The main result of this chapter is the formula, due to Weyl, for the character of any finite-dimensional irreducible module for a split semi-simple Lie algebra over a field of characteristic 0. If the base field is the field of complex numbers and R is a finitedimensional irreducible representation, then the character of R is the function on a Cartan subalgebra
defined by
where, as usual, exp a = 1 + a + a2/2! + …. If is a connected semi-simple compact Lie group, then (1) gives the character in the ordinary sense of an irreducible representation of
. Thus in this case
corresponds to a maximal torus and it is known that any element of
is conjugate to an element in this torus. Then (1) gives the values of the characters for elements of the torus. We know that hR acts diagonally, that is, a basis can be chosen so that the matrix of hR is
where Λ(h), M(h), … are the weights of in the representation. Then the matrix exp hR is diag {exp Λ(h), exp M(h), …}; hence
where nM is the multiplicity of the weight M(h), that is, the dimensionality of the weight space M.
One obtains a purely algebraic form of the definition of the character x(h) by replacing the exponentials exp M(h) of (3) by formal exponentials which are elements of a certain group algebra. Weyl’s formula gives an expression for the character xΛ of the finite-dimensional irreducible module with highest weight Λ as a quotient of two quite simple elementary alternating expressions in the exponentials.
Weyl derived his formula originally by using integration on compact groups. An elementary purely algebraic method for obtaining the result is due to Freudenthal and we shall follow this in our discussion. A preliminary result of Freudenthal’s gives a recursion formula for the multiplicities nM. Weyl’s formula can be used to derive a formula for the dimensionality of the irreducible module with highest weight Λ. It can also be used to obtain the irreducible constituents of the tensor product of two irreducible modules.
In this section we derive some properties of the Weyl group which are needed for the proof of Weyl’s formula and for the determination of the automorphisms of semi-simple Lie algebras over an algebraically closed field of characteristic 0 (Chapter IX). As usual, denotes a finite-dimensional split semi-simple Lie algebra over a field Φ of characteristic zero, ei, fi, hi, i = 1,2, …, l, are canonical generators for
as in Chapters IV and VII, and
is the splitting Cartan subalgebra spanned by the hi. Let
* be the conjugate space of
,
*0 the rational vector space spanned by the roots of
in
. If α is a non-zero root, then the Weyl reflection is the mapping
in *0. Here (ξ, η) is the positive definite scalar product in
*0 defined as in §4.1. The mapping Sα is characterized by the properties that it is a linear transformation in
*0 which maps a into —α and leaves fixed every vector orthogonal to α. The reflection Sα is an orthogonal transformation relative to (ξ, η) and Sα permutes the weights of
in any finite-dimensional module for
. The Sα generate the Weyl group W which is a finite group. If T is an orthogonal transformation such that αT is a root for some root α ≠ 0, then a direct calculation shows that
In particular, this holds for every T ∈ W. We note also that if α is a root then — α is a root and it is clear from (4) that Sα = S_α.
The roots αi, i = 1,2, …, l, such that αi(hj) = Aji, (Aij) the Cartan matrix, form a simple system. The notion of a simple system of roots is that given in § 4.3 and depends on a lexicographic ordering of the rational vector space *0. We recall that l roots form a simple system if and only if every root α = ±(Σkiαi) where the ki are non-negative integers. This criterion implies that if π = {α1, α2, …, αi} is a simple system and T is a linear transformation in
*0 which leaves the set of roots invariant, then πT = {α1T, …, αlT} is a simple system of roots. The set P of positive roots in the lexicographic ordering which gives rise to the simple system π is the set of non-zero roots of the form Σkiαi, ki
0. On the other hand, if the ordering is given, so that P is known, then π is the set of elements of P which cannot be written in the form β + γ, β, γ, ∈ P. Thus P and π determine each other. Any simple system of roots defines a set of canonical generators ei, fi, hi(cf. §4.3).
LEMMA 1. Let π = {α1, α2, …, αl} be a simple system of roots and let a be a positive (negative)root. Then αSαi > 0 (< 0) if α ≠ -α; and αSαi < 0 (> 0) if α = α (α = — α i).
Proof. If α is a negative root, then — α is a positive root and the assertion on α will follow from that on — α. Hence it suffices to assume α > 0. We have α = Σkjαj, kj 0 and
If α ≠ αi, then since no multiple except 0, ± α of a root is a root, α ≠ kiαi and so kj ≠ 0 for some j ≠ i. Then the expression for αSαi has a positive coefficient for some αj and hence αSαi > 0. If α = αi, then αSαi = - α i < 0. This completes the proof.
Our results on W will be given in two theorems. The first of these is
THEOREM 1. If π = {α1, α2, …, αi} is a simple system of roots, then the Weyl group W (relative to ) is generated by the reflections Sαi, αi ∈ π. If a is any non-zero root, then there exists α ∈ π and S ∈ W such that αjS = α.
Proof: Let W′ be the subgroup of W generated by the Sαi. We show first that any positive root α has the form αjS, αj ∈ π, S ∈ W. We recall that if α = Σkiαi then Σki is the level of α and we shall prove the result by induction on the level. The result is clear if the level is one since this condition is equivalent to α ∈ π. If Σki > 1, and there exists an αi such that (α, αi) > 0. Otherwise (α, αi)
0 for all i and since α = Σkiαi, ki
0 contrary to (α, α)
0. Choose αi so that (α, α) > 0. Then β = αSαi. > 0 by Lemma 1 and we have β = Σj≠ikkjαj + (ki – [2(α, αi)])αi. Since (α, αi) > 0 it is clear that the level of β is lower than that of α. Hence the induction shows that β = αjS′ where S ∈ W′, αj ∈ π, S′ ∈ W′. Then α = αjS′Sαi as required. Since α = αj S where S ∈ W′, Sα = S-1SαjS ∈ W′. Since Sα, = S-α this shows that every Sα ∈ W′ and so W′ = W. It remains to show that if α is any non-zero root then α = αiS, α ∈ π, S ∈ W. This has been shown for α > 0. Since αSα = — α the result is clear also for α < 0.
THEOREM 2. Let π and π′ be any two simple systems of roots. Then there exists one and only one element S ∈ W such that πS = π′.
Proof. Let P and P′, respectively, denote the sets of positive roots determined by π and π′. It is clear that P and P′ contain the same number q of elements, since this is half the number of non-zero roots. It is clear also that P = P′ if and only if π = π′ and if ≠ P′ then π P′ and π′
P. Let r be the number of elements in the intersection P
P′. If r = q, P = P′ and S = 1 satisfies πS = π′, so the result holds in this case. We now employ induction on q — r and we may assume r < q, or P ≠ P′. Then there exists αi ∈ π such that
and hence
. If
, by Lemma 1. Hence
. Also αi = (−αi)Sαi ∈ P
P′ Sαi. Hence P
P′ Sαi contains at least r + 1 elements. The simple system corresponding to P′Sαi is π′Sαi so the induction hypothesis permits us to conclude that there exists a T ∈ W such that πT = π ′Sαi. Then S = TSαi satisfies πS = π′. This proves the existence of S. To prove uniqueness it suffices to show that if S ∈ W satisfies πS = π, or equivalently, PS = P, then S=1. We give an elementary but somewhat long proof of this here and in an exercise (Exercise 2, below) we shall indicate a short proof which is based on the existence theorem for irreducible modules. We now write Si = Sαi and by Theorem 1, S = Si1, Si2 … Sim, ij = 1, 2, …, l. We cannot have m = 1 since αi1Si1 = - αi1
π and if m = 2, then αi1Si = — αi1 and since (— αi1)Si1 = - αi1Si1Si2 > 0, i2 = i1 by Lemma 1. Thus
. We now suppose m > 2 and we may assume that if T = Sj1Sj2 … Sjr, r < m, and πT = π, then T = 1. Let S′ = Si1Si2 … Sim-1, so that S = S′Sim. Since S ≠, Sim, S′ ≠ 1 and since S′ is a product of m — 1 Sj, PS′ ≠ P by the induction hypothesis. Then there exists an αj ∈ π such that αjS′ < 0. Since αjS′Sim > 0, Lemma 1 implies that αjS′ = -αim. Similarly, if β is any positive root such that βS′ < 0, then βS′Sim > 0 implies that βS′ = — αim = αjS′ hence β = αi. Thus S′ has the following two properties: αjS′ = — αim and βS′ > 0 for every positive β ≠αj. Set Si0 = 1. Then αjSi0 > 0 and
Hence there exists a k, 1 k
m — 1, such that αjSi0Si1 … Sik-1 > 0 αjSi0 … Sik-1 > 0 and (αjSi0 … Sik-1)Sik < 0, αjSi0 … Sik-1 = αik. Hence if we set T = Si0Si1 … Sik-1, then T-1SjT = Sik or SjT = TSik. Then, if T′ = Sik+1 … Sim-1 for k < m — 1 and T′ = 1 for k = m — 1, S′ = TSikT′ = SjTT′. Hence TT′ = SjS′. If α is a positive root ≠ αj, then β = αSj is a positive root ≠ αj. Hence αTT′ = αSjS′ = β S′ > 0. Also αjTT = αjSjS′ = (— αj)S′ = αim > 0. Hence πTT′ = π and since TT′ is a product of m — 1 Sj’s the induction hypothesis gives TT′ — 1. Then SjS′ = 1 and so S′ = Sj and S = Si1Sj. Hence S = 1 by the case m = 2 which we considered before.
Let be a finite-dimensional irreducible module for
with highest weight Λ = Λ(h) a dominant integral linear function on
. We know that
is a direct sum of weight spaces relative to
and that the weights are integral linear functions on
of the form Λ — Σkiαi, ki a non-negative integer. If M = M(h) is any integral linear function on
we define the multiplicity nMof M in
to be 0 if M is not a weight and otherwise, define nM = dim
M, where 9ftis the weight space in
of
corresponding to the weight M. We recall that for the highest weight Λ we have nΛ = 1. We shall now derive a recursion formula due to Freudenthal which expresses nM in terms of the nM, for M′ > M in the lexicographic ordering of
0* determined by the simple system of roots π = {α1, α2, …, αl}.
Let α be a non-zero root of and choose
so that (eα, e-α) = — 1 where (x, y) = tr ad x ad y, the Killing form on
. Then we know (§4.1) that [eαe-α] = hα where, in general for ρ ∈
*, hρ is the element of
such that (h, hρ) = ρ(h). As in §4.2, let
(α) be the subalgebra
+ Фeα + Фe-α. Then we know that
is completely reducible as
(α)-module (Theorem 4.1). We consider a particular decomposition of
as a direct sum of irreducible
(α)-submodules and we shall refer to these as the irreducible
(α)-constituents of
. Let
be one of these. Then we know that
has a basis (y0, y1, …, ym) such that
We know also that
where (ρ, σ) = σ(hσ) = σ (hρ), as in §4.1. Equations (7) imply that
We note also that is a direct sum of weight spaces, that the weights are M, M — α, M — 2α, …, M — mα and the weight spaces
M-pα =
M-pα are one-dimensional.
Let M be a weight of in
such that M + α is not a weight. Then the α-string of weights in
containing M is M, M — α, …, M — mα where m = 2(M, α)/(α, α). Let 0
p
m. Then M — pα is a weight and
M-Tα is a direct sum of the M — pα weight spaces of those irreducible
(α) -constituents which have this as weight. If 0
p
(M, α)/(α, α), then these are the irreducible
(α)-constituents having maximal weight M, M – α, …, M – pα. Let mj, 0
j
(M, α)/(a, a), denote the number of irreducible
(α)-constituents of highest weight M — jα. Then it is now clear that
so that,
If 0 j
p
(M,α)/(α, α), then the weight space corresponding to M — pα in an irreducible
(α) -constituent having highest weight M — jα is spanned by the vector yp-j in the notation of (7). The dimensionality of this module is 2(M—jα, a)/(α, a) + 1 = (m — 2j) + 1.
Hence we may replace i by p — j and m by m— 2j in (9) to obtain
We can use this to compute tr , the trace of the restriction to tr
. We remark that it is clear that
so that is invariant under
. By (12), the contribution to tr
of the mj irreducible
(α) -constituents with highest weight M — jα is
hence,
Since m/2 = (M, α)/(α, α). we have the formula
for 0 p
(M, α)/(α, α).
Next let (M, α)/(α, α)< p m. If we apply the Weyl reflection Sα to M — pα we obtain
since m = 2(M, α)/(α, α).We have
and we see that M — pα is a weight for the irreducible
(α)-constituents having the following highest weights:
and only these. The reasoning used to establish (13) now gives
for (M, α)/(α, α;) <p m. We recall that if M is a weight, the MSα is a weight and nM= nMSα (Theorem 4.1). Hence
. On the other hand,
. Hence
This and (14) imply that (13) is valid also for p > (M, α)/(α, α).We recall that M was any weight of in
such that M + α is not a weight. Hence M — pα can represent any weight of
in
. We now change our notation and write M for M — pα. If we recall that nM′ = 0 if M′ is not a weight, then we can re-write (13) as
for any weight M and any root α ≠ 0.
We consider next the Casimir element defined by the Killing form. This is where (ui), (ui) are dual bases for
relative to (x,y). We know that [aRΓ] = 0 for all a є
. Since R is absolutely irreducible it follows from Schur’s lemma that
,
(cf. Jacobson, Lectures in Abstract Algebra II, p. 276). We choose dual bases in the following way: (h1, …, hl), (h1, …, hl) are dual bases in
relative to (x,y). Then since (eα, e—α) = — 1 and eα is orthogonal to
and to every
the following are dual:
Then
If we take the trace of the induced mappings in M, M a weight, we obtain
Since hR is the scalar multiplication by M(h) in M we have
, so the first term on the right hand side of (18) is
. We proceed to show that
(M, M). Thus we write
. Then
and
and
.
Hence . It follows that
If we use this result and (16) in (18), we obtain
The terms nM(M, α) and nM(M, —α) cancel in this formula so we may replace by
. The resulting formula is valid also if M is not a weight. In this case nM=0. If, for a particular α,no M + jα, j
1, is a weight then nM + jα — 0 and the particular sum
. If M + jα is a weight for some j
1, then no M — kα is a weight for k
1 since the α-string containing M+jα does not contain M. Hence the set {M+kα | k
1}contains the complete α-string containing M + jα. Then
, since
Thus in all cases and (19) holds. The argument shows also that
holds for any integral linear function M on . This implies that
, so if we substitute in (19) with
we obtain
Setting this gives the following formula:
If M = Λ the highest weight of in
, then nM = 1 and nM+kα = 0 for α > 0, k
1. Hence (21) gives
and substitution in (21) gives Freudentha’s recursion formula:
We shall now show that this gives an effective procedure for calculating nM beginning with nΛ = 1. For this we require a couple of lemmas.
LEMMA 2. Let . Then δ(hi) = 1, i= 1,2, …,l, if ei, fi, hi is a set of canonical generators for
. Also if S ≠ 1 is in W, then δ — δS is a non-zero sum of distinct positive roots.
Proof: If α is a positive root we know (Lemma 1) that αSi ≡ αSαi = α-α(hi)αi > 0 unless α = αi in which case αiSi = —αi.Hence
Also, (δSi,αi) = (δ,αiSi) = (δ,- αi). Hence (δ- αi, α i)= (α,- αi) and 2(δ, αi) = (αi, αi). Thus δ(hi) = 2(δ, αi)/(αi, α i) = 1, i = 1,2,…,l. Since any S α W maps roots into roots we evidently have δS = δ-Σβ, where the summation is taken over the β = -αS > 0. If there are no such β, then αS > 0 for all α > 0. This implies S = 1, by Theorem 2, contrary to hypothesis.
LEMMA 3. Let Λ be the highest weight of in
. Then
for any weight M ≠ Λ of in
.
Proof: We shall prove that there exists a M' > M such that . This will prove the result by an evident ascending chain argument. Assume first there exists an i such that M(hi) < 0, hi in the set of canonical generators. Then we take M’ = MS = M-M(hi)αi > M and we have
since M(hi) = 2(M, αi)/(αi, α i). Since (α, αi) > 0 this shows that Next suppose M(hi)
0, i= 1, 2,..l. M +αi is a root and x ≠ 0 satisfies xh = Mx, then xei = 0 for all i and x is canonical generator of an e-extreme
-module. This must coincide with
, so M is the highest weight, contrary to hypothesis. Now let αi be one of the simple roots such that M’ = M+ αi is a weight. Then M' > M and
Since M(hi) 0, (M,α)
0. Also (δ,αi) > 0 and (αi, αi) >0, so again we have
It is now clear that (22) can be used to determine the weights and their multiplicities. Thus we begin with nA = 1. Suppose that for a given non-negative integers with at least one ki > 0, we already know nM' for every
integers, 0
k′i
k′i, M' ≠ M. Then every term on the right hand side of (22) is known. Moreover, if (Λ+α, Λ+α) = (M+α, M+α), then, by Lemma 3, M is not a weight and nM = 0. Otherwise, the coefficient of nM in (22) is not zero so we can solve for nM using the formula.
In order to obtain a general formulation of Weyl's formula valid for any field it is necessary to replace the exponentials which appear in this formula by “formal” exponentials. This notion can be made precise by introducing the group algebra over the base field of the group of integral linear functions on . We recall that an element M ∈
* is called integral if M(hi) is an integer for i = 1,2, …, l, where ei, fi, hi, are canonical generators for
. The set
- of integral linear functions is a group under addition, and it is clear that
is the direct sum of the cyclic groups generated by elements λi of
such that λi(hj) = δij. We now introduce an algebra
over α with basis
in 1:1 correspondence with the elements of
in which the multiplication table is
Then is the group algebra over Φ of
and e(0) = 1 is the identity element of
. The elements of
are the formal exponentials to which we alluded before. We now define the character X of a finite-dimensional module
to be the formal exponential
where nM is the multiplicity of M ∈ as defined before: nM = 0 if M is not a weight and nM = dim
M the dimensionality of the weight space if M is a weight. The summation in (25) is taken over all M ∈
. This is a finite sum since nM ≠ 0 for only a finite number of M.
Let xi = e(λi), λi (h j) = δij. Since any integral linear function can be written in one and only one way as M = Σmiλi where the mi are integers, the base element
. We have called M dominant if M(hi)
0 for i = 1, 2, …,l. This is equivalent to mi
0. The set of linear combinations of the e(M), M dominant, is a subalgebra of
which is the same as the set of linear combinations of the monomials
. The set of these monomials obtained from the sequences of integers (m1,…, ml) with mi
0 are linearly independent. Hence the subalgebra we have indicated can be identified with the commutative polynomial algebra Φ[x1, x2, …, xl] in the algebraically independent elements x1 …, xl. Every element of
has the form
lf where the ri
0 and f∈ Φ[x1 …, xl]. It is well known that Φ[x1 …, xl] is an integral domain. It follows that is a commutative integral domain.
With each element S in the Weyl group W we associate the linear mapping in such that e(M)S = e(MS). Since
it is clear that S is an automorphism in . The set of S in
is a group of automorphisms isomorphic to W. An element α ∈
is called symmetric if aS = a,S ∈ W, and alternating if aS = (det S)a, S ∈ W where det S is the determinant of the orthogonal transformation S in
. Thus det S = ± 1 and if S = Sα the Weyl reflection determined by the root α then det Sα =—1. In particular, det Si =- 1 for Si =Sαi. Since the Si generate W, a is symmetric if and only if a Si = a, i — 1, 2, …, l, and a is alternating if and only if aSi = a for all i. The set of symmetric elements is a subalgebra; the set of alternating elements is a subspace. The product of two alternating elements is symmetric and the product of an alternating element and a symmetric element is alternating. Since nMs = nM for S ∈ W, it follows that the character
is a symmetric element. We note next that the element
where δ = (1/2)(Σα > 0 α) the integral linear function defined in § 2, is an alternating element of . Thus we have seen (proof of Lemma 2) that δSi = δ – αi so (-δ)Si = -δ + αi and e(-δ)Si = e(-δ)e(αi). Also we know that Si permutes the positive roots ≠ αi and sends αi into —αi. Hence
and
We set
This element is alternating. It turns out that we can obtain a formula for f and this will give the desired formula for χ.
Let σ = ΣS∈W(det S)S as linear operator in . We have for any T ∈ W that σT = ΣS(det S)ST = (det T)-1 Σs(det ST)ST = (det T)σ and similarly Tσ = (det T)σ. If a is any element of
, then (aσ)T = (det T)(aσ). Hence aσ is an alternating element. If a is alternating to begin with, then aS = (det S)a and so aσ = wa where w is the order of the Weyl group. It follows that (1/w)σ is a projection operator of
onto the space of alternating elements. Hence any alternating element has the form aσ, a ∈
, and consequently such an element is a linear combination of the elements e(M)σ, M an integral linear function. Since Sσ = (MS)σ it is clear that e(M)σ can be replaced by e(M)σ and so we may express any alternating element as a linear combination of elements e(M)σ where M is highest among its conjugates MS under the Weyl group. In particular, we may suppose that M
MSi = M — M(hi)αi, that is, we may suppose that M(hi)
0 for i = 1,2,…, l. Suppose M(hi) = 0 for some i, or equivalently MSi = M. Then e(M)σ = — e(M)Siσ = -σ(MSi) σ= -e(M)σ and e(M)σ = 0. We therefore conclude that every alternating element is a linear combination of the elements e(M)σ where M(hi) > 0, i=1, 2, …, l.
We now apply this argument to the element Q defined in (26). If we multiply out the product in (26) we see that Q is a linear combination of elements e(M) where M = δ — ρ and ρ is a sum of a subset of positive roots. Thus where ∈α = ±1 and any conjugate MS of M under the Weyl group again has the form
where ρ′ is a sum of positive roots and ∈′α = ±1. If we apply the projection operator (1/w)σ to the expression indicated for Q we obtain an expression for Q as linear combination of elements e(M)σ where M is of the form δ – ρ, ρ a sum of positive roots. We have seen also that we may assume that M satisfies M(hi) > 0, i = 1,2, …, l. These conditions imply that M = δ. Thus, M = δ – ρ, ρ = Σkiαi where the ki are nonnegative integers, then
. Since
is an integer. Hence we must have ρ(hi)
0. On the other hand,
. Hence (ρ, ρ) =0 and ρ = 0. Thus we have shown that
. Now
. By Lamma 2,
. Hence
where M = δ - ρ, ρ ≠ 0 and a sum of positive roots. Also it is clear from (26) that Q = e(δ) + Σ ± e(M). Since Q = ηe(δ)σ a it follows that η = 1. We have therefore proved the following
LEMMA 4. Let Q = e(- –)IIα > 0(e(α) - 1). Then
We shall next introduce the vector space and we shall define certain linear mappings of this space and of the algebra
. The elements of
have the form
. The algebra composition in
provides a linear mapping of
⊗
into
such that a ⊗ b → ab. This gives a linear mapping of
so that
. It follows that if we set
then this product of Σρi ⊗ ai and b is single-valued and coincides with the image of Σρi ⊗ ai ⊗ b in
* ⊗
. It is clear that the product (Σρi ⊗ ai)b = Σρi ⊗ aib turns
*⊗
into a right
-module. We recall that we have the bilinear form (ρ, σ) = (hρ, hσ) on
* which defines the linear mapping
* ⊗
* into
so that ρ ⊗ π → (ρ, π). If we combine this with the linear mapping of
⊗
into
we obtain the linear mapping of (
* ⊗
) ⊗ (
* ⊗
) into
⊗
=
such that (ρ ⊗ a) ⊗ (τ ⊗ b) → (ρ, τ)ab. This defines a
-bilinear mapping of
* ⊗
such that the value (ρ ⊗ a, τ ⊗ b) = (ρ, τ)ab. Since (ρ, σ) is symmetric and
is commutative this is a symmetric bilinear form. Also if c ∈
, then
and (ρ ⊗ a, τ ⊗ b)c = (ρ, τ)abc and similarly
which implies that (x, y), x, y ∈
* ⊗
, is
-bilinear.
Next we define a linear mapping, the gradient, of into
* ⊗
: a → aG such that e(M)G = M ⊗ e(M) and a linear mapping, the Laplacian, of
into
: a →aα such that e(M)Δ = (M, M)e(M). We have
The linearity then implies that
We have
This implies that
We now return to the formulas which we developed for the multiplicity nM of the integral linear function M in a finite-dimensional irreducible module of highest weight Λ, γ the element of (rational number) determined by the Casimir operator. We consider again (19):
We multiply both sides by e(M) and sum on M. This gives
which we multiply through by
by (26). This gives
The coefficient of e(M + α) in the right-hand side of (31) is
Hence (31) can be written in the form
Hence we have
and canceling Q ≠ 0 in the integral domain , we obtain
If we set f = χQ, as before, we obtain
Since Q = ΣS∈Wdet S(e(δS)) and (δS, δS) = (δ, δ) we have QΔ = (δ, δ)Q. Since γ = (Λ + δ, Λ + δ) — (δ, δ) (eq. below (21)) these substitutions convert (33) to the following fundamental equation for f
The element f = χQ is an alternating element and consequently this element is a linear combination of elements of the form e(M)σ. Moreover, we can limit the M which are needed here by looking at the form of χ and Q. Thus we have χ = ΣnMe(M) where we now consider the summation as taken just over the weights M of the representation. Also we have seen that Q = ΣS∈W(det S)e(δS). If we multiply we obtain χQ as a linear combination of terms e(M + δS) where M is a weight: M a weight, S ∈ W. Now
so that e(M + δS) is in the characteristic space of the characteristic root (MS-1 + δ, MS-1+ δ) of Δ. Since f belongs to the characteristic root (Λ + δ, Λ + δ) for α and characteristic spaces belonging to distinct roots are linearly independent it follows that f is a linear combination of e(M + δS) such that
By Lemma 3 (MS-1 + δ, MS-1 + δ) = (Λ + δ, Λ + δ) for the weight MS′ implies that MS′ = Λ. Hence we see that f is a linear combination of the terms e(Λ + δ)S. If we apply the projection operator (1/w)σ to f we see that . Since
. Hence the coefficient of e(Λ + δ) in our expression for f is η. On the other hand, the coefficient of this term in χQ is nΛ = 1. Hence η = 1 and we have proved
Weyl’s Theorem. Let be the irreducible module for
with highest weight Λ. Then the character χΛ = ΣnMe(M) of
in
is given by the formula
where δ = (1/2) Σα>0α, α a root.
This theorem means that the expression on the right is divisible by Q = Σs (det S)e(δS) in and the quotient is the character χ of the representation.
It is easy to see that this result gives Weyl’s original formula
in the complex case. Weyl employed his result to obtain by a limiting process the dimensionality of = ΣnM = χΛ(0). We proceed to obtain the same result by a somewhat similar device.
We introduce the algebra 〈t〉 of formal power series in an indeterminate t with coefficients in
. We recall that the mapping of power series into their constant terms is a homomorphism ζ of
〈t〉 onto
. We can also define homomorphisms of
into
〈t〉 by employing exponentials: exp z = 1 + z +(z2/2!) + … which is defined for any z ∈
〈t〉 with zero constant term. We have the relation exp (z1 + z2) = (exp z1)(exp z2); hence, if λ, μ, ρ ∈
*, then exp (λ, ρ)t exp (μ, ρ)t = exp (λ + μ, ρ)t. In particular, this holds for λ = M, μ = M′, integral linear functions on
, which implies, in view of (24), that we have a homomorphism ζρ of
into
〈t〉 such that
. Now consider χΛ ζρ ζ. Since χΛ = ΣnMe(M) we have χΛ ζ ρ = ΣnMe xp (M, ρ)t and since the constant term of an exponential is 1, χΛ ζ ρζ = Σn M= dim
. We shall obtain the formula for dim
by applying ζρζ to (35), taking ρ = δ = (1/2) Σα > 0 α.
Let σ = ΣS∈W (det S)S as before and let M, M′ be integral linear functions. Then
Hence
since δ = (1/2)Σα > 0 α. Applying this to (35) we obtain
Now
where k is the number of positive roots. Hence if we divide both sides of (38) by tk and then apply the homomorphism ζ which picks out the constant term, we obtain
or
We now indicate how dim can be calculated from the Dynkin diagram for
. Let wi = 1, 2 or 3 be the weight of the vertex αi in the diagram and let αr be one of the roots in the simple system such that wr = 1. We replace the scalar product (α, β) by (α, β)′ = 2(α, β)/(αr, αr). Then we evidently have dim
. We write Λ = Σmiλi, mi
0, α = Σkiαi, kj
0. Then since
and δ(hj) = 1, j = 1, …, l, δ = Σλi Hence we require (Σmi + 1)λi, Σkjαj)′ = Σi,j(mi + 1)kj (λi, αj)′ and Σi,jkkj(λi, αj)′ Now
Hence and (δ, α)′ = Σwiki and
where the product is taken over all the sequences (k1, k2, …, kl ki 0, such that Σkjαj is a root. We have seen that this set as well as the wi can be determined from the Dynkin diagram.
G2. Here w1 = 3, w2 = 1 and the roots are α1, α2, α1 + α2, α1 + 2α2, α1 + 3α2, 2α1 + 3α2. Then (41) gives
if the highest weight of is m1λ 1+ m2λ2. For Λ = λ1 and Λ = λ2 we obtain, respectively, 14 and 7, which we had obtained previously.
Bl, l 2. Here w1 = w2 = … wl-1 = 2, wl = 1 and the weights are
These contribute the following factors to the dimensionality formula:
Let Λ = λk, 1 k
l – 1, so that mk = 1, mi = 0 if i ≠ k. The product taken over the first set of factors in (44) is
The second set of factors gives
and the last is
Multiplication of these results gives
We recall that this result was what was required to complete our proof (§ 7.6) of the irreducibility of the module of k-vectors relative to Bl.
We shall now call a character of a finite-dimensional irreducible representation of a primitive character. Such a character has the form
, where the summation is taken over M < Λ in the lexicographic ordering in
0* or in
. Since the e(Λ) constitute a basis for the group algebra
of
it is clear that χ Λ1= χ Λ2 implies Λ1 = Λ2 and this implies that the associated representations are equivalent. Conversely, equivalence of finite-dimensional irreducible representations implies equality of the characters. It is clear also from the expression we have indicated for a primitive character that distinct primitive characters are linearly independent. If
is any finite-dimensional module for
,
is a direct sum of, say, m1 irreducible modules with character χΛ1 = m2 with character χΛ2 ≠ χΛ1 etc. Then the character χ of
has the form
Since this expression is unique we see that if χ and the primitive characters are known then the mi can be determined. This gives the isomorphism classes and multiplicities of the irreducible constituents of . As a corollary, we see that these classes and multiplicities are independent of the particular decomposition of
into irreducible constituents.
The characters can be used to determine the isomorphism classes and multiplicities of the tensor product of two irreducible modules. Suppose (x1, …, xm) is a basis for a module such that xih = Λi((h)xi and (y1, …, yn) is a basis for a module
such that yjh = Mj(h)yj. Then the mn products xi ⊗ yj form a basis for
⊗
and we have
Hence we have the following expressions for the characters of ,
and
⊗
. Thus we see that the character of
⊗
is the product in
of the characters of
and
. One obtains the structure of
⊗
by writing the product of two characters as a sum of primitive characters. In the case of the Lie algebra A1 there is a classical formula for this called the Clebsch-Gordon formula. We shall give an extension of this which is due to R. Steinberg. This is based on an explicit formula for the multiplicity nM of the weight M in the irreducible module with highest weight Λ. This formula is due to Kostant; the simple proof we shall give is due to Cartier and to Steinberg (independently).
We introduce first a partition function P(M) for M ∈ , which is analogous to the (unordered) partition function of number theory. If M ∈
, P(M) is the number of ways of writing M as a sum of positive roots, that is, P(M) is the number of solution (kα, kβ,…,kρ), of Σα0kα& α = M where the kα are non-negative integers and {α, β, … ρ} is the set of positive roots. We have P(0) = 1 and P(M) = 0 unless M = Σmiαi, mi-non-negative integer and defined as before. Then
. It is convenient to replace the group algebra
by the field
of power series of the form
is an infinite series with coefficients in
in the elements
ni non-negative integral. In
we can consider the “generating function” ΣM∈
P(M)e(M) which is defined since P(M) = 0 unless M = Σmαi, mi
0. It is clear that we have the identity
Since (1 – e(α))-1 = 1 + e(α) + e(2α) + …, we have the identity
We re-write Weyl’s formula (35) as
We replace M by — M in this and multiply the result through by e(δ) to obtain
By Lemma 4,
Hence if we multiply both sides of (48) by , we obtain, by (47), that
Comparison of coefficients of e(— M) gives Kostant’s formula for the multiplicity nM in the irreducible module with highest weight Λ, namely,
We now consider the formula for χΛ1χΛ2 where is the character of the irreducible module
i with highest weight Λi. We have seen that χΛ1χ Λ2= ΣmΛχΛ, where mΛ is the multiplicity in
1 ⊗
2 of the irreducible module with highest weight Λ. The summation is taken over all the dominant Λ. If we multiply through by ΣS∈W (det S)e(δS) and apply Wey’s formula we get
The summation on the right-hand side can be taken for all if we define mΛ = 0 for non-dominant Λ. Applying the formula (50) for
we get
Hence
If we put on the left we get
Hence
It is easy to see that if Λ is dominant, then (Λ + δ)S and hence (Λ + δ)S - δ is not dominant if S ≠ 1. Hence if Λ is dominant then m(Λ+ δ)S- δ = 0 if S ≠ 1 and so we obtain the formula
for the multiplicity of the module with highest weight Λ in 1 ⊗
2
The notations and conventions are as in Chapters VII and VIII.
1. Let ρ ∈ 0*. Give a direct proof that ρ
ρS for every S ∈ W if and only if ρ(hi)
0, i = 1, 2, …, l.
2. (Seligman). Prove the uniqueness assertion in Theorem 2 by using the fact that there exists a finite-dimensional irreducible module with highest weight Λ satisfying: Λ(hi) are distinct and positive. Note that ΛS = Λ if πS = S, so that
. This leads to Λ(hi) = Λ(hj) if αiS = αj.
3. Call a weight Λ in a frontier weight if for every root α ≠ 0 either Λ + α or Λ - α is not a weight. Show that if
is finite-dimensional irreducible then any two frontier weights are conjugate under the Weyl group.
4. Let the base field be the field of real numbers. If a is a root let Pα be the hyperplane in defined by α(h) = 0. A chamber is defined to be a connected component (maximal connected subset) of the complement of ∪α≠0Pα, in
. Show that every chamber C is a convex set. A set of roots Σ is a defining system for C if C is the set of elements h satisfying α(h) > 0 for all α ∈ Σ. Defining systems which are minimal are called fundamental systems. Show that these are just simple systems of roots determined by the lexicographic orderings in
*.
5. Show that the group algebra (of the group
of integral functions on
) is a domain with unique factorization (into elements).
6. Prove that if P ∈ is alternating, then P is divisible by Q = ΣS(det S)e(δS).
7. Let η be the automorphism of such that e(Λ)η = e(— Λ). Show that if χ is the character of a finite-dimensional module
then χη is the character of the contragredient module
*.
8. Let be a finite-dimensional irreducible module whose character satisfies χη = χ. Assume the base field algebraically closed. Show that the image
R under the representation R in
is a subalgebra of an orthogonal or a symplectic Lie algebra of linear transformations in
.
9. Let be the split three-dimensional simple Lie algebra with canonical basis e, f, h. Show that the character of the (m + 1)-dimensional irreducible module
m+1 for
is xm + xm-1 + … + xm where x = e(λ), λ(h) = 1. Use this to obtain the irreducible constituents of
m-1 ⊗mn+1.
10 (Dynkin). Let and ⊗ be finite-dimensional irreducible modules for
. Prove that
⊗
is irreducible if and only if for every l in any simple ideal of
either
l = 0 or
l = 0.
11. Use Weyl’s formula to show that the dimensionalities of the four basic irreducible modules for F4 are: 26, 52, 273, 1274. Use Freudenthal’s formula to obtain the character of the 26-dimensional basic module.
12. Use Weyl’s formula to prove dim = 2l if has highest weight λl for Bl.
13 (Steinberg). Let be the finite-dimensional irreducible module with highest weight Λ. Show that M is a weight of
if Λ — MS is a sum of positive roots for every S ∈ W.
14 (Kostant). Prove the following recursion formula for the partition function P(M):