Arun Ram
Department of Mathematics and Statistics
University of Melbourne
Parkville, VIC 3010 Australia
aram@unimelb.edu.au
Last update: 13 August 2012
Abstract.
This is a typed version of I.G. Macdonald's lecture notes on Kac-Moody Lie algebras from 1983.
Introduction
Before we get down to serious business, let me begin by telling you a little in general terms about this subject, and what I propose to do
(and what I do not propose to do) is this course of lectures. Basically it is an outgrowth of the theory of finite-dimensional
complex simple (or semisimple) Lie algebras developed by Killing and E. Cartan nearly 100 years ago. If is a complex
finite-dimensional semisimple Lie algebra, one can associate canonically with (I will go into details later) a certain
matrix of integers,
called the Cartan matrix of , which determines up to isomorphism. This matrix
satisfies the following conditions:
All principal minors of are positive.
Conversely, any matrix of integers satisfying these two conditions is the Cartan matrix of some semisimple Lie algebra
, and one can write down generators and relations for which involve only the integers
. (SOMETHING GOES HERE)
(I will write them down explicitly a little later.)
In the late 1960's, V. Kac and R. Moody more or less simultaneously and independently, had the idea of starting with a "generalized Cartan matrix"
(satisfying but not ), writing down the
same set of generators and relations as in Serr's theorem, and looking at the resulting Lie algebra (which is now infinite-dimensional). These are
the so-called Kac-Moody Lie algebras, or Lie algebras defined by generalized Cartain matrices.
Of course, nearly always when one takes an attractive and elegant piece of mathematics, such as the classical theory of finite-dimensional semisimple
Lie algebras over , and starts tinkering with it by weakening the axioms, the result is usually of no interest at all.
The surprising thing in this case is that one does continue to get a coherent theory (which of course includes the classical theory as a special case)
in which all the main features of the classical theory have their counterparts: root system, Weyl group, representations and characters – culminating
in a generalization (due to V. Kac) of Weyl's character formula.
Moreover it has become clear during the last 10 years or so that these Kac-Moody Lie algebras impinge on many other areas of mathematics:
Number Theory
(Modular forms)
Combinatorics
(Partitions, Rogers-Ramanujan identities)
Topology
(Loop spaces and Loop groups)
Linear Algebra
(Representations of Quivers)
Singularities
Completely integrable systems
Mechanics and Particle Physics
My aim in this course of lectures is to cover the basic structure and representation theory of Kac-Moody Lie algebras, but not
any of the applications listed above. Also I don't propose to assume any particular knowledge of Lie algebras, so I will begin by briefly reviewing
some of the basic notions. We shall work always over a field of characteristic 0 (and fairly soon
will be , the field of complex numbers). A Lie algebra is then a vector space
over endowed with a bilinear multiplication
(Lie bracket)
satisfying
(1)
for all
(2)
for all (Jacobi identity).
By applying (1) to and using the bilinearity of the bracket we have
(1)
and conversely (1) (1) (take ).
Examples
any vector space over , define
for all . This is an abelian Lie algebra.
any associative –algebra, define
.
Check that (2) holds. We have a Lie algebra L().
any –algebra – i.e. is a –vector
space endowed with a bilinear multiplication
.
A derivation is a –linear mapping which satisfies
If are derivations, so is
(check this). Again the Jacobi identity is satisfied (the verification is the same as in Ex. 2), so we have a Lie algebra Der().
It follows from that
by induction on ; hence if is nilpotent
is well-defined (because the sum is finite) and is an automorphism of :
so that is a –algebra
homomorphism, hence an automorphism because
.
Let be a finite-dimensional –vector space, End().
Then L() (Ex. 2) is a Lie algebra denoted by .
If , so that
is the algebra of
matrices over , we write
in place of
.
is a subalgebra of , because
.
a (real or complex) Lie group, tangent space
to at the identity element
( or here).
inherits from the group a Lie algebra structure: roughly speaking, addition in
corresponds to multiplication in near the identity, and the bracket
ERROR HERE? to formation of the commutator
(for near ). Here of course is finite-dimensional:
. This is
the origin of the subject: Lie algebra linear approximation to at .
Basic concepts
Many notions for groups have counterparts for Lie algebras. This is hardly surprising, given the origin of the subject.
Let be a Lie algebra. If are vector subspaces
(or just subsets) of , let denote the
subspace of spanned by all
with . Observe that
(because
).
Subalgebra: a vector subspace of is a subalgebra if
(so that
is a Lie algebra in it's own right).
Ideal: a vector subspace of 𝔤 is an ideal of
𝔤 if [𝔤,𝔫]⊂𝔫 (normal subgroup).
Quotient algebra: Let 𝔫 be an ideal in 𝔤, and form the vector space quotient
𝔤/𝔫, whose elements are the cosets
x‾=x+𝔫. Define
[x‾,y‾]=[x,y]‾,
this does not depend on the choice of representations, and makes 𝔤/𝔫 into a Lie Algebra.
(G/N)
Homomorphism: a homomorphism from 𝔤 to 𝔥 is a k–linear map
f:𝔤→𝔥 such that
f⁡([x,y])=[f⁡(x),f⁡(y)]
for all x,y∈𝔤.
It's kernel𝔫=f-1⁡(0)
is an ideal in 𝔤, it's image𝔞=f⁡(𝔤) is a subalgebra of 𝔥,
and f induces an isomorphism 𝔤/𝔫→∼𝔞
If x,y∈𝔤 are such that
[x,y]=0, we say that x,ycommute. In particular, if any two elements of 𝔤 commute, ie if
[𝔤,𝔤]=0, we say that 𝔤 is
abelian (Ex. 1 above).
Centre of
𝔤=z={x∈𝔤:[x,𝔤]=0}.
It is an ideal in 𝔤.
Derived algebra𝔤′=D𝔤=[𝔤,𝔤]
consists of all linear combinations of brackets [x,y].
D𝔤 is an ideal in 𝔤 (by virtue of the Jacobi identity):
[[x,y],z]=-[[y,z],x]-[[z,x],y]∈D𝔤,
hence [D𝔤,𝔤]⊂D𝔤.
Moreover 𝔤/D𝔤 is abelian (and D𝔤 is the
smallest ideal with abelian quotient).
Derived series, upper and lower central series; nilpotent, solvable Lie algebras.
Adjoint representation
For each x∈𝔤 we define
ad(x):𝔤→𝔤 by
ad(x)y=[x,y].
Then
ad:𝔤→𝔤𝔩(𝔤)
is a homomorphism of Lie algebras, because for all x,y,z∈𝔤 we have
This is another equivalent form of the Jacobi identity.
The kernel of ad is the centre z of 𝔤.
Inner automorphisms
If x∈𝔤 is such that adx is nilpotent
((x)N=0for someN>0)
then we can form eadx (Ex. 3 above) which is an
automorphism of 𝔤. The subgroup Int(𝔤) of Aut(𝔤)
generated by these eadx is the group of
inner automorphisms of 𝔤. It is a normal subgroup of Aut(𝔤), because if
φ∈Aut(𝔤) we have
φ(adx)φ-1=adφ(x)
and therefore also
φ(eadx)φ-1=eadφx.
Representations
Let 𝔤 be a Lie algebra. A representationρ of 𝔤 on a
k–vector space V is by definition a Lie algebra homomorphism
ρ:𝔤→𝔤𝔩(V).
In other words, for each x∈𝔤 we have a linear transformation
ρ(x):V→V depending on linearity on
x:
ρ(αx+βy)=αρ(x)+βρ(y)(x,y∈𝔤;α,β∈k)
and satisfying
ρ([x,y])=ρ(x)ρ(y)-ρ(y)ρ(x).
An equivalent notion is that of a 𝔤–module, which is a vector space V on which
𝔤 acts linearly, i.e. we are given a bilinear mapping
(x,v)↦x·v:𝔤×V→V
satisfying
[x,y]·v=x·y·v-y·x·v(x,y∈𝔤;v∈V)
To connect the two notions, define x·v=ρ(x)v.
Usual notions of irreducibility, direct sums etc.
Universal enveloping algebra of a Lie algebra
If G is a group, a G–module (or representation of G) is the same thing as a
kG–module, where kG is the group algebra of G over k.
The analogue of this for Lie algebras is the universal enveloping algebraU(𝔤) of
a Lie algebra 𝔤, which may be defined as follows: for the tensor algebra of the vector space 𝔤
T(𝔤)=⊕n≥0Tn(𝔤)
where
T0(𝔤)=k,T1(𝔤)=𝔤,Tn(𝔤)=𝔤⊗…⊗𝔤(nfactors)forn≥2.
Let J𝔤 be the two-sided ideal of IS THIS A T? U(𝔤) generated by all
x⊗y-y⊗x-[x,y](x,y∈𝔤)
and define
U(𝔤)=T(𝔤)/J𝔤.
U(𝔤) is functional in 𝔤: if φ:𝔤→𝔥
is a homomorphism of Lie algebras, it induces
T(φ):T(𝔤)→T(𝔥),
and
T(φ)(x⊗y-y⊗x-[x,y]) φx⊗φy-φy⊗φx-[φx,φy]∈J𝔥
so that T(φ) maps J𝔤 into J𝔥
and hence induces
U(φ):U(𝔤)→(𝔥)
U is the left adjoint of the functor L (Ex. 2) from associative algebras to Lie algebras (over k): for each Lie algebra 𝔤
and each associative algebra A, there is a canonical bijection
which is a Lie algebra homomorphism 𝔤→L(A). Finally verify that the mappings
φ→φ#,θ→θb
are inverses of each other.
In particular, if
ρ:𝔤→𝔤𝔩(V)=L(End(V))
is a representation, we have
ρ#:U(𝔤)→End(V),
i.e. V is a U(𝔤)–module.
The Poincaré-Birkhoff-Witt Theorem
Recall that
U(𝔤)=T(𝔤)/J𝔤;T(𝔤)=⊕n≥0Tn(𝔤)
is a graded algebra, but J𝔤 is not a graded ideal, because the generators
x⊗y-y⊗x-[x,y]
are not homogeneous:
x⊗y-y⊗x∈T2(𝔤),[x,y]∈T1(𝔤).
So U(𝔤) is not a graded algebra; but it does carry a filtration, defined as follows: Let
Tn=⊕i=0nTi(𝔤)
Let
π:T(𝔤)→U(𝔤)
be the canonical homomorphism, and let
Un=π(Tn)
The Un are vector subspaces of U:
k=U0⊂U1⊂…;U=⋃n≥0Un
and since
Tm⊗Tn⊂Tm+n
we have
Um·Un⊂Um+n,
ie U(𝔤) is a strong filtered associative k–algebra. Now form the associated graded algebra:
define
Gn=Un/Un-1(n≥0;U-1=0)
then the multiplication in U(𝔤) induces bilinear mappings
Gm×Gn→Gm+n
namely (for x∈Um,y∈Un )
(x+Um-1)(y+Un-1)=xy+Um+n-1
which make
G=Gr(U(𝔤))=⊕n≥0Gn
into a graded associated k–algebra. For each n≥1 we have
commutative with exact rows;
φ:Tn(𝔤)→Gn
is surjective, hence we have a surjective algebra homomorphism
φ:T(𝔤)→G
defined by
φ(x)=π(x)+Un-1(x∈Tn)
In particular, if x,y∈𝔤 we have
π(x⊗y-y⊗x)=π([x,y])∈U1
and therefore
φ(x⊗y-y⊗x)=0inG2
Hence the kernel of φ contains the two-sides ideal I of
T(𝔤) generated by all
x⊗y-y⊗x(x,y∈𝔤);
now T(𝔤)/I is by definition the symmetric algebraS=S(𝔤) of the vector space 𝔤; hence
φ induces a surjective homomorphism
ω:S(𝔤)→Gr(U(𝔤))
(P – B – W) ω is an isomorphism.
For the proof, see the standard texts (Bourbaki (Ch. I), Humphreys, Jacoborn)
Let
σ:T(𝔤)→S(𝔤)
be the canonical homomorphism.
Let V be a vector subspace of
Tn(𝔤)
which is mapped isomorphically by σ onto
Sn(𝔤).
Then π(V)(≅V)
is a complement of Un-1 in
Un.
Proof.
The diagram
commutes, hence θ maps V isomorphically onto
Gn=Un/Un-1
(because ω is an isomorphism, by P-B-W). Hence the result.
□
The canonical map
g↪T(𝔤)→πU(𝔤)
is injective.
We may therefore identify𝔤 with it's image in U(𝔤).
Let
(xλ)λ∈L
be a totally ordered k–basis of 𝔤. Then the elements
xλ1…xλn=π(xλ1⊗…⊗xλn)
such that
λ1≤…≤λn(for alln≥0)
form a k–basis of U(𝔤).
Proof.
Let Vn be the subspace of
Vn(𝔤) spanned by all
xλ1⊗…⊗xλn
with
λ1≤…≤λn.
Clearly σ maps Vn isomorphically onto
Sn(𝔤), hence by Corollary 7.2
π(Vn) is a complement of
Un-1 in Un.
By induction on n it follows that U(𝔤)
is the direct sum of the π(Vn) for all
n≥0.
(Corollary 7.4 is also known as the P-B-W theorem).
□
Recap
To recapitulate from last time:– to each Lie algebra 𝔤 we associate U(𝔤),
its universal enveloping algebra:
U(𝔤)=T(𝔤)/J𝔤
where T(𝔤) generated by all
x⊗y-y⊗x-[x,y](x,y∈𝔤).
𝔤embeds in U(𝔤)
(by virtue of the P-B-W theorem) and we identify 𝔤 with its image in U(𝔤).
In U(𝔤) we have
[x,y]=xy-yx(x,y∈𝔤).
Moreover U(𝔤) is universal in the following sense:
if φ:𝔤→A is any k–linear mapping of 𝔤
into an associative k–algebra A such that
φ[x,y]=φ(x)φ(y)-φ(y)φ(x)
– i.e. if
φ:𝔤→L(A)
is a Lie algebra homomorphism, then φ extends uniquely to a homomorphism
φ#:U(𝔤)→A,
as follows :– first extend φ to
φ∼:T(𝔤)→A
in the obvious way:
φ∼(x1⊗…⊗xn)=φ(x1)…φ(xn)(x1,…,xn∈𝔤)
and then observe that
J𝔤⊂Kerφ∼,
so that φ∼ induces
φ#:
U(𝔤)→A
as desired. Thus φ↦φ# is a mapping
which one easily verifies to be bijective (i.e. U is a left adjoint of the functor L, as I said last time).
Recall also (Corollary 7.4 of P-B-W th.) that if
(xλ)λ∈L
is an ordered k–basis of 𝔤, then the monomials
xλ1…xλn
with
λ1≤…≤λn
and n≥0
(if n=0, the product is empy and conventionally is said to be read as 1, the identity element
of U(𝔤)) form a k–basis of
U(𝔤). This has the following consequence: if
𝔤=𝔞⊕𝔟
where 𝔞,𝔟 are subalgebras and 𝔤 is the direct
sum of the vector spaces 𝔞,𝔟, then
U(𝔞),U(𝔟)
are subalgebras of U(𝔤) and
U(𝔤)=U(𝔞)U(𝔟)=U(𝔟)U(𝔞)
(Take ordered bases
(yμ),(zν)
of 𝔞,𝔟 respectively; the monomials
yμ1…yμm
with μ1≤…≤μm
form a k–basis of U(𝔞), the monomials
zν1…zνn
with ν1≤…≤νn
form a k–basis of U(𝔟), and the monomials
yμ1…yμmzν1…zνn
with μ1≤…≤μm and
ν1≤…≤νn
form a k–basis of U(𝔤).)
Free Lie algebras
Let X be a set, k a field. We want to define the free Lie algebra
Lie(X) on the set X. There are two ways of proceeding: one involves P-B-W, the other doesn't.
(1) Form the free non-associative algebra F(X) on X. How does one do this?
Define inductively sets Xn,n≥1 by
X1=XX2=X1×X1X3=(X2×X1)⊔(X1×X2)and in generalXn=∐p=1n-1Xp×Xn-p
and put
M(X)=∐n≥1Xn
(disjoint union) (the "free magma" on X).
If a,b∈M(X),
say a∈Xp and
b∈Xq, then
(a,b)∈Xp×Xq⊂Xp+q⊂M
so we have a multiplication ab=(a,b)
in M(X). Then F(X)
is the k–algebra with M(X) as basis,
i.e. it consists of all finite linear combinations
∑λiai with
λi∈k and
ai∈M(X),
and multiplication defined in the obvious way:
(∑λiai)(∑μjbj)=∑i,jλiμjaibj.
Now let J be the 2-sided ideal in F(X)
generated by all
xx,x(yz)+y(zx)+z(xy)(x,y,z∈F(X))
and define
Lie(X)=F(X)/J
where F(X) is a graded algebra and J
is a homogeneous ideal. It is clear that Lie(X) is a Lie algebra and that if
j:X↪F(X)→Lie(X)
is the canonical embedding, then any mapping φ of the setX
into a Lie algebra 𝔤 extends uniquely to a Lie algebra homomorphism
φ#:Lie(X)→𝔤
(extend φ in the obvious way to
φ∼:F(X)→𝔤
and observe that the generators of J in the kernel of
φ∼, by definition).
i.e. the functor Lie (from sets to Lie algebras) is a left adjoint of the forgetful functor Φ
(from Lie algebras to sets).
(2) Let A(X) be the free associative algebra
on X
(=F(X)/I, where I
is the 2-sided ideal generated by all
(xy)z-x(yz)(x,y,z∈F(X))).
The embedding
X↪A(X)=L(A(X))
induces, as we have just seen, a Lie algebra homomorphism
α:Lie(X)→L(A(X))
hence also
β:U(Lie(X))→A(X)
But also we have a mapping
X↪Lie(X)→U(Lie(X)),
hence (by the universal property of A(X)) a homomorphism
of associative algebras
γ:A(X)→U(Lie(X))
Check that these two homomorphisms β,γ are inverses of each other,
hence that
U(Lie(X))≅A(X)
By P-B-W, Lie(X) embeds in
U(Lie(X)),
hence in A(X), so that the mapping α
above is injective. In other words, the free Lie algebra Lie(X) may be described as
the subalgebra ofL(A(X))generated by X.
We have βk=alpha and γα=k,
also αj=i, hence
γβk=γα=k, hence
γβ=1 (because k injective);
βγi=βγαj=βkj=αj=i,
hence βγ=1 (because i is injective).
Finally, if R is any subject of Lie(X), the Lie algebra generated by
X subject to the relations R is by definition
Lie(X)/𝔞, where 𝔞
is the ideal of the Lie(X) generated by R (i.e. the intersection of all ideals
of Lie(X) which contain R).
Finite-dimensional simple Lie algebras / ℂ
This is the classical theory we intend to generalise. A Lie algebra 𝔤 is said to be simple
if its only ideals are 0 and 𝔤, and if also 𝔤 is non-abelian (thus deliberately
excluding the 1-dimensional abelian Lie algebra). Take k=ℂ, and dim
𝔤<∞.
An element x∈𝔤 is semisimple if
ad x:𝔤→𝔤 is a semisimple
(i.e., diagonalizable) linear transformation.
Let 𝔤 be simple, finite-dimensional. Then 𝔤 has nonzero subalgebras consisting
of semisimple elements (toral subalgebras); they are necessarily abelian.
Let 𝔥 be a maximal toral subalgebra (or Cartan subalgebra) of
𝔤. Certainly such exist, for dimensional reasons. Moreover (a non-trivial fact) any two such are
conjugate in 𝔤 (i.e. transforms of each other under the group Int(𝔤)
of inner automorphisms). Fix 𝔥 once for all.
l=dim𝔥 is called
the rank of 𝔤.
Let 𝔥* be the vector space dual of 𝔥. Introduce
the killing form
〈x,y〉=trace(adx)(ady)(x,y∈𝔤)
This is symmetric, nondegenerate and invariant, ie
〈[x,z],y〉=〈x,[z,y]〉(x,y,z∈𝔤)
Moreover its restriction to 𝔥 is nondegenerate, hence defines an isomorphism
ω:𝔥→∼𝔥*(ω(x)(y)=〈x,y〉)
and a symmetric bilinear form
〈λ,μ〉 on
𝔥*(〈λ,μ〉=〈ω-1λ,ω-1μ〉).
Example𝔤=𝔰𝔩n(ℂ)=
Lie algebras of n×n matrices with trace 0. Here we may take
𝔥 to consist of the diagonal matrices
k=(h1⋱hn)with∑1nhi=0
(so that l=rank(g)=n-1).
Roots
Consider the adjoint representation ad𝔤 of 𝔤,
restricted to 𝔥: this is a representation of 𝔥 on 𝔤.
Since 𝔥 is abelian, all its irreducible representations are 1-dimensional, so that 𝔤
splits up into a direct sum of 1-dimensional 𝔥–modules. explicitly, for each
α∈𝔥* define
𝔤α={x∈𝔤:(adx)=α(h)xfor allh∈𝔥}
Then it turns out that 𝔤0=𝔥; the nonzero
α∈𝔥* such that
𝔤α≠0 are called the roots of
𝔤 (relative to 𝔥). They form a finite subject R
of 𝔥*, called the root system of
(𝔤,𝔥), and we have
𝔤=𝔥+∑α∈R𝔤α(direct sum)
Moreover each 𝔤α(α∈R)
is 1-dimensional, and
[𝔤α,𝔤β]⊂𝔤α+β
(hence is 0 if α+β∉R⋃{0}).
If α is a root, so is -α.
In the case of 𝔰𝔩n(ℂ), let
eij(1≤i,j≤n)
be the matrix units, and for 1≤i≤n let
ui:𝔥→ℂ be the
ith projection:
ui(h)=hi.
Then
which shows that the roots are
α=ui-uj(i≠j);𝔤α=ℂeij,
and the root space decomposition is clear. We compute the Killing form on 𝔥 as follows from above
(adh)(adh′)eij=(hi-hj)(h′i-h′j)eij
so that
〈h,h′〉=∑i,j(hi-hj)(h′i-h′j)=2n∑1nhih′i
(remember that
∑hi∑h′i=0).
So it is a multiple of the obvious scalar product.
It is possible to choose roots
α1,…,αl(l=dim𝔥)
such that each root α∈R is of the form
α=∑1lniαi
with coefficients ni∈ℤ and either
ni≥0 (positive roots) or all
ni≤0 (negative roots). The
αi are called a set of simple roots or a
basisB of R (they are also a basis of
𝔥*). Choose such a basis once for all. There is then a unique
highest root, for which ∑ni is a maximum;
and a unique lowest root, for which ∑ni is
a minimum.
Weyl group
For each α∈R, let wα
denote the reflection in the hyperplane orthogonal to α in
𝔥*, so that
wα(λ)=λ-〈λ,α∨〉α(λ∈𝔥*)
where
α∨=2α/||α||2
is the coroot of α. The reflections
wαi corresponding to the simple roots generate
a finite group of isometries of 𝔥*, called the
Weyl groupW of (𝔤,𝔥).
Each reflection wα lies in W; R
is stable under W; and each root α∈R is of the form
wαi for some w∈W
and some simple root αi. Moreover, any other basis of R
is of the form
(wα1,…,wαl)=wB
for some (unique) w∈W.
Cartan matrix
The numbers
αij=〈αi∨,αj〉=2〈αi,αj〉〈αi,αj〉
are integers, and the matrix
(aij)1≤i,j≤l
is called the Cartan matrix of. It is independent of the choices of
𝔥 and of basis of R. It satisfies the following conditions:
(C)
aii=2(1≤i≤l);aij≤0ifi≠j;aij=0⇔aji=0
(P)
All principal minors of A are positive.
In the case of 𝔰𝔩n(ℂ)
we may take
αi=ui-ui+1(1≤i≤n-1).
The Weyl group W is the symmetric group Sn
(for wαi interchanges ui
and ui+1 and leaves the other
uj fixed). Here the Cartan matrix is
A=2-1-12-1-12⋱-1-12
(with l=n-1 rows and columns).
Generators and relations
The Cartan matrix A determines 𝔤 up to isomorphism.
Choose generators
ei∈𝔤αi,fi∈𝔤-αi(1≤i≤l)
such that
〈ei,fi〉=1,
and elements
h1,…,hl∈𝔥
such that
〈hi,h〉=αi∨(h),
so that
αj(hi)=〈αi∨,αj〉=aij
Then the 3l elements
ei,fi,hi
generate 𝔤 subject to the following relations (Serre):
The idea now is (roughly) the following: start with any matrix A of integers satisfying
(C), and form the Lie algebra with the above generators and relations.
However there is one remark that should be made at this point. In the classical set up (which I have just been describing) the
Cartan matrix A is nonsingular, the
hi(1≤i≤l)
form a basis of the Cartan subalgebra 𝔥, and the simple roots
αj∈𝔥*. Now a generalized
Cartan matrix may well be singular (and it would be foolish to exclude this possibility, because for the affine Lie algebras
the Cartan matrix is singular).
References
I.G. Macdonald
Issac Newton Institute for the Mathematical Sciences
20 Clarkson Road
Cambridge CB3 OEH U.K.