• Keine Ergebnisse gefunden

Equal rank subgroups

Im Dokument Lie algebras (Seite 133-138)

n+ 1 n−1.

In multiplying all of these terms together there is a huge cancellation and what is left for the dimension of this fundamental representation is

(2n+ 1)(2n−2)

2 .

Notice that this equals 2n

2

−1 = dim ∧2V −1.

More generally this dimension argument will show that the fundamental repre-sentations are the kernels of the contraction maps i(Ω) : ∧k → (V)∧k−2(V) where Ω is the symplectic form.

ForBn it is easy to check that ωi :=L1+· · ·+Li (i≤n−1), andωn =

1

2(L1+· · ·+Ln) are the basic weights and the Weyl dimension formula gives the value

2n+ 1 j

for j ≤ n−1 as the dimensions of the irreducibles with these weight, so that they are∧j(V), j= 1, . . . n−1 while the dimension of the irreducible corresponding toωn is 2n. This is the spin representation which we will study later.

Finally, forDn=o(2n) the basic weights are ωj =L1+· · ·+Lj, j≤n−2, and

ωn−1:= 1

2(L1+· · ·+Ln−1+Ln) andωn:= 1

2(L1+· · ·+Ln−1−Ln).

The Weyl dimension formula shows that the the first n−2 fundamental repre-sentations are in fact the representation on ∧j(V), j= 1, . . . , n−2 while the last two have dimension 2n−1. These are the half spin representations which we will also study later.

7.10 Equal rank subgroups.

In this section we present a generalization of the Weyl character formula due to Ramond-Gross-Kostant-Sternberg. It depends on an interpretation of the Weyl denominator in terms of the spin representation of the orthogonal group O(g/h), and so on some results which we will prove in Chapter IX. But its logical place is in this chapter. So we will quote the results that we will need.

You might prefer to read this section after Chapter IX.

Letpbe an even dimensional space with a symmetric bilinear such that p=p+⊕p

is a direct sum decomposition ofpinto two isotropic subspaces. In other words p+andp are each half the dimension ofp, and the scalar product of any two vectors inp+ vanishes, as does the scalar product of any two elements of p. For example, we might takep=n+⊕n and the symmetric bilinear form to be the Killing form. Thenp±=n± is such a desired decomposition.

The symmetric bilinear form then putsp± into duality, i.e. we may identify p with p+ and vice versa. Suppose that we have a commutative Lie algebra hacting onpas infinitesimal isometries, so as to preserve eachp±, that thee+i are weight vectors corresponding to weightsβi and that the ei form the dual basis, corresponding to the negative of these weights−βi. In particular, we have a Lie algebra homomorphismνfromhtoo(p), and the two spin representations ofo(p) give two representations ofh. By abuse of language, let us denote these two representations by Spin and Spin−ν. We can also consider the characters of these representations of h. According to equation (9.22) (to be proved in Chapter IX) we have

In the case that his the Cartan subalgebra of a semi-simple Lie algebra and andp±=n± we recognize this expression as the Weyl denominator.

Now letgbe a semi-simple Lie algebra andr⊂ga reductive subalgebra of the same rank. This means that we can choose a Cartan subalgebra ofgwhich is also a Cartan subalgebra ofr. The roots ofrform a subset of the roots ofg.

The Weyl groupWgacts simply transitively on the Weyl chambers ofgeach of which is contained in a Weyl chamber forr. We choose a positive root system forg, which then determines a positive root system forr, and the positive Weyl chamber forgis contained in the positive Weyl chamber forr.

Let

C⊂Wg

denote the set of those elements of the Weyl group ofgwhich map the positive Weyl chamber ofginto the positive Weyl chamber forr. By the simple transi-tivity of the Weyl group actions on chambers, we know that elements ofCform coset representatives for the subgroup Wr⊂Wg. In particular, the number of elements ofC is the same as the index ofWrinWg.

Let

ρg and ρr

denote half the sum of the positive roots of g and r respectively. For any dominant weight λ of g the weight λ+ρg lies in the interior of the positive Weyl chamber for g. Hence for each c ∈C, the element c(λ+ρg) lies in the interior forrand hence

c•λ:=c(λ+ρg)−ρr

7.10. EQUAL RANK SUBGROUPS. 135 is a dominant weight forr, and each of these is distinct.

LetVλdenote the irreducible representation ofgwith highest weightλ. We can consider it as a representation of the subalgebra r. Also the Killing form (or more generally any ad invariant symmetric bilinear form) on ginduces an invariant form on r. Let p denote the orthogonal complement of r in g. We thus get a homomorphism ofrinto the orthogonal algebrao(g/r), which is an even dimensional orthogonal algebra, and hence has two spin representations.

To specify which of these two spin representations we shall denote by S+ and which byS, we note that there is a one dimensional weight space with weight ρg−ρr, and we let S+ denote the spin representation which contains that one dimensional space. The spaces S± are o(g/r) modules, and via the homomor-phismr→o(g/r) we can consider them asrmodules.

Finally, for any dominant integral weightµof r we let Uµ denote the irre-ducible module of rwith highest weightµ.

With all this notation we can now state

Theorem 16 [G-K-R-S]In the representation ring R(r) we have Vλ⊗S+−Vλ⊗S=X

c∈C

(−1)cUc•λ. (7.31) Proof. To say that the above equation holds in the representation ring of r means that when we take the signed sums of the characters of the representations occurring on both sides we get equality. In the special case thatr=h, we have observed that (7.31) is just the Weyl character formula:

χ(Irr(λ)(χ(S+g/h)−χ(S−g/h)) = X

w∈Wg

(−1)we(w(λ+ρg)).

The general case follows from this special case by dividing both sides of this equation byχ(S+r/h)−χ(S−r/h). The left hand side becomes the character of the left hand side of (7.31) because the weights that go into this quotient via (9.22) are exactly those roots ofgwhich are not roots ofr. The right hand side becomes the character of the right hand side of (9.22) by reorganizing the sum and using the Weyl character formula forr. QED

Chapter 8

Serre’s theorem.

We have classified all the possibilities for an irreducible Cartan matrix via the classification of the possible Dynkin diagrams. The four major series in our clas-sification correspond to the classical simple algebras we introduced in Chapter III. The remaining five cases also correspond to simple algebras - the “excep-tional algebras”. Each deserves a discussion on its own. However a theorem of Serre guarantees that starting with any Cartan matrix, there is a corresponding semi-simple Lie algebra. Any root system gives rise to a Cartan matrix. So even before studying each of the simple algebras in detail, we know in advance that they exist, provided that we know that the corresponding root system exists.

We present Serre’s theorem in this chapter. At the end of the chapter we show that each of the exceptional root systems exists. This then proves the existence of the exceptional simple Lie algebras.

8.1 The Serre relations.

Recall that ifαandβ are roots,

hβ, αi:= 2(β, α) (α, α)

and the string of roots of the form β+jαis unbroken and extends from β−rα to β+qα where r−q=hβ, αi.

In particular, ifα, β∈∆ so thatβ−αis not a root, the string is β, β+α, . . . , β+qα

where

q=−hβ, αi.

Thus

(adeα)−hβ,αi+1eβ= 0, 137

foreα∈gα, eβ ∈gβ but

(adeα)keβ6= 0 for 0≤k≤ −hβ, αi, ifeα6= 0, eβ6= 0. So if ∆ ={α1, . . . , α`} we may choose

ei∈gαi, fi∈g−αi

so that

e1, . . . , , e`, f1, . . . , f` generate the algebra and

[hi, hj] = 0, 1≤i, j,≤` (8.1)

[ei, fi] = hi (8.2)

[ei, fj] = 0 i6=j (8.3)

[hi, ej] = hαj, αiiej (8.4) [hi, fj] = −hαj, αiifi (8.5) (adei)−hαjii+1ej = 0 i6=j (8.6) (adfi)−hαjii+1fj = 0 i6=j. (8.7) Serre’s theorem says that this is a presentation of a (semi-)simple Lie algebra.

In particular, the Cartan matrix gives a presentation of a simple Lie algebra, showing that for every Dynkin diagram there exists a unique simple Lie algebra.

Im Dokument Lie algebras (Seite 133-138)