Affine Kac-Moody Algebras Graded by Affine Root Systems: Josiane Nervi
Affine Kac-Moody Algebras Graded by Affine Root Systems: Josiane Nervi
www.academicpress.com
0. Introduction
The aim of this paper is to give the complete classification of all affine Kac–
Moody algebras graded by affine root systems.
An affine Lie algebra ĝ is said to be graded by the affine root system Σ if ĝ
contains an affine subalgebra a whose root system relative to a Cartan subalgebra
ha is equal to Σ and such that ĝ = λ∈Σ∪{0} Vλ where Vλ = {x ∈ ĝ: [h, x] =
λ(h)x for all h ∈ ha }.
The Lie algebra a is then called the grading subalgebra. This is slight
modification of the notion of grading by a finite root system which was introduced
by Berman and Moody [1] and developed by Benkart and Zelmanov [2],
Neher [3], Allison et al. [4]. In the preceding paper [5] we gave the complete
classification of all simple finite-dimensional Lie algebras graded by a finite root
system.
Let us give now a description of the contents of this paper.
In Section 1 we construct a large class of affine subalgebras of a given affine
algebra ĝ, the so-called admissible subalgebras. In the finite-dimensional case, the
admissible subalgebras were constructed by H. Rubenthaler [6] and they played
a crucial role in the classification of dual pairs in reductive Lie algebras [7]. These
algebras are also important for the classification of the simple finite-dimensional
Lie algebras graded by finite root systems [5].
0021-8693/02/$ – see front matter 2002 Elsevier Science (USA). All rights reserved.
PII: S 0 0 2 1 - 8 6 9 3 ( 0 2 ) 0 0 0 5 3 - 4
J. Nervi / Journal of Algebra 253 (2002) 50–99 51
Let us recall here some basic facts from the theory of parabolic prehomoge-
neous vector spaces (details can be found in [8,9]). If G is an algebraic group
over C and (ρ, V ) a rational representation of G in a finite-dimensional vector
space, the triple (G, ρ, V ), or briefly (G, V ), is called a prehomogeneous vec-
tor space if G has a Zariski open orbit in V . This is, of course, an infinitesimal
condition; more precisely if g is the Lie algebra of G and if (g, dρ, V ) is the
corresponding derived representation of g, the condition of prehomogeneity is
equivalent to the existence of an element v0 ∈ V such that the map
g −→ V
X −→ dρ(X) v0
is surjective.
For that reason, a finite-dimensional representation (g, V ) verifying the above
mentioned surjectivity condition will also be called a prehomogeneous vector
space. A prehomogeneous vector space (G, V ), where the group G is reductive
is called regular if the isotropy subgroup of the open orbit is reductive. This is
known to be equivalent to the fact that the complementary set of the open orbit is
an hypersurface.
Let now g = gk be a Z-gradation of a finite-dimensional reductive Lie
algebra. By a result of Vinberg [10], the representation (g0 , gi ) (i = 0), is always
a prehomogeneous vector space.
Let now g be a simple Lie algebra. Let h be a Cartan subalgebra of g and let R
be the root system of the pair (g, h). We choose a basis Ψ of R and denote by θ
a subset of Ψ . Let Hθ the unique element defined by the equations:
α(Hθ ) = 0 ∀α ∈ θ, α(Hθ ) = 2 ∀α ∈ Ψ \θ.
Let dp (θ ) = {X ∈ g: [Hθ , X] = 2pX}. Then g = p∈Z dp (θ ) is a Z-gradation
of g and the prehomogeneous vector spaces (d0 (θ ), dp (θ )) are said to be of
parabolic type. Moreover, the representation (d0 (θ ), d1 (θ )) is irreducible if and
only if |Ψ \θ | = 1 (see, for example, [9]). The following theorem will be used
later.
Theorem 1.1.1 [9]. Suppose that the prehomogeneous vector space of parabolic
type (d0 (θ ), d1 (θ )) is irreducible. Then it is regular if and only if there exist
Y ∈ d−1 (θ ) and X ∈ d1 (θ ) such that (X, Hθ , Y ) is a sl2 -triple.
If [d1 (θ ), d1 (θ )] = {0} then we will say that the prehomogeneous vector space
of parabolic type is commutative (and this is equivalent to say that dp (θ ) = {0}
for any p 2).
J. Nervi / Journal of Algebra 253 (2002) 50–99 53
Let us first recall some facts about affine algebras and fix some notations
(see [11] for details).
An affine generalized Cartan matrix A = (aij )i,j =0,...,k indecomposable matrix
of integers satisfying: aii = 2, aij 0 if i = j ; aij = 0 implies aij = 0.
Furthermore all the proper principal minors of A are positive and det A = 0.
A realization of the matrix A is a triple (h, Π, Π ∨ ), unique up to isomorphism,
where h is a complex vector space, Π = {α0 , . . . , αn } ⊂ h , and Π ∨ =
{α0∨ , . . . , αn∨ } ⊂ h satisfying the conditions:
us put α ∨ = w(αi∨ ). One has α, α ∨ = 2 and there exist Xα ∈ ĝα and Yα ∈ ĝ−α
such that (Xα , α ∨ , Yα ) is a sl2 -triple. For a real root α, we also have dim ĝα = 1
and ∆ ∩ Zα = {α, −α}.
Let us denote by δ = ni=0 ai αi the unique element of h such that
At (a0 , . . . , an ) = 0 and ai are positive relatively prime integers; then ∆Im = {j δ:
j ∈ Z }. The imaginary root spaces are finite-dimensional too [11, Corollary 8.3].
Denote by g̊ the underlying simple subalgebra of ĝ generated by the triples
(ei , αi∨ , fi ), i = 1, . . . , n and by h̊ = ni=1 Cαi∨ the Cartan subalgebra of g̊. One
has h = h̊⊕CK ⊕Cd and the root system ∆ = (ĝ, h) can be completely described
in terms of δ and the root system ∆˚ of the pair (g̊, h̊) [11, Proposition 6.3]. The
set Π̊ = {α1 , . . . , αn } is a basis of ∆. ˚ The Cartan matrix of finite type Å of Π̊
is obtained from the matrix A by deleting the 0th row and column. The Dynkin
diagram S(Å) is obtained by deleting the 0th vertex on the Dynkin diagram, S(A),
of the matrix A.
Proposition 1.3.3.
• If α0 ∈
/ Γ , one has
h ∈ hΓ ⇔ α(h) = α(H ) = 0 ∀α ∈ Γ.
• If α0 ∈ Γ , one has
α(h) = α(H ) = 0 ∀α ∈ Γ \{α0 },
h ∈ hΓ ⇔ ✷
µ = −α0 (H ).
J. Nervi / Journal of Algebra 253 (2002) 50–99 55
For an element α = αi ∈Π nαi αi in the root lattice Q, the integer htΓ (α) =
αi ∈Π\Γ nαi is called the Γ -height of α.
For any integer p we set
α∈Ωp ĝ
α if Ω = ∅,
p
Ωp = α ∈ ∆ ∪ {0}: htΓ (α) = p , dp =
{0} if Ωp = ∅.
Proof. Using Remark 1.2.4(2) and since dim ĝα < ∞ for any α ∈ ∆ ∪ {0}, it is
sufficient to prove that |Ωp ∩ ∆Re | < ∞. Let α be a real root in Ωp . Suppose
(2)
that ĝ is not of type A2l , then α = j δ + γ , j ∈ Z, γ ∈ ∆˚ [11, Proposition 6.3].
Hence we have htΓ (α) = j htΓ (δ) + htΓ (γ ) = p. Since the set {htΓ (γ ), γ ∈ ∆} ˚ is
(2)
finite, the result follows. If ĝ is of type A2l one can also have α = 2 (2j − 1)δ + γ ,
1
In this section, we will decompose the d0 -module d1 (under the adjoint action)
into a direct sum of irreducible d0 -modules.
Let us make the convention which consists of circling the vertices ω0 , . . . , ωk
of Π\Γ in the Dynkin diagram S(A). We denote by Πj the connected component
of Γ ∪ {ωj } which contains ωj . As |Π\Γ | 2, the set Πj , can be identified with
a proper subdiagram of S(A); hence it is of finite type. Denote by ∆j = Πj the
set of roots in ∆ which are linear combinations of elements in Πj . Let gj be the
associated
simple algebra which can be viewed as a subalgebra of ĝ. The algebra
hj = α∈Πj Cα ∨ is a Cartan subalgebra of gj and ∆j is the set of roots of the
pair (gj , hj ).
56 J. Nervi / Journal of Algebra 253 (2002) 50–99
(1)
The algebra ĝ is of type B14 . The connected components Πj are of type D5 ,
A8 , and B9 respectively:
Proposition 1.3.8.
k j
• We have d1 = j =0 d1 .
j
• For any j , the space d1 is an irreducible d0 -module for the adjoint action.
Proof. Recall that d1 = α∈Ω1 ĝα where Ω1 = {α ∈ ∆: htΓ (α) = 1}. We also
j j
have d1 = α∈Ω j ĝα where Ω1 = {α ∈ ∆j : α = ωj (mod Γj )}. Since the
1
j
support of a root is a connected subgraph of S(A), one has Ω1 = Ω1 ∩ ∆j and Ω1
j j j
is the union of the disjoint sets Ω1 . Therefore d1 = kj =0 d1 . The space d1 is an
j j j
irreducible d0 -module which is d0 -stable. As d0 ⊂ d0 , the space d1 is also an
irreducible d0 -module. ✷
Proposition 1.4.1.2.
j j
(i) For any j = 0, . . . , k, we have Hj = α∈Πj xα α ∨ where each xα is a positive
integer.
(ii) The elements H0 , . . . , Hk are linearly independent in hΓ .
Proof. Let us fix j in {0, . . . , k}; let Mj denote the Cartan matrix of finite type
j
associated to the quadruple (gj , hj , Πj , Πj∨ ). Let us set Hj = α∈Πj xα α ∨ . Since
Hj verifies
α(Hj ) = 0 ∀α ∈ Γj , ωj (Hj ) = 2,
j
the column X = t((xα )α∈Πj ) is the solution of the linear system tMj X = B
where B is a column whose coefficients are all zero except one (say the qth
one) which is equal to 2. Using [11, Theorem 4.3] we deduce that X > 0 (i.e.
xα > 0 ∀α ∈ Πj ) and we have X = (tMj )−1 B; more precisely X = 2Λq where
j
Λq is the qth column of the matrix (tMj )−1 . These matrices (tMj )−1 are well
known (they give the fundamental weights in terms of the simple roots and we
find these relations in tables in [12]). It is easy to check that for every regular
j j j
space (d0 , d1 ) we have the expected result, namely that each xα is a positive
integer. More precisely:
j
Let us prove now that Hj = xα α ∨ lies in hΓ . Since for any γ ∈ Γ \Γj
α∈Πj
j
and any α ∈ Πj one has (γ , α) = 0, then γ (Hj ) = α∈Πj xα γ (α ∨ ) = 0. For
any γ ∈ Γj , one also has α(Hj ) = 0 by definition of the element Hj . Therefore
Hj ∈ hΓ .
Let us set now H = kj =0 aj Hj = 0 for some (a0 , . . . , ak ). For any j , the
j j
coefficient of ωj∨ in H is aj xωj (where xωj is the coefficient of ωj∨ in Hj ). Since
j j
the elements of Π ∨ are linearly independent, we have aj xωj = 0. Since xωj > 0,
we have aj = 0 for any j , and therefore the elements H0 , . . . , Hk are linearly
independent. ✷
Proposition 1.4.1.4. The matrix M = (ωj (Hi ))ij is a generalized Cartan matrix.
Proof. For any (i, j ), put mij = ωj (Hi ). Then for all j we have mjj = ωj (Hj ) =
2. If i = j , we have
i ∨
mij = ωj xγ γ = xγi ωj (γ ∨ ). ()
γ ∈Πi γ ∈Πi
On the other hand, the constants in the preceding sum are positive integers
(Proposition 1.4.1.2) and the constants ωj , γ ∨ are non-positive integers (since
ωj ∈
/ Πi ). Therefore the mij are non-positive integers.
J. Nervi / Journal of Algebra 253 (2002) 50–99 59
Suppose now that mij = 0 for some i = j . For all γ ∈ Πi , one has
ωj , γ ∨ 0. As xγi > 0 for all γ ∈ Πi , the equality () above implies that
ωj , γ ∨ = 0 for all γ ∈ Πi . Then Πj ∩ Πi = ∅ and furthermore (ωi , ωj ) = 0.
Hence ωi , γ ∨ = 0 for all γ ∈ Πj , therefore mj i = 0. ✷
Proof of Theorem 1.4.1.5. The matrix M is not invertible. In fact, the non-
zero canonical central element K which is obviously in h ∩ hΓ , can be written
K = x0 H0 + · · · + xn Hn (Lemma 1.4.1.6) and one has M t(x0 , . . . , xn ) = 0.
The matrix M is indecomposable. In fact, recall that for any j = 0, . . . , k the
coefficients of the element Hj in the basis Πj∨ are positive (Proposition 1.4.1.2).
As S(A) is a connected graph, one deduces that it is impossible to decompose the
set {0, . . . , k} into I1 ∪ I2 with ωj (Hi ) = 0 for all j ∈ I1 and i ∈ I2 .
Let M be a proper principal submatrix of M associated to a proper subset I of
{0, . . . , k}. Then M corresponds to the C-admissible diagram of semisimple type
constituted by the union of Πj for j ∈ I . Using [6, Theorem 3.1], we deduce that
det M > 0. We conclude by applying [11, Proposition 4.7.b]. ✷
For any j , the numerical labels in the following diagrams are coefficients of Hj
in the coroot basis Πj∨ .
The connected component Π0 is of type Dl :
J. Nervi / Journal of Algebra 253 (2002) 50–99 61
One remarks that only the circled coefficients are useful to calculate the
ωj (Hi )’s. The Cartan matrix associated to the pair (Π, Γ ) is then
2 −2 0 . . . 0
−1 2 −1 . . . 0
. ..
M = .
. ..
.
..
.
..
.
.
0 . . . −1 2 −1
0 . . . 0 −2 2
and the Dynkin diagram is
• [hΓ , hΓ ] = {0};
j
• as [d1i , d−1 ] = {0} for i = j , one has [Xi , Yj ] = δij Hj for all i, j ;
• as d1 = α=ωj (mod Γj ) ĝα (respectively d−1 = α=ωj (mod Γj ) ĝ−α ) one has
j j
not.
Remark 1.4.2.5. The subalgebra ĝΓ depends on the choice of the elements
Xj , Yj . Obviously all admissible subalgebras corresponding to the same subset Γ
are isomorphic.
Remark 1.4.2.6. If the subset Γ = ∅ then one has hΓ = h and ĝΓ = ĝ.
J. Nervi / Journal of Algebra 253 (2002) 50–99 63
there exists in Π\Γ a unique root such the associated connected component of
S(A) (as defined in Section 1.2) contains α0 . Such a root will be denoted by ω0
(possibly ω0 = α0 ). The root ω0 of Ψ corresponds to the so-called 0th vertex on
(1)
the Dynkin diagram of ĝΓ . In the following example ĝ is of type E7 , ĝΓ is of
(2)
type A5 and one has ω0 = α0 :
E7(1) A(2)
5
One has δ = ni=0 aαi αi where aα0 = 1 (see Remark 1.5.1). Then δ =
k
j =0 aωj ωj . As Hj ∈ h for any j , one has δ(Hj ) = 0 and
k
δ(dΓ ) = aωj ωj (dΓ ) = aω0 .
j =0
Hence
aω0
δ= δΓ .
b0
(2)
If ĝΓ is not of type A2l then b0 = 1 and the result follows.
If ĝΓ is of type A(2)
2l then b0 = 2. One checks that for each case where the
(2)
C-admissible algebra ĝΓ is of type A2l the number aω0 is an even integer. From
Table A the list of such (Π, Γ ) is given by the following diagrams (where the
numerical labels are the coefficients of the root δ).
(1) (2)
E7 A2
(1) (2)
E8 A2
E8(1) A(2)
4
(1) (2)
E8 A2
E6(2) A(2)
4
Proposition 1.5.5. ∆Γ ⊂ ∆.
Proof.
One recall that a ĝΓ -module V is called hΓ -diagonalizable if V =
λ∈hΓ Vλ where Vλ = {v ∈ V : h(v) = λ(h)v for h ∈ hΓ } and it is called
integrable if all Xj , Yj for j = 0, . . . , k are locally nilpotent on V .
The ĝΓ -module ĝ is clearly hΓ -diagonalizable since one has ĝ = α∈∆∪{0} Vα
where
Vα = ĝβ = x ∈ ĝ: [h, x] = α(h)x, h ∈ hΓ .
β=α
j j j
• If αi ∈ Γj then ei ∈ d0 and (ad Xj )2 (ei ) ∈ [d1 , d1 ] = {0}.
• If αi ∈ Γ \Γj then (ad Xj )(ei ) = 0.
• If αi ∈
/ Γ (i.e. α i = 0) then there exists l such that αi = ωl and, for some
integer N , one has
(ad Xj )N (ei ) ∈ VNωj +ωl = ĝβ .
β=Nωj +ωl (mod Γ )
For N big enough one has ĝβ = {0} and the result follows.
For fi ∈ ĝ−αi let us consider the following cases:
j j
• If αi ∈ Γj then (ad Xj )2 (fi ) ∈ [d1 , d1 ] = {0}.
• If αi ∈ Γ \Γj then (ad Xj )(fi ) = 0.
j
• If αi = ωl (l = j ) then (ad Xj )(fi ) = 0 because [d1 , d−1
l
] = {0}.
j
• If αi = ωj then (ad Xj )(fi ) ∈ d0 ; so (ad Xj )3 (fi ) = 0.
66 J. Nervi / Journal of Algebra 253 (2002) 50–99
Remark 1.5.7. A slight modification of the preceding proof shows that ĝ is still
an integrable ĝΓ -module even if (Π, Γ ) is admissible.
Theorem 1.5.8. The root system ∆Γ of the C-admissible algebra ĝΓ is ∆\{0}.
Therefore one has ∆ ⊂ ∆Γ ∪ {0} and Proposition 1.5.5 implies the final
result. ✷
J. Nervi / Journal of Algebra 253 (2002) 50–99 67
2.1. Definition
The definition of a Lie algebra graded by a finite root system has been
introduced by S. Berman and M. Moody (see also the papers by Benkart and
Zelmanov [2], Allison et al. [4], and Nervi [5]). In this work we study more
particularly affine algebras graded by affine root systems. This needs a slight
extension of the definition given in [4].
Definition 2.1.1. Let Σ be an affine root system. An affine Lie algebra ĝ over C
is graded by Σ or Σ-graded if
(ii) ĝ = α̃∈Σ∪{0} Vα̃ where Vα̃ = {x ∈ ĝ: [h, X] = α̃(h)x for all h ∈ ha }.
Remark 2.1.2. For any α̃ ∈ Σ ∪ {0} the root space aα̃ is included in Vα̃ and one
has Vα̃ = {0}.
Remark 2.1.3. As α̃∈Σ ([Vα̃ , V−α̃ ] ⊕ Vα̃ ) is a non-zero ideal of ĝ that contains
[a, a], we have [ĝ, ĝ] = α̃∈Σ ([Vα̃ , V−α̃ ] ⊕ Vα̃ ) [11, Proposition 1.7.b].
In this section we establish a link with the grading affine subalgebras of ĝ and
the C-admissible subalgebras.
Let ĝ be an affine algebra graded by an affine root system Σ. Denote by
B = { ω0 , . . . ,
ωk } (k 1) a root basis of Σ and by a the grading subalgebra.
The elements of the Cartan subalgebra ha of a are semisimple since ĝ is ha -
diagonalizable (condition (ii) of Definition 2.1.1). One deduces by applying the
conjugacy theorem of Kac–Peterson [14] that there exists a Cartan subalgebra h
of ĝ containing ha .
Put ∆ = (ĝ, h) and ρ : h → ha , α → α|ha .
Remark
2.3.1. For any α̃ ∈ Σ ∪ {0}, the subspace
Vα̃ of the Σ-gradation is equal
to ρ(γ )=α̃ ĝγ and one has V0 = Zĝ (ha ) = h ⊕ ρ(γ )=0 ĝγ .
∆+ (Π) ⊂ P.
Im
with 2j − 1 > 0, and the conclusion is the same. Hence we always have
|∆+ (Π) ∩ ((−P)\S)| < ∞. If the previous set is empty then one has obviously
∆+ (Π) ⊂ P.
Suppose now that ∆+ (Π) is not a subset of P, then there exists a root
αi ∈ Π such that ρ(αi ) < 0. Consider the reflection s−αi [11, p. 35] and set
Π = s−αi (Π). The set Π is a root basis of ∆ [11, Proposition 5.9] and s−αi
preserves the root δ. We clearly have ρ(−αi ) > 0; so −αi ∈ / (P)\S (1). On the
other hand, one has s−αi (αi ) = −αi ∈ Π hence −αi ∈ ∆+ (Π ) (2).
Let us prove now that ∆+ (Π ) ∩ ((−P)\S) is a proper subset of ∆+ (Π) ∩
((−P)\S). Note that ∆+ (Π ) ∩ ((−P)\S) ⊂ ∆+ (Π )\{−αi } (cf. (1) and (2)).
We know that the set ∆+ (Π )\{−αi } is s−αi -invariant [11, Lemme 3.7]. Hence
for any β ∈ ∆+ (Π )\{−αi }, one has
β = (s−αi )2 (β) ∈ s−αi ∆+ (Π ) .
Since Π = s−αi (Π ), we have s−αi (∆+ (Π )) ⊂ ∆+ (Π). Therefore
∆+ (Π ) ∩ (−P)\S ⊂ ∆+ (Π) ∩ (−P)\S .
70 J. Nervi / Journal of Algebra 253 (2002) 50–99
Corollary 2.3.5. The map ρ is a surjection from ∆+ ∪ {0} onto Σ+ ∪ {0} and
ρ −1 (Σ+ ) ⊂ ∆+ .
Proposition 2.3.7. If the set S = P ∩ (−P) is not empty then there exists a non-
empty subset Γ of Π , such that ∆+ ∪ Γ − = P. This condition is equivalent to
Γ = P ∩ (−P).
Proof. The condition S = ∅ implies that there exists a simple root in (−P)
(Remark 2.3.6). Set Γ = Π ∩ (−P) = {γ ∈ Π: ρ(γ ) = 0}. Γ is a proper subset
of Π and Γ is a finite root system.
Let us prove now that Γ − = ∆− ∩ P. Let α be an element of Γ − , write
α = −β1 − · · · − βn (βj ∈ Γ ). Let us prove by induction on n that Γ − ⊂
∆− ∩ P. For n = 1, one has α = −β1 ; since β1 ∈ Γ ⊂ (−P) then −β1 ∈ ∆− ∩ P.
Let n be a positive integer, suppose that if α = −β1 − · · · − βn , βj ∈ Γ then
α ∈ ∆− ∩ P. Let α be a root such that α = −β1 − · · · − βn − βn+1 (βj ∈ Γ ).
There exists a permutation σ such that α = −βσ (1) − · · · − βσ (n) − βσ (n+1) and
the partial sums are roots. Set γ = −βσ (1) − · · · − βσ (n) . One has γ ∈ ∆− ∩ P
and −βσ (n+1) ∈ ∆− ∩ P. Since α = γ − βσ (n+1) ∈ ∆− and P is closed, then
α ∈ ∆− ∩ P.
Let us prove now the converse inclusion by induction too. Let α be an
element of ∆− ∩ P, one can write α = −β1 − · · · − βn (βj ∈ Π) If n = 1 then
α = −β1 ∈ P and β1 ∈ Π ∩ (−P) = Γ . Let n be a positive integer, suppose
that if α = −β1 − · · · − βn (βj ∈ Π) and if α ∈ ∆− ∩ P, then α ∈ Γ − . Let α
be an element of ∆− ∩ P such that α = −β1 − · · · − βn − βn+1 (βj ∈ Π).
There exists a permutation σ such that α = −βσ (1) − · · · − βσ (n) − βσ (n+1) and
the partial sums are roots. Set γ = −βσ (1) − · · · − βσ (n) . One has γ ∈ ∆− and
γ = α + βσ (n+1) ; since P is closed then γ ∈ P. We deduce that γ ∈ Γ − .
Furthermore, α − γ = −βσ (n+1) and −γ , α ∈ P, we deduce that βσ (n+1) ∈
(−P) ∩ Π = Γ . ✷
J. Nervi / Journal of Algebra 253 (2002) 50–99 71
k
= mj
ωj .
j =0
The set of the q-uplets (n1 , . . . , nq ) which are solutions of the preceding system
of equations is finite. And for each q-uplet (n1 , . . . , nq ), the number of the roots
in ∆ whose coefficients relative to the simple roots in a non-empty subset of Π
(here Π\Γ ) are fixed, is finite. One deduces that |ρ −1 ({α̃})| < ∞. ✷
Recall that since ρ is a surjection from ∆ ∪ {0} onto Σ ∪ {0}, the set Γ =
{α ∈ Π: ρ(α) = 0} is a proper subset (possibly empty) of Π . Hence we can define
as in Section 1 a Z-gradation in ĝ relating Γ . One has
d0 = h ⊕ ĝΓ = V0 , d1 = ĝγ .
γ =ω (mod Γ )
ω∈Π\Γ
Proof. From Proposition 2.3.10 it is enough to prove that for any ω ∈ Π\Γ one
has ρ(ω) ∈ B. Suppose that there exists ω ∈ Π\Γ such that ρ(ω) ∈ Σ+ \B. One
can write
ρ(ω) =
ωj1 + · · · +
ωjp , ()
where
ωl ∈ B and p 2. We can also assume that the partial sums are roots.
Suppose first that ρ(ω) is a real root. Let (Xρ(ω) , ρ(ω)∨ , Yρ(ω) ) be the
corresponding sl2 -triplet. Since ĝ is a-graded, any Z ∈ Vρ(ω) verifies
2Z = ρ(ω)∨ , Z = [Xρ(ω) , Yρ(ω) ], Z = Xρ(ω) , [Yρ(ω) , Z] .
We deduce that Vρ(ω) = [V0 , Xρ(ω) ]. From () we deduce that
Xρ(ω) = X ωjp . . . [X ωj1 ] . . .
ωj2 , X
ωjm ∈ a m ⊂ V ωjm ⊂ d1 (Corollary 2.3.11). Hence we have Xρ(ω) ∈ dp
with X ωj
(p 2). Furthermore,
ĝω ⊂ [V0 , Xρ(ω) ] ⊂ [d0 , dp ] ⊂ dp .
This is a contradiction to the fact that we have ĝω ⊂ d1 . Hence if ω ∈ Π\Γ and
if ρ(ω) is real, then ρ(ω) ∈ B.
Suppose now that ρ(ω) is an imaginary root. Since |B| 2, there exists
at least two simple roots in Π\Γ whose restrictions to ha are simple roots of B;
hence one has |(Π\Γ )| 3. Let ωl0 be one of these roots. We set ρ(ωl0 ) = ω ∈ B.
It is possible to choose ωl0 and
ω such that there exists a sequence (α1 , . . . , αs )
in Γ (possibly empty) such that the sum ξ = ω + α1 + · · · + αs + ωl0 is a positive
root of ∆. Since |(Π\Γ )| 3, the support of ξ is a proper subset of Π , hence ξ is
a real root of ∆. One has ρ(ξ ) = p(ω) +
ω ∈ Σ+ and since ρ(ω) is an imaginary
J. Nervi / Journal of Algebra 253 (2002) 50–99 73
root, ρ(ξ ) is real. Since ρ(ω) is imaginary, supp(ρ(ω)) = B and ρ(ξ ) is the sum
of p simple roots (p 3). As in the first case, we can write
ĝω = ĝξ , ĝ−α1 −···−αs −ωl0 ⊂ [Vρ(ξ ), V− ω ⊂ dp−1 .
ω ] = [V0 , Xρ(ξ ) ], V−
Let ω ∈
Π\Γ . Let Πω be the connected component of Γ ∪ {ω} containing ω.
Let gω = p∈Z dp (ω) be the Z-gradation of the corresponding finite-dimensional
subalgebra gω corresponding to Πω (see Section 1.1).
Corollary 2.3.14. The algebra ĝ and the grading algebra a have the same one-
dimensional center CK.
Proof. Let
ωj ∈ B. Denote by ωj its unique preimage in Π\Γ , one has
ωj∨ ∈ a
ωj −
, a ωj ⊂ [V ωj ] ⊂ d1 (ωj ), d−1 (ωj ) ⊂ d0 (ωj ).
ωj ,V−
74 J. Nervi / Journal of Algebra 253 (2002) 50–99
Theorem 2.4.3. Let ĝ be an affine algebra and Σ be an affine root system. Denote
a basis of Σ by B = { ω0 , . . . ,
ωk }. The following conditions are equivalent:
ĝ = C t, t −1 ⊕ sl4 (C) ⊕ CK ⊕ Cd,
where C[t, t −1 ] is the algebra of Laurent polynomials in t and sl4 (C) the
simple Lie algebra of complex matrices of trace zero. (For details about concrete
constructions of this affine algebra see [11, Chapter 7].) In the underlying
subalgebra g̊ = sl4 (C), consider h̊ the Cartan subalgebra of the diagonal matrices.
Denote by (α1 , α2 , α3 ) the standard basis of ∆˚ = (g̊, h̊). The subalgebra h =
h̊ ⊕ CK ⊕ Cd is a Cartan subalgebra of ĝ. Let ∆ be the root system of the pair
(ĝ, h). Let δ be the root of ∆ defined by δ|h̊ ⊕ CK = 0, δ, d = 1.
Put α0 = δ − α1 − α2 − α3 = δ − θ , the sequence (α0 , α1 , α2 , α3 ) is a basis
of ∆ and the coroot of α0 is α0∨ = K − θ ∨ . Denote by (Xαi , αi∨ , Yαi )i=0,...,3 the
standard sl2 -triples. Consider the subalgebra å of sl4 (C) generated by the two sl2 -
J. Nervi / Journal of Algebra 253 (2002) 50–99 75
triples (Xα1 + Xα3 , α1∨ + α3∨ , Yα1 + Yα3 ) and (Xα2 , α2∨ , Yα2 ). The algebra å is the
simple subalgebra (of type C2 ) of the matrices
a c −d e
−a + f
c b d
.
−d b a−f c
e d c −a
Denote by hå the Cartan subalgebra generated by α1∨ + α3∨ and α2∨ . Put ω1 =
α1 |hå = α3 |hå and ω2 = α2 |hå ; (
ω1 ,
ω2 ) is a basis of Σ̊ = (å, hå ). (For details
about these calculations see [5, Example 5.5].) Put a = C[t, t −1 ] ⊕ å ⊕ CK ⊕ Cd,
then a is an affine subalgebra of ĝ of type C2(1) and ha = hå ⊕ CK ⊕ Cd is a
Cartan subalgebra of a. Denote by Σ the root system of (a, ha ) and define δa
in ha by the relations δa |h̊å ⊕ CK = 0, δa , d = 1. Put ω0 = δa − (2 ω1 + ω2 );
ω1 ,
( ω1 ,
ω2 ) is a basis of Σ. Define as precedently ρ : h → ha , α → α|ha . We
have ρ(α1 ) = ρ(α3 ) = ω1 , ρ(α2 ) =
ω2 , ρ(δ) = δa , ρ(α0 ) =
ω0 . It is easy to verify
that ρ(∆) = Σ. The algebra ĝ is Σ-graded and for any α̃ ∈ Σ ∪ {0}, we have
Vα̃ = ĝγ .
γ ∈∆∪{0}
ρ(γ )=α̃
A(1) (1)
3 graded by C2
In this example, as in the sequel of the paper, we have joined by an arrow the
simple roots having the same restriction.
Let ĝ be an affine algebra graded by an affine root system Σ. The notations are
the same as precedently. The following definition is useful.
Definition 2.5.1. The two different roots α and β of Π\Γ are said Γ -connected
if (α, β) = 0 or if there exists in S(A) a string whose extremities are α and β and
whose vertices (except α and β) are all in Γ .
Proposition 2.5.2. Let α and β in Π\Γ be two roots which are Γ -connected, one
has ρ(α) = ρ(β) and (ρ(α), ρ(β)) = 0.
76 J. Nervi / Journal of Algebra 253 (2002) 50–99
and the last sum is direct since, from Proposition 2.5.2, the roots ωl are not Γ -
connected in Π\Γ . Hence the decomposition ω∨ = h1 + · · · + hq is unique. One
ω ) = α(hl ) = 0 for all α ∈ Γl , and ωl (
has α( ∨
ω∨ ) = ωl (hl ) = ω( ω∨ ) = 2. We
deduce that hl = Hl for all l = 1, . . . , q. ✷
Theorem 2.5.5. The pair (Π, Γ ) is C-admissible and one can choose the C-
admissible subalgebra ĝΓ such that ĝΓ contains the grading subalgebra a.
we can write
ω = X1 + · · · + Xq
X ω = Y1 + · · · + Yq ,
and Y
where Xl ∈ d1 (ωl ) and Yl ∈ d−1 (ωl ). One has, from Proposition 2.5.3,
q
q
ω∨ = [X
ω] =
ω , Y [Xl , Yl ] = Hl .
l=1 l=1
The decomposition is unique, so we have [Xl , Yl ] = Hl for all l and the space
(d0 (ωl ), d1 (ωl )) is regular (Section 1.1). Since ρ is a surjection from Π\Γ
J. Nervi / Journal of Algebra 253 (2002) 50–99 77
onto B, it is clear that the same is true for any j = 0, . . . , card(Π\Γ ); hence
the pair (Π, Γ ) is admissible.
Suppose now that there exist j ∈ {0, . . . , card(Π\Γ )} and p 2 such that
dp (ωj ) = {0}. This means that there exists a root σ ∈ ∆+ that verifies σ =
pωj (mod Γ ), hence ρ(σ ) = pρ(ωj ) ∈ Σ+ . This is impossible since ρ(ωj ) is a
real root of Σ. The pair (Π, Γ ) is therefore C-admissible. Consider now the affine
subalgebra ĝΓ generated by hΓ = {h ∈ h: α(h) = 0 ∀α ∈ Γ } and the sl2 -triples
(Xj , Hj , Yj ). The algebra ĝΓ contains ha and the generators X ω , Y ω ∈ B).
ω (
Hence ĝΓ contains a. ✷
Proposition 2.5.6. The root system Σ grades ĝΓ and the corresponding grading
subalgebra is a. Moreover, if we apply the preceding procedure to this grading,
the associated C-admissible subalgebra is ĝΓ itself.
ρ : h −→ hΓ
α −→ α|hΓ = α.
One knows that RΓ = ∆\{0} and Π\Γ is a basis of RΓ (Theorem 1.5.9).
Consider
ρΓ : hΓ −→ ha
α −→ α|ha .
One has ρ = ρΓ ◦ ρ. Since ρ is a surjection from ∆+ ∪ {0} onto Σ+ ∪ {0} and
since Π\Γ is a basis of RΓ , one has ρΓ−1 ({0}) ∩ R+ Γ = ∅ and ρΓ is clearly a
surjection from R+ onto Σ + . Put V α̃ = ĝ α for any α̃ ∈ Σ ∪ {0}. Note
Γ ρΓ (α)=α̃ Γ
that V0 = ĝ0Γ = hΓ and ĝΓ = α̃∈Σ∪{0} Vα̃ . Hence ĝΓ is Σ-graded. In this new
gradation we can construct a C-admissible subalgebra of ĝΓ that contains a by
the procedure of Theorem 2.5.5. Since the set {α ∈ Π\Γ : ρΓ (α) = 0} is empty,
then this subalgebra is ĝΓ itself. ✷
Proof. Denote by ĝ2 the grading subalgebra in ĝ1 which is associated to Σ2 . One
has
ĝ2 ⊂ ĝ1 ⊂ ĝ.
Let h2 be a Cartan subalgebra of ĝ2 , by applying the conjugacy theorem of Cartan
subalgebras [14, Theorem 2] one can find a Cartan subalgebra h1 (respectively h)
of ĝ1 (respectively ĝ) such that
h2 ⊂ h1 ⊂ h.
In the corresponding root systems ∆, Σ1 , Σ2 , one can choose basis Π, Π1 , Π2
such that
ρ : h −→ h1
α −→ α|h1 = α
is a surjection from ∆+ ∪ {0} onto Σ1+ ∪ {0} (Corollary 2.3.5) and
ρ̃ : h1 −→ h2
α −→ α|h2 = α̃
is a surjection from Σ1+ ∪ {0} onto Σ2+ ∪ {0} (Corollary
2.3.5). The spaces of the
Σ1 -gradation are denoted by Vα and one has ĝ = α∈Σ1 ∪{0} Vα . For any α̃ ∈ Σ2 ,
let us set:
Uα̃ = Vα .
α|h2 =α̃
It is easy to prove that ĝ is Σ2 -graded and that the spaces of the gradation are the
Uα̃ ’s, α̃ ∈ Σ2 ∪ {0}. ✷
Theorem 2.5.10. Let ĝ an affine algebra and Σ be an affine root system. The two
following conditions are equivalent:
(i) ĝ is Σ-graded.
(ii) There exist a root system ∆ of ĝ, a basis Π of ∆, a subset Γ of Π such that
– the pair (Π, Γ ) is C-admissible;
– the C-admissible subalgebra ĝΓ , associated to the pair (Π, Γ ) is graded
by Σ and Σ is maximal in ĝΓ .
J. Nervi / Journal of Algebra 253 (2002) 50–99 79
It is now clear that the study of the Σ-gradations of the affine algebras
can be reduced to the study of the maximal Σ-gradations of its C-admissible
subalgebras. Recall that maximality is equivalent to say that the subset Γ of Π
precedently defined is empty.
We can assume without loss of generality that the homomorphism ρ is not
a bijection from Π onto B = { ω0 , . . . ,
ωk } the basis of Σ that we have fixed once
for all. Recall that it is equivalent to say that the representations (V0 , V ωj ) are
not all irreducible (Proposition 2.4.1). Such maximal gradations will be called
non-trivial (abbreviated n.t.m.).
For the trivial maximal gradations the grading subalgebra is the algebra ĝ itself
(Theorem 2.4.3).
In this section we assume that the gradation of the affine algebra ĝ by the root
system
Σ is maximal non-trivial. Let Π i = ρ −1 ({
ωi }) ∩ Π and set H i = ωi∨ =
γ ∈Π i γ for all i = 0, . . . , k.
∨
(a) (α, β) = 0;
(b) α(H j ) = β(H j ) for all j = 0, . . . , k;
(c) let γ be a root in Π\{α} such that (α, γ ) = 0, then γ ∈
/ Π i and there exists
a simple root µ such that ρ(µ) = ρ(γ ) and (µ, β) = 0. If (γ , β) = 0 then
µ = γ .
Theorem 2.5.5:
X ωi = Xα + Xβ + Xλ ,
λ∈Π i \{α,β}
j j
ωi =
H , X ωi = α H X
ωi H j X ωi ,
j
H , X
ωi = γ∨ + ξ ∨ , Xα + Xβ + Xλ
ξ ∈Π j \{γ } λ∈Π i \{α,β}
80 J. Nervi / Journal of Algebra 253 (2002) 50–99
= α(γ ∨ )Xα + β(γ ∨ )Xβ + α(ξ ∨ )Xα
ξ ∈Π j \{γ }
+ β(ξ ∨ )Xβ + λ(ξ ∨ )Xλ .
ξ ∈Π j \{γ } ξ,λ
Note that α(γ ∨ ) < 0 and α(ξ ∨ ) 0 for all ξ ∈ Π j . Hence the coefficient of Xα
in the previous sum (i.e. α(γ ) + ξ ∈Π j \{γ } α(ξ ∨ )) is not zero.
∨
In the following we give the list of the affine algebras which are not graded by
an n.t.m. root system. For the others we give the possible n.t.m. Σ-gradations. We
give too some representative proofs.
Proof. The result is straightforward for the three first algebras by applying
Lemma 2.6.1(a). For D4(3) , note that the only possibility can be illustrated by the
following arrow diagram:
Lemma 2.6.3. If the algebra ĝ is of one of types F4(1) , E6(2) , A(2) (2) (1)
2l , Dl+1 , or Cl ,
−1 −1
and if ρ ({ρ(α0 )}) ∩ Π = {α0 } then for all i = 0, . . . , n, ρ ({ρ(αi )}) ∩ Π =
{αi } and the gradation is trivial.
Proof. Note that the Dynkin diagrams of such algebras are linear (i.e. non-
circular) and without branching point. Let j0 the smallest index 1 such
that |ρ −1 ({ρ(αj0 )}) ∩ Π| 2. Denote by αj0 +k a root in this set. One has
k 2 (Lemma 2.6.1(a)) and ρ(αj0 −1 ) = αj0 +k−1 (or αj0 +k+1 if this root exists)
(Lemma 2.6.1(c)). It follows that |ρ −1 ({ρ(αj0 −1 )}) ∩ Π| 2 and this is not
possible from to the choice of j0 . ✷
Proof. Suppose that ĝ (of type E6(2) ) admits an n.t.m. Σ-gradation, then one has
|ρ −1 ({ρ(α0 )}) ∩ Π| 2. If ρ(α0 ) = ρ(α2 ), by applying Lemma 2.6.1(c) one has
ρ(α1 ) = ρ(α3 ) and ρ(α4 ) = ρ(α2 ). Hence Π 0 = {α0 , α2 , α4 }, Π 1 = {α1 , α3 },
α0 (H 1 ) = α0 (α1∨ ) = −1 and α2 (H 1 ) = α2 (α1∨ + α3∨ ) = −2; this is not possible
by Lemma 2.6.1(b). This case is illustrated by the following diagram:
(2)
The other two possibilities for E6 (illustrated by the following diagrams) are
treated similarly.
(2)
The case A2l . The case l = 1 has been studied in Proposition 2.6.2. The case
l = 2 is straightforward by Lemma 2.6.1(b).
Let us assume l 3 and choose notations such that α0 ∈ Π 0 and α1 ∈ Π 1 .
Denote by j0 the smallest non-zero integer such that ρ(α0 ) = ρ(αj0 ). We clearly
have j0 2. If j0 < l then ρ(α1 ) = ρ(αj0 −1 ) = ρ(αj0 +1 ) by Lemma 2.6.1(c), one
has α0 (H 1 ) = −1 and αj0 (H 1 ) = −2. This is not possible by Lemma 2.6.1(b)
then j0 = l. It follows that ρ(α1 ) = ρ(αl−1 ). This is not possible by Lemma
2.6.1(b) and the fact that α1 (H 0 ) = −2 and αl−1 (H 0 ) = −1. The following
diagram illustrates this case:
(1)
The case E8 is straightforward. ✷
Proposition 2.6.5.
– If the algebra G(1)2 admits an n.t.m. Σ-gradation then Σ is necessarily of
(2)
type A2 .
(1)
– If the algebra C2 admits an n.t.m. Σ-gradation then Σ is necessarily of
type A(1)
1 .
– If the algebra F4(1) admits an n.t.m. Σ-gradation then Σ is of type D4(3) .
(2)
– If l + 1 is even then the algebra Dl+1 does not admit any n.t.m. Σ-gradation.
(2)
– If l + 1 is odd and if the algebra Dl+1 admits an n.t.m. Σ-gradation then Σ
(2)
is of type A2p with p = l/2.
82 J. Nervi / Journal of Algebra 253 (2002) 50–99
or A(2) (2)
2p+1 or A2p with l = 2p + 2 in these two last cases.
(2) (2)
– If the algebra A2l−1 admits an n.t.m. Σ-gradation then Σ is of type A2l−2 .
(2)
The case Dl+1 . Note that l 2. By Lemma 2.6.1(b), we prove that if |Π 0 | 2
then Π 0 = {α0 , αl }. We prove then by induction that for all j = 0, . . . , [l/2], one
has ρ −1 ({ρ(αj )}) = {αj , αl−j }. If l + 1 = 2p then ρ(αp ) = ρ(αp−1 ). This is
impossible by Lemma 2.6.1(a). The case l + 1 = 2p + 1 is illustrated by the
following diagram:
The computation of the ωj (H i )’s gives the type of the Cartan matrix of the
possible grading root system: A(2)
2p . ✷
For the other cases the proofs are very similar and use essentially Lemma 2.6.1.
(1)
In the last proposition we give the result for the circular case Al .
J. Nervi / Journal of Algebra 253 (2002) 50–99 83
Remark 2.6.6. Because of the circularity of the diagram, one can replace α0 with
any other simple root and change the indices. The result is the same.
Proposition 2.6.7.
– If algebra Al (l 3) admits an n.t.m. Σ-gradation and ρ −1 ({ρ(α0 )}) ∩
(1)
Π = {α0 } then for all j ∈ {0, . . . , l} one has ρ −1 ({ρ(αj )})∩Π = {αj , αl+1−j }
(one sets αl+1 = α0 ). The integer l is odd and if we set l = 2p + 1, then Σ is
(1)
of type Cp+1 .
(1)
– If algebra Al admits an n.t.m. Σ-gradation and satisfies the hypothesis:
() it does not exist any simple root αj such that ρ −1 ({ρ(αj )}) ∩ Π = {αj }
then Σ is of type A(1) (1)
p with l = mp − 1 or Cp with l = 2mp + 1.
We prove in this section that the possible non-trivial maximal gradations given
in Section 2.6 are effective gradations.
Let ĝ be an affine algebra. Let h be a Cartan subalgebra and let ∆ be the
root system of the pair (ĝ, h). We use the standard notations from the preceding
sections. We take now one of the diagrams obtained in Section 2.6 (listed in
Appendix B) and consider
the corresponding partition of a basis Π of ∆. That it
is to say that Π = ki=0 Π i where Π i is either a simple root or a set of orthogonal
simple roots connected together by arrows. Let us set α0 ∈ Π 0 . Put
Hi = α∨ , Xi = Xα , Yi = Yα ,
α∈Π i α∈Π i α∈Π i
Lemma 2.7.1. One has α(K ) = 0 for all α ∈ Π and K is an element of the
one-dimensional center CK of ĝ.
Remark 2.7.2. Since a0∨ = 1 [11, p. 79], the elements K , H 1 , . . . , H k are linearly
independent.
The integer N is such that vN = 0 and vN−1 = 0. From (1), one deduces that
vj = 0 for all j N and from (2), that −m01 − N + 1 = 0 therefore N = 1 − m01 .
And the result follows.
The others equalities are proved in the same way. ✷
Denote by Σ the root system of the pair (a, c). The type of Σ is given by the
matrix M.
If αi0 ∈ Π j then (ad Xj )(Yαi0 ) = α∈Π j [Xα , Xαi0 ] = αi∨0 . One deduces that
(ad Xj )2 (Yαi0 ) = Xα , αi∨0 = −2Xαi0 ;
α∈Π j
Proposition 2.7.8. The root system Σ of the pair (a, c) is the set ρ(∆). The
algebra ĝ is maximally graded by Σ and the corresponding algebra is a.
Moreover, all non-trivial maximal gradings are obtained this way. The list of the
non-trivial maximal gradings of affine algebras is given in Appendix B.
Proof. We know that Σ ⊂ ρ(∆), let us prove the converse inclusion. Using
Proposition 2.7.7 we show as in Theorem 1.5.8 that for any α ∈ ∆, there exist
an integer n and a root ω in Σ such that ρ(α) = nω. If ω is an imaginary root
then ρ(α) ∈ Σ. If ω is a real root then up to conjugation by an element of the
Weyl group of a, we can assume that ω is a simple root. Assume that n is a
positive integer. This means that α is the sum of n roots of Π whose restrictions
to c are ω. Since such simple roots are orthogonal, this is impossible if n 2.
Hence ρ(α) = ω and ρ(∆) ⊂ Σ.
Let us prove now that ĝ is Σ-graded. Let α ∈ Σ ∪ {0} and set
Vα = ĝγ .
ρ(γ )=α
In the left column of Table A is the Dynkin diagram of the affine algebra ĝ
and in the right column is the Dynkin diagram of the corresponding admissible
(eventually C-admissible) a.
J. Nervi / Journal of Algebra 253 (2002) 50–99 87
Table A.1
The algebra ĝ is of type (Aff1)
(1) (1)
A(k+1)(p+1)−1 C-admissible Ak
(1) (2)
Bn C-admissible if p = 1 Dk if k 2
(1)
A1 if k = 1
(1) (2)
Bn C-admissible A2k
(1) (1)
Ckp Ck
(1) (1)
Bn C-admissible Bk
(1) (1)
B3 C-admissible A1
88 J. Nervi / Journal of Algebra 253 (2002) 50–99
(1) (1)
Dn C-admissible Bk
(1) (2)
Cn A2k
(1) (2)
Cn Dk
(1) (2)
Dn C-admissible if p = 1 Dk
(1) (2)
Dn A2k
(1) (1)
Dkp C-admissible Ck
(1) (1)
E6 C-admissible A2
J. Nervi / Journal of Algebra 253 (2002) 50–99 89
(1) (1)
E6 C-admissible A2
(1) (1)
E6 C-admissible G2
(1) (2)
E7 C-admissible A2
(1) (2)
E7 A2
(1) (1)
E7 A1
(1) (2)
E7 A2
90 J. Nervi / Journal of Algebra 253 (2002) 50–99
(1) (2)
E7 A2
(1) (2)
E7 A2
(1) (2)
E7 A4
(1) (1)
E7 G2
(1) (1)
E7 C2
(1) (2)
E7 A5
(1) (1)
E7 C-admissible C3
J. Nervi / Journal of Algebra 253 (2002) 50–99 91
(1) (1)
E7 C-admissible F4
(1) (2)
E8 C-admissible A2
(1) (2)
E8 A2
(1) (1)
E8 A1
(1) (2)
E8 A2
(1) (2)
E8 A2
(1) (2)
E8 A2
(1) (1)
E8 A1
92 J. Nervi / Journal of Algebra 253 (2002) 50–99
(1) (2)
E8 A2
(1) (2)
E8 C-admissible A4
(1) (1)
E8 C2
(1) (2)
E8 A4
(1) (3)
E8 C-admissible D4
(1) (1)
E8 G2
(1) (1)
E8 C-admissible G2
(1) (2)
E8 C-admissible E6
J. Nervi / Journal of Algebra 253 (2002) 50–99 93
(1) (1)
F4 C-admissible G2
(1) (2)
F4 A2
(1) (2)
F4 A2
(1) (2)
G2 A2
(2) (2)
A2l Dk+1
(2) (2)
A2l A2k
(2) (2)
A2l−1 Dk+1
94 J. Nervi / Journal of Algebra 253 (2002) 50–99
Table A.2
The algebra ĝ is of type (Aff2)
(2) (2)
A2l−1 A2k
(2) (2)
A2l−1 A2k
(2) (1)
A2kp−1 C-admissible Ck
(2) (2)
Dl+1 C-admissible if p = 1 Dk+1
(2) (2)
E6 A2
(2) (2)
E6 A2
(2) (2)
E6 A2
J. Nervi / Journal of Algebra 253 (2002) 50–99 95
(2) (2)
E6 C-admissible A4
The Dynkin diagram of ĝ takes place in the left column (with arrows joining
roots which have the same restriction) while the right column contains the Dynkin
diagram of the maximal grading subalgebra a.
Table B
(1) (2)
G2 A2 ρ(δ) = 2δa
(1) (3)
F4 D4 ρ(δ) = 3δa
(2) (2)
Dp+1 A2p+2 ρ(δ) = δa
(1) (2)
E7 E6 ρ(δ) = 2δa
96 J. Nervi / Journal of Algebra 253 (2002) 50–99
Table B (continued)
(1) (1)
E6 F4 ρ(δ) = δa
(1) (3)
E6 D4 ρ(δ) = 3δa
(1) (1)
B3 G2 ρ(δ) = δa
(1) (2)
Bl Dl ρ(δ) = 2δa
(1) (2)
B2p+1 A2p ρ(δ) = 2δa
(1) (1)
Dl Bl−1 ρ(δ) = δa
J. Nervi / Journal of Algebra 253 (2002) 50–99 97
Table B (continued)
(1) (2)
Dl Dl−1 ρ(δ) = 2δa
(1) (2)
D2p+2 A2p+1 ρ(δ) = 2δa
(1) (2)
D2p+2 A2p+2 ρ(δ) = 2δa
(2) (2)
A2l−1 A2l−2 ρ(δ) = δa
(1) (1)
Cl A1 ρ(δ) = (l − 1)δa
(1) (1)
C2mp Cp ρ(δ) = 2mδa
98 J. Nervi / Journal of Algebra 253 (2002) 50–99
Table B (continued)
(1) (1)
A2p+1 Cp+1 ρ(δ) = 2δa
(1) (1)
Amj −1 Aj −1 ρ(δ) = (m − 1)δa
0 0
(1) (1)
A2mp−1 Cp ρ(δ) = (m − 1)δa
References
[1] S. Berman, R. Moody, Lie algebras graded by finite root systems, Invent. Math. 108 (1992) 323–
347.
[2] G. Benkart, E. Zelmanov, Lie algebras graded by finite root systems and intersection matrix
algebras, Invent. Math. 126 (1996) 1–45.
[3] E. Neher, Lie algebras graded by J-graded root systems and Jordan pairs covered by grids, Amer.
J. Math. 118 (2) (1996) 439.
[4] B. Allison, G. Benkart, Y. Gao, Central extensions of Lie algebras graded by finite root systems,
Math. Ann. 316 (2000) 499–527.
J. Nervi / Journal of Algebra 253 (2002) 50–99 99
[5] J. Nervi, Algèbres de Lie simples graduées par un système de racines et sous-algèbres C-
admissibles, J. Algebra 223 (2000) 307–343.
[6] H. Rubenthaler, Construction de certaines sous-algèbres remarquables dans les algèbres de Lie
semi-simples, J. Algebra 81 (1983) 268–278.
[7] H. Rubenthaler, Les paires duales dans les algèbres de Lie réductives, Astérisque 219 (1994).
[8] M. Sato, T. Kimura, A classification of irreducible prehomogeneous vector spaces and their
relative invariants, Nagoya Math. J. 65 (1977) 1–155.
[9] H. Rubenthaler, Espaces préhomogènes de type parabolique, Thèse, Université de Strasbourg,
1982.
[10] E.B. Vinberg, On the classification of the nilpotent elements of graded Lie algebras, Soviet Math.
Dokl. 16 (1975) 1517–1520.
[11] V.G. Kac, Infinite Dimensional Lie Algebras, third edition, Cambridge Univ. Press, 1990.
[12] N. Bourbaki, Groupes et algèbres de Lie, Masson, Paris, 1981, Chapters 4, 5, 6.
[13] J.P. Serre, Algèbres de Lie semi-simples complexes, Benjamin, 1966.
[14] V.G. Kac, D.H. Peterson, Infinite flag varieties and conjugacy theorems, Proc. Natl. Acad. Sci.
USA 80 (1983) 1778–1782.