g), γ ∈ Γ, g ∈ G(R)} = Ind G Γ (V ).

(We write G for G(R) in much of this section.) This is a representation space for G and g, under the right regular representation R. There is an isomorphism of G-modules

C ( ˜ V ) −→ C (Γ\G(R)) ⊗ V ; f 7→ F (g) = ρ(g) −1 f(g).

Indeed, if h ∈ G then R h (F )(g) = ρ(gh) −1 f (gh), hence under this isomorphism, R h (f ) maps to

[g 7→ ρ(g −1 )R h (f)(g) = ρ(h)·ρ((gh) −1 )R h (f )(g) = ρ(h)·R h F (g) = (R⊗ρ(h)F )(g)].

We want to compute the cohomology groups H (Γ\G, V ˜ ) and especially H (Γ\X, V ˜ ).

Now ˜ V is locally constant (in the euclidean topology) and so its cohomology is com- puted by the complex of global sections of the (twisted) de Rham complex:

0→ V ˜ →A 0 ( ˜ V ) →A d 1 ( ˜ V ) → d . . . →A d N ( ˜ V )→0.

Here A i ( ˜ V ) is the locally constant sheaf that on a small ball U is just the C differential i-forms on U with coefficients in V . Then

A i ( ˜ V ) := Γ(Γ\G, V ˜ ) = A i (G, V ) Γ .

Now A i (G, V ) is the space of smooth functions on G that at the point g takes V i

T G,g to V . But V i

T G,g = V i

g. So A i (G, V ) = C (G, Hom(

^ i

g, V )) = Hom(

^ i

g, C (G, V )).

We will compute exterior differentiation and then look at what this does to the Γ-invariant subspaces.

Typeset by AMS -TEX

1

(2)

Differential forms and differentials.

The exterior derivative of a differential form is calculated by linearity using the formula

d(f ω 1 ∧ ω 2 ∧ . . . ω q ) = df ∧ ω 1 ∧ ω 2 ∧ . . . ω q ) + X q

i=1

(−1) i +1 f ω 1 ∧ · · · ∧ dω i ∧ . . . ω q .

This simplifies if ω i = dx i for a coordinate system x 1 , . . . , x n , whose derivatives commute, but we have identified differentials with linear maps from the Lie alge- bra to functions, and elements of the Lie algebra certainly don’t commute. Let X 1 , . . . , X N be a basis for g, ω 1 , . . . , ω N the dual basis that parallelizes the cotan- gent space. Note in any case that if f ∈ C (G) then df = P

X j (f )ω j . On the other hand, d 2 f = 0, and this allows us to compute dω i for each i, as follows:

0 = d 2 f = d( X

X j (f )ω j ) = X

j

d(X j (f ) ∧ ω j ) + X

k

X k (f )dω k

= X

i,j

X i ◦ X j (f )ω i ∧ ω j + X

k

X k (f)dω k

= X

i<j

(X i ◦ X j − X j ◦ X i )(f )ω i ∧ ω j + X

k

X k (f)dω k

= X

i<j

[X i , X j ](f )ω i ∧ ω j + X

k

X k (f)dω k .

Now we write [X i , X j ] = P

k c j ij X k and thus 0 = X

k

X k (f)[ X

i<j

c k ij ω i ∧ ω j + dω k ].

But this is true for any f , and by letting f vary, we see that the term in brackets vanishes, in other words

dω k = − X

i<j

c k ij ω i ∧ ω j .

Continuing the manipulations, we eventually find that if

ω ∈ A q (G, V ) = Hom(

^ q

g, C (G, V ))

dω(Y 0 ∧ · · · ∧ Y q ) = X q

j =0

(−1) j Y j (ω(Y 0 ∧ · · · ∧ Y ˆ j ∧ · · · ∧ Y q ))

+ X

r<s

ω([Y r , Y s ] ∧ Y 0 ∧ · · · ∧ Y ˆ r ∧ · · · ∧ Y ˆ s ∧ · · · ∧ Y q )

The action Y j (ω(•)) is right-differentiation of functions. The complete calculation

can be found in Knapp, Lie Groups, Lie Algebras, and Cohomology, pp. 155-160.

(3)

The Lie algebra complex.

If (π, W ) is any (complex) U (g)-module, we can define a complex C (g, W ) = Hom( V •

g, W ) with differential given by the same formula df (Y 0 , . . . , Y q ) = X

(−1) j π(Y i )f (Y 0 , . . . , Y ˆ j , . . . , Y q )+ X

r<s

f ([Y r , Y s ], Y 0 , . . . , Y ˆ r , . . . , Y ˆ s , . . . , Y q ).

Let H q (g, W ) = ker(d q )/im(d q −1 ). Then

H 0 (g, W ) = {f ∈ Hom( C , W ) | df (X) = 0, ∀X ∈ g} = {f ∈ Hom( C , W ) | π(X ) = 0, ∀X ∈ g}.

In other words

H 0 (g, W ) = W g := Hom U ( g ) (C, W ).

Theorem. The functor W 7→ W g is left-exact and W 7→ H q (g, W ) are its right- derived functors.

Proof. Since W g := Hom U(g) (C, W ), we know that the functor is left-exact and its right-derived functors are given by Ext q U(g) (C, W ). So we need to identify C (g, W ) with Hom U ( g ) (C , W ) where C is an acyclic resolution of C in the category of U (g)-modules. This will be sketched later in the setting of (g, K )-modules.

We certainly don’t want to compute the cohomology of C (G, V ); but we have seen that

A i ( ˜ V ) = A i (G, V ) Γ = Hom(

^ i

g, C (Γ\G) ⊗ V ) So

H q (Γ\G, V ˜ ) = H q (g, C (Γ\G) ⊗ V ).

This is still not what we want to compute, which is H (X Γ , V ˜ ) = H (Γ\G/K, V ˜ ).

We can compute this by a complex of differential forms A (X Γ , V ˜ ) = A (X, V ) Γ and there is a functorial embedding

i K : A q (X, V ) Γ ֒ → A q (G, V ) Γ .

given by G 7→ X ; g 7→ gK (pullback of differentials). The image consists of f : V q

g→C (Γ\G) ⊗ V such that

(1) f (Y 0 , . . . , Y q−1 ) = 0 if one of the Y i ∈ k = Lie(K );

(2) f ∈ (A q (G, V ) Γ ) K (right-invariant under K).

Here (1) says that the pullback of differentials from X to G vanish on vectors tangent to K , and (2) says that the coefficients of the differential are functions on X = G/K. More precisely – ignore Γ for the moment, since the right and left actions don’t interfere: Say f (g) ∈ Hom( V q

T G,g , V );

f (gk) ∈ Hom(

^ q

T G,gk , V ) −→

R(k

−1

) Hom(

^ q

T G,g , V )

and the condition that f be K -invariant is that R(k −1 )f (gk) = f(g). But R(k −1 )

acting on left-invariant vector fields is just Ad(k) acting on g. So to conclude

(4)

Lemma. The image of i K (A q (X, V ) Γ ) is Hom K (

^ q

(g/k), C (Γ\G, V )) = Hom K (

^ q

(g/k), C (Γ\G) ⊗ V ).

Define C (Γ\G) 0 ⊂ C (Γ\G) to be the subspace of K -finite vectors: a vector whose translates under K generate a finite-dimensional subspace. Then it is clear that for any q,

Hom K (

^ q

(g/k), C (Γ\G) ⊗ V ) = Hom K (

^ q

(g/k), C (Γ\G) 0 ⊗ V ).

because both V and V q

(g/k) are finite-dimensional.

Thus for any representation (π, W ) of U (g) on which the action of k integrates to a consistent K -finite action of K, we define

C q (g, K; W ) = Hom K (

^ q

(g/k), W ).

This is a subspace of C q (g, W ) and it is easy to see that it is preserved by the differential, so that we have a complex and can define the relative Lie algebra cohomology H q (g, K; W ).

Proposition. H (X Γ , V ˜ ) −→ H (g, K; C (Γ\G) 0 ⊗ V ).

Again H q (g, K ; •) is the right-derived functor of a left-exact functor on the cat- egory of (g, K)-modules (see below). In what follows, G is a connected reductive Lie group, K ⊂ G is a maximal compact subgroup (modulo the center of G) and g = Lie(G), k = Lie(K ).

Definition. 1. A (g, k)-module is a U (g)-module whose restriction to U (k) is semi- simple and a sum of finite-dimensional k-modules.

2. A (g, K )-module is a (g, k)-module (π, V ) whose k action integrates to an action of K (note: K may be disconnected) in such a way that, for k ∈ K and X ∈ g, we have π(k)π(X)π(k −1 )v = π(ad(k)X)v for all v ∈ V .

3. A (g, K)-module V is admissible if for any irreducible representation τ of K, Hom K (τ, V ) is finite-dimensional.

One also calls (g, K )-modules Harish-Chandra modules. They form an abelian category (a full subcategory of the category of U (g)-modules) with enough projec- tives and injectives. If V, W are (g, K)-modules, one can compute Ext (V, W ) – extensions in the category of (g, K)-modules – by using a projective resolution of V :

. . . →P i →P i −1 → . . . →P 0 →V ; C i = Hom(P i , W ); Ext q (V, W ) = H q (C ).

Let P i = U (g) ⊗ U (k) V i

(g/k) and define ∂ q : P q →P q−1 by the expected formula

∂ q (r ⊗ x 1 ∧ · · · ∧ x q ) = X

(−1) i−1 x i · r ⊗ x 1 ∧ · · · ∧ x ˆ i ∧ · · · ∧ x q )

+ X

i<i

(−1) i+j r ⊗ ([x i , Y j ] ∧ x 1 ∧ · · · ∧ x ˆ i ∧ · · · ∧ x ˆ j ∧ · · · ∧ x q )

Let ε : P 0 = U (g) ⊗ U ( k ) C→C be the augmentation.

(5)

Theorem. The P j are projective, the maps ∂ q and ε are well-defined, and the sequence . . . →P q →P q −1 → . . . →→P 0 →C is exact.

Proof. The detailed proofs are contained in §VII.8 of Knapp’s yellow book. There are two points:

(1) tensor product U (g) ⊗ U (k) • takes projectives to projectives and everything in the category is U (k)-projective;

(2) The sequence is exact.

Exactness is a long calculation that reduces ultimately using filtrations to the exactness of the Koszul complex. The first point can be easily explained. Let V be any (g, K )-module and let U be a K-module. Let I(U ) = U (g) ⊗ U (k) U . We check that this is a (g, k)-module (that it is a (g, K )-module follows easily). Let W ⊂ U (g) be a finite-dimensional subspace invariant under ad(k), X ∈ W , Y ∈ U (k); then

Y X ⊗ u = [Y, X ] ⊗ u + X ⊗ Y u ⊂ W ⊗ U.

So the action is semisimple and I (U ) is a (g, k)-module.

Suppose f : B→A is a surjective morphism of (g, K )-modules. Now Frobenius reciprocity is a canonical isomorphism

Hom (g,K ) (I (U ), •) = Hom K (U, •).

So the map Hom (g,K) (I (U ), B)→Hom (g,K) (I(U ), A) is surjective if and only if Hom K (U, B)→Hom K (U, A) is surjective; but this is clear because B and A are semisimple as K-modules. Thus I(U ) is projective for any U , and in particular all the P j are projective.

Corollary. The functors H (g, K ; •) are the right-derived functors of W 7→ Hom g,K (C, W ).

In particular short exact sequences give rise to long exact sequences in the usual way.

Definition. Let U be a (g, K )-module. The contragredient of U is the subspace U ˜ ⊂ Hom(U, C ) of vectors on which K acts finitely.

Suppose U is admissible, so that as K -module, U = ⊕ σ∈ K ˆ U (σ) with U (σ) = Hom K (σ, U ) ⊗ σ finite-dimensional. Then ˜ U = ⊕ σ∈ K ˆ U (σ) where ∗ is the usual contragredient for finite-dimensional representations. In particular, ˜ U is again ad- missible.

Definition. Let χ : Z (g)→C be an algebra homomorphism. The (g, K )-module U has infinitesimal character χ if zu = χ(z)u for all z ∈ Z(g), u ∈ U .

Corollary. Suppose U and V are (g, K )-modules with infinitesimal characters χ U and χ V , respectively. Suppose V is finite-dimensional and χ U 6= χ V ˜ . Then H q (g, K ; U ⊗ V ) = 0 for all q.

Proof. The natural equivalence of (bi)-functors:

Hom ( g,K ) (C, U ⊗ V ) −→ Hom ( g,K ) ( ˜ V , U )

(6)

gives rise to isomorphisms of derived functors

H q (g, K ; U ⊗ V ) −→ Ext q ( ˜ V , U ).

So it suffices to show that Ext q ( ˜ V , U ) = 0 for all q. Now by hypothesis, there is z ∈ Z(g) that acts as 1 on U and as 0 on ˜ V . If q = 0 and h : ˜ V →U is a (g, K )- morphism then h(v) = zh(v) = h(zv) = 0. If q > 0, let S ∈ Ext q ( ˜ V , U ) correspond to a Yoneda extension, i.e. a long exact sequence of (g, K)-modules

0→U →E q −1 → . . . →E 0 → V ˜ →0.

Then z acts consistently on this sequence and as the identity on U and as 0 on ˜ V , and defines an equivalence with the trivial Yoneda extension.

Alternatively, multiplication by z defines two natural transformations of bifunc- tors (A, B) 7→ Ext q (g,K) (A, B), say z 1 and z 2 , by acting in the first and second variable, respectively. For q = 0 we have z 1 = z 2 . Suppose they are equal up to q−1. Let B→C→B be an exact sequence of (g, K )-modules with C injective. Then the exact cohomology sequence breaks up as Ext j−1 (U, V )→Ext j (U, V ) which is surjective for j = 1 and an isomorphism for j ≥ 2, and which commutes with z 1 and z 2 . This reduces the case of q to that of q − 1. In our setting, z 1 = 0 and z 2 = 1, so we are done.

Compact quotients.

Suppose Γ\G is compact. Then C (Γ\G) 0 ⊂ L 2 (Γ\G) where the measure defin- ing L 2 is any right-invariant Haar measure dg on G = G(R). It is not hard to show that if G is reductive, then any right-invariant Haar measure is also left-invariant.

Now L 2 (Γ\G) is a Hilbert space on which G acts by a unitary representation – unitary because

< r(h)f, r(h)f > L

2

= Z

Γ\ G

f(gh)f (gh)dg = Z

Γ\ G

f (g)f (g)dg =< f, f > L

2

, where the second equality follows from invariance of dg.

The following theorem is due to Gelfand and Piatetski-Shapiro:

Theorem. Assume Γ\G is compact. Then L 2 (Γ\G) decomposes as G-module as a countable Hilbert space direct sum:

L 2 (Γ\G) = M d

π

m(π, Γ)π

where π runs over irreducible unitary representations of G and m(π, Γ) ∈ N. In particular, the multiplicity of π is always finite, and is zero except for a countable set of π.

The classification of irreducible unitary representations is incomplete, but the

following theorem is fundamental. Choose a maximal compact (or compact con-

nected) subgroup K of G. We define a K -finite vector in a Hilbert space repre-

sentation (π, V) as above; a smooth vector v ∈ V is one for which, for any element

X ∈ Lie(G) and any w ∈ V the function t 7→< π(exp(tX )v, w > from R to V is

infinitely differentiable.

(7)

Theorem(Harish-Chandra). Let π be an irreducible unitary Hilbert space rep- resentation of the reductive Lie group G and let π 0 ⊂ π denote the subspace of K- finite smooth vectors. Then U (g) acts naturally on π 0 by π(X)v = dt d π(exp(tX)v and makes it an irreducible (g, K) module.

This is proved in several stages, the most important being that any irreducible representation of K occurs with finite multiplicity in π 0 – in other words, that π 0 is an admissible (g, K) module. This implies that every K -finite vector is automati- cally smooth. The proofs can be found in the beginning of Chapter VIII of Knapp’s book Representation Theory of Semisimple Groups.

We see that L 2 (Γ\G) 0 = C (Γ\G) 0 and that this in turn is L d

π m(π, Γ)π 0 . Thus Theorem. Assume Γ\G is compact. Then for any finite-dimensional representa- tion V of G,

H (X Γ , V ˜ ) −→ M d

π

m(π, Γ)H (g, K; π 0 ⊗ V ).

Thus the calculation of the cohomology of X Γ divides into two parts: a global part, which is the determination of m(π, Γ), and a purely Lie-theoretic part, which is the calculation of H (g, K ; π 0 ⊗ V ). The first part is very hard, the second part was solved some time ago. First I switch to the adelic setting.

Theorem (Borel-Harish-Chandra). Let G be a reductive group over Q with

center Z , and for any open compact subgroup K ⊂ G(A f ), let K S(G) = G(Q)\G(A)/K Z(R)×

K . Then K S(G) is compact for one K if and only it it is compact for all K if and only if G/Z has Q-rank 0; in other words, if G/Z does not contain a subgroup isomorphic to GL(1).

One says that G/Z is anisotropic if it is of Q-rank 0. The adelic version of the Gelfand-Piatetski-Shapiro theorem is the following:

Theorem. Let G be a (connected) reductive group over Q with center Z , and assume G/Z is anisotropic. Then L 2 (G(Q)\G(A)) decomposes as G(A)-module as a countable Hilbert space direct sum:

L 2 (G(Q)\G(A)) = M d

π

m(π)π

where π runs over irreducible unitary representations of G(A).

We fix a maximal compact subgroup K ⊂ G(R). With π as in the theorem, a vector v ∈ V π is smooth if it is C with regard to the action of G(R) and if there is an open compact subgroup K f ⊂ G(A f ) that fixes v. Write π 0 ⊂ π for the space of K -finite smooth vectors. Then

(1) π 0 determines π (and vice versa); in particular, π 0 is irreducible as a repre- sentation of (U (g), K ) × G(A f );

(2) π 0 admits a unique factorization (up to scalar multiples) π 0 −→ π ⊗ ⊗ p π p

where the product is taken over all prime numbers, π is an irreducible

(g, K )-module, and each π p is an irreducible (smooth) representation of

G(Q p ).

(8)

If V is a representation of G, we can then define H (g, K ∞ ; π⊗V ) := H (g, K ∞ ; π ∞ ⊗ V ) ⊗ π f where π f = ⊗ p π p is an irreducible smooth representation of G(A f ). If K f ⊂ G(A f ) is open compact, we can similarly define

H (g, K ; π K

f

⊗ V ) := H (g, K ; π ⊗ V ) ⊗ π f K

f

.

Working through the comparison of K S(G) with a union of spaces of the form Γ i \G(R)/K ∞ Z(R), we find

Proposition. For any representation V of G, there is a canonical isomorphism

H ( K S(G), V ˜ ) −→ M

π

H (g, K ; π ⊗ V ) ⊗ π f K

f

;

H (S(G), V ˜ ) −→ M

π

H (g, K ; π ⊗ V ) ⊗ π f ,

where the latter isomorphism commutes with the action of G(A f ) on both sides.

Automorphic vector bundles. Now assume (G, X ) a Shimura datum, so that X is of Hermitian type. Fix h ∈ X and let [W ] be the automorphic vector bundle on S(G, X ) attached to a finite-dimensional representation of K h . We can replace the de Rham complex by the Dolbeault complex in the discussion above. Instead of

A i (G, V ) = C (G, Hom(

^ i

g, V )) = Hom(

^ i

g, C (G, V )).

we have

A 0 ,q ([W ]) = [C (G(Q)\G(A), Hom(

^ i

(p ), W ))] K

h

= Hom K

h

(

^ i

(p ), C (G(Q)\G(A)) ⊗ W )

= [

^ i

(p ) ⊗ C (G(Q)\G(A)) ⊗ W ] K

h

.

Now k h ⊕ p is the Lie algebra of a subgroup P h ⊂ G C (not defined over R and the definition of relative Lie algebra cohomology in this case identifies A 0 ,q ([W ]) with C q (Lie(P h ), K h ; C (G(Q)\G(A)) ⊗ W ). Thus Dolbeault’s theorem gives us an isomorphism

H q (S(G, X ), [W ] −→ H q (Lie(P h ), K h ; C (G( Q )\G(A)) ⊗ W ).

Again, we can replace C (G(Q)\G(A)) by L

π m(π)π 0 and then we find an iso- morphism of G(A f )-spaces

H q (S(G, X ), [W ]) −→ M

π

m(π)H q (Lie(P h ), K h ; π ⊗ W ) ⊗ π f .

(9)

Hodge theory for (g, K)-modules.

The analytic arguments used to prove the Hodge theorem in complex geometry becomes completely algebraic in the setting of relative Lie algebra cohomology.

Change notation: let π be a (g, K )-module that comes from a unitary representation of G. We want to compute H q (g, K ; π ⊗ V ) for any irreducible finite-dimensional representation (ρ, V ). We already know that this vanishes unless π and V have opposite infinitesimal characters. To apply Hodge theory, we endow W with an admissible scalar product: one that is invariant under K and with respect to which the action of p is self-adjoint. The existence of such a scalar product is easy to show:

V is a representation of G, therefore of the compact real form G u , and therefore possesses a G u -invariant (hermitian) scalar product (unique up to positive scalar multiples), say (, ) V . This means that Lie(G u ) = k ⊕ ip acts by skew-adjoint operators: if X ∈ Lie(G u ) then (Xv, v ) + (v, Xv ) = 0. For X = iY with Y ∈ p this means

0 = i(Y v, v ) + ¯ i(v, Y v ) = i[(Y v, v ) − (v, Y v )]

which means that p is self-adjoint.

Let D q (π ⊗ V ) = V q

p ⊗ π ⊗ V with scalar product given by the tensor product of the hermitian Killing form on p C :

(x, y) = B(x, y) ¯

(dualized and taken to the q-th power) with the scalar products already defined on the other two factors. All of these scalar products are invariant under k. Write τ = π ⊗ ρ, so that τ(x) = π(x) ⊗ 1 + 1 ⊗ ρ(x) for x ∈ g. The adjoint with respect to the scalar product on τ is given by

τ (x) = −τ (x), x ∈ k; τ (x) = −π(x) + ρ(x), x ∈ p.

In what follows, we choose a (real) basis x 1 , . . . , x D of p (and don’t confuse the dimension D with the differential d. If η ∈ D q (π ⊗ V ) = Hom( V q

p, π ⊗ V ) and J ⊂ {1, . . . , D}, |J| = q, we write

η J = η(x j

1

, . . . , x j

q

).

Proposition. Define d : D q (π ⊗ V )→D q −1 (π ⊗ V ) by the formula (d η) J =

X D

j=1

τ (x j ) η {j}∪J .

Then d commutes with K, maps the K -invariant subspace C q (π ⊗ V ) into the K -invariant subspace C q −1 (π ⊗ V ), and is formally adjoint to d:

(d η, f ) = (η, df ) for η ∈ D q , f ∈ D q−1 .

Proof. We recall the formula for df (with a shift of indices):

df (Y j

1

, . . . , Y q+1 )

= X

j

(−1) j−1 τ(Y j )f (Y 1 , . . . , Y ˆ j , . . . , Y q +1 ) + X

r<s

f ([Y r , Y s ], Y 1 , . . . , Y ˆ r , . . . , Y ˆ s , . . . , Y q +1 )

= X

j

(−1) j −1 τ(Y j )f (Y 1 , . . . , Y ˆ j , . . . , Y q +1 )

(10)

because [p, p] ⊂ k so the bracket terms vanish. Thus (replacing the Y u ’s by x j

u

(η, df ) = X

|I|=q

(η I , (df ) I ) τ = X

|I|=q

(η I , X

u

(−1) u −1 τ (x j

u

)f I(u) ) τ

where I = {j 1 , . . . , j q } and I(u) is I with the j u term removed. And thus by adjunction

(η, df) = X

I,u

((−1) u−1 τ (x j

u

)η I , f I ( u ) . Using the anticommutation formula

η I = (−1) u −1 η {j

u

}∪I(u)

this shows after reviewing the notation that (η, df) = (d η, f ).

Now if x ∈ k we find

−(τ (x)d η, f ) = (τ (x) d η, f ) = (d η, τ (x)f) = (η, dτ (x)f )

= (η, τ (x)df ) = (−τ(x)η, df ) = −(d τ (x)η, f ).

Thus d commutes with K and in particular takes the K-invariants to K-invariants.

Let ∆ = dd + d d. For each q, ∆ is an endomorphism of D q (π ⊗ V ) and of C q (π ⊗ V ). Moreover, for η ∈ D q (π ⊗ V )

(∆η, η) = (dη, dη) + (d η, d η) and since the scalar product is positive non-degenerate,

∆η = 0 ⇔ dη = 0 and d η = 0 ⇔ (∆η, η) = 0.

In this case we say η is harmonic; and we let H q (π ⊗ V ) ⊂ C q (π ⊗ V ) denote the subspace of harmonic forms.

Proposition. For every q, the map H q (π ⊗ V )→H q (g, K ; π ⊗ V ) is an isomor- phism.

Proof. This is equivalent to the orthogonal decomposition C q (π ⊗ V ) = H q (π ⊗ V ) ⊕ Im(d) ⊕ Im(d ).

Note for example that for η ∈ C q ,

dη = 0 ⇔ (dη, f) = 0 ∀f ∈ C q+1 ⇔ η ⊥ Im(d ).

So C q = ker(d) ⊕ Im(d ) is an orthogonal decomposition. Similarly, Z q := ker d = H q ⊕ Im(d) is an orthogonal decomposition because H q = Z q ∩ ker d .

Kuga’s formula calculates the action of ∆ in terms of a specific element, the

Casimir element, in Z (g).

Références

Documents relatifs

In this section, basically following [18] and [2], we recall some basic facts about the cohomology of Lie-Rinehart algebras, and the hypercohomology of Lie algebroids over

Let f rs ∈ F p [g] ∩ k[g] G be the polynomial function from Section 1.4 with nonzero locus the set of regular semisim- ple elements in g, and let f e rs be the corresponding

The complex structure of a complex manifold X is a rich source of different cohomological structures, for example, the Dolbeault cohomology or the holomorphic de Rham cohomology,

Let us start with giving a few well-known facts about the inhomogeneous pseudo-unitary groups n2) is the real Lie group of those complex transformations of an

It is known in general (and clear in our case) that i\0y has as its unique simple quotient the socle of i^Oy. Passing to global sections we see that r(%»0y) is generated by

Similar complexes were introduced later by other authors in the calculus of variations of sections of differential fibrations and cohomology theorems were proved [1], [5],

We close this section by making some remarks on the relation between the cohomology of M.(n} and the cohomology of maximal nilpotent subalgebras of semisimple Lie algebras. Hence

Another important property satisfied by the cohomology algebra of a compact K¨ ahler manifold X is the fact that the Hodge structure on it can be polarized using a K¨ ahler class ω ∈