arxiv:math/ v1 [math.rt] 13 Oct 2005 June 7, 2018

Similar documents
Topics in Representation Theory: Roots and Weights

The Spinor Representation

Spectral Theorem for Self-adjoint Linear Operators

arxiv: v1 [math.gr] 8 Nov 2008

LECTURE 4: REPRESENTATION THEORY OF SL 2 (F) AND sl 2 (F)

REPRESENTATION THEORY OF S n

Math 210C. The representation ring

ON NEARLY SEMIFREE CIRCLE ACTIONS

THE EULER CHARACTERISTIC OF A LIE GROUP

IRREDUCIBLE REPRESENTATIONS OF SEMISIMPLE LIE ALGEBRAS. Contents

CHAPTER 6. Representations of compact groups

arxiv: v1 [math.rt] 14 Nov 2007

Lemma 1.3. The element [X, X] is nonzero.

Elementary linear algebra

Classification of root systems

implies that if we fix a basis v of V and let M and M be the associated invertible symmetric matrices computing, and, then M = (L L)M and the

Representations of algebraic groups and their Lie algebras Jens Carsten Jantzen Lecture III

A PRIMER ON SESQUILINEAR FORMS

MAT 445/ INTRODUCTION TO REPRESENTATION THEORY

Given a finite-dimensional vector space V over a field K, recall that a linear

Highest-weight Theory: Verma Modules

The Cayley-Hamilton Theorem and the Jordan Decomposition

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

Pacific Journal of Mathematics

Problems in Linear Algebra and Representation Theory

A PIERI RULE FOR HERMITIAN SYMMETRIC PAIRS. Thomas J. Enright, Markus Hunziker and Nolan R. Wallach

Lecture 7: Semidefinite programming

Linear Algebra (part 1) : Vector Spaces (by Evan Dummit, 2017, v. 1.07) 1.1 The Formal Denition of a Vector Space

Since G is a compact Lie group, we can apply Schur orthogonality to see that G χ π (g) 2 dg =

ON SUM OF SQUARES DECOMPOSITION FOR A BIQUADRATIC MATRIX FUNCTION

A NOTE ON REAL ENDOSCOPIC TRANSFER AND PSEUDO-COEFFICIENTS

A PROOF OF BURNSIDE S p a q b THEOREM

Representations of su(2)

Factorization in Polynomial Rings

INTRODUCTION TO LIE ALGEBRAS. LECTURE 7.

32 Divisibility Theory in Integral Domains

LIE ALGEBRAS: LECTURE 7 11 May 2010

SCHUR-WEYL DUALITY FOR U(n)

There are two things that are particularly nice about the first basis

AN UPPER BOUND FOR SIGNATURES OF IRREDUCIBLE, SELF-DUAL gl(n, C)-REPRESENTATIONS

4.4 Noetherian Rings

Group Gradings on Finite Dimensional Lie Algebras

The Cartan Decomposition of a Complex Semisimple Lie Algebra

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

Problem 1A. Suppose that f is a continuous real function on [0, 1]. Prove that

REPRESENTATION THEORY NOTES FOR MATH 4108 SPRING 2012

5 Quiver Representations

LECTURE 3: TENSORING WITH FINITE DIMENSIONAL MODULES IN CATEGORY O

1 Invariant subspaces

Theorem 5.3. Let E/F, E = F (u), be a simple field extension. Then u is algebraic if and only if E/F is finite. In this case, [E : F ] = deg f u.

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

Topics in linear algebra

SPRING 2006 PRELIMINARY EXAMINATION SOLUTIONS

REPRESENTATION THEORY WEEK 7

INVARIANT PROBABILITIES ON PROJECTIVE SPACES. 1. Introduction

Large automorphism groups of 16-dimensional planes are Lie groups

Math 249B. Nilpotence of connected solvable groups

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

Notes 10: Consequences of Eli Cartan s theorem.

Homework 4 Solutions

Math 210B. Artin Rees and completions

ALGEBRAIC GEOMETRY COURSE NOTES, LECTURE 2: HILBERT S NULLSTELLENSATZ.

where m is the maximal ideal of O X,p. Note that m/m 2 is a vector space. Suppose that we are given a morphism

Boolean Inner-Product Spaces and Boolean Matrices

A MODEL-THEORETIC PROOF OF HILBERT S NULLSTELLENSATZ

INTRODUCTION TO LIE ALGEBRAS. LECTURE 2.

Intrinsic products and factorizations of matrices

A PROOF OF BOREL-WEIL-BOTT THEOREM

Decomposing Bent Functions

An Algebraic View of the Relation between Largest Common Subtrees and Smallest Common Supertrees

What is A + B? What is A B? What is AB? What is BA? What is A 2? and B = QUESTION 2. What is the reduced row echelon matrix of A =

SPHERICAL UNITARY REPRESENTATIONS FOR REDUCTIVE GROUPS

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

Chapter 2 The Group U(1) and its Representations

Category O and its basic properties

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

Real representations

ELEMENTARY SUBALGEBRAS OF RESTRICTED LIE ALGEBRAS

Topics in Representation Theory: The Weyl Integral and Character Formulas

SOME GOOD-FILTRATION SUBGROUPS OF SIMPLE ALGEBRAIC GROUPS CHUCK HAGUE AND GEORGE MCNINCH

arxiv: v1 [math.rt] 11 Jun 2012

Representation Theory

arxiv:math/ v3 [math.qa] 16 Feb 2003

Background on Chevalley Groups Constructed from a Root System

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Chapter 1. Preliminaries

Lecture 19 - Clifford and Spin Representations

Linear Algebra. The analysis of many models in the social sciences reduces to the study of systems of equations.

Clifford Algebras and Spin Groups

MATH 423 Linear Algebra II Lecture 33: Diagonalization of normal operators.

THE SEMISIMPLE SUBALGEBRAS OF EXCEPTIONAL LIE ALGEBRAS

Representations of semisimple Lie algebras

Chapter 2: Linear Independence and Bases

Honors Algebra 4, MATH 371 Winter 2010 Assignment 4 Due Wednesday, February 17 at 08:35

Classification of semisimple Lie algebras

Thus, the integral closure A i of A in F i is a finitely generated (and torsion-free) A-module. It is not a priori clear if the A i s are locally

Topics in Representation Theory: SU(n), Weyl Chambers and the Diagram of a Group

n WEAK AMENABILITY FOR LAU PRODUCT OF BANACH ALGEBRAS

Transcription:

The representation ring of a simply connected Lie group as a λ-ring arxiv:math/0510283v1 [math.rt] 13 Oct 2005 June 7, 2018 Abstract Adams and Conway have stated without proof a result which says, roughly speaking, that the representation ring R(G) of a compact, connected Lie group G is generated as a λ-ring by elements in 1-to-1 correspondance with the branches of the Dynkin diagram. In this note we present an elementary proof of this. 1. Introduction Let G be a 1-connected, compact Lie group. A celebrated theorem of Weyl s describes the representation ring R(G) as a polynomial ring R(G) = [V 1,...,V r ], where the fundamental representations V i are in 1-to-1 correspondance with the nodes of the Dynkin diagram. Therefore, in order to understand the representations of G, one basically needs to understand r of them. In his 1985 article [2], J F. Adams presents a method, which he ascribes to J. H. Conway, to cut down on this number. The idea, simply enough, is to exploit the structure of λ-ring which R(G) possesses since it is possible to take exterior powers of representations (see [4] for this we shall not need anything from the general theory of λ-rings, though). The key result announced by Adams is the following. Suppose that the Dynkin diagram of G has an arm of length k; by this we mean that there are nodes v 1,v 2,...,v k such that v i is connected to v i+1 by a single bond for 1 i < k, and such that no other edge runs to any of v 1,v 2,...,v k 1. Then one has: Theorem 1.1 Suppose that the Dykin diagram of the 1-connected, compact Lie group G has an arm of length k. Let V i be the representation corresponding to v i. Then for 1 i k, one has l i V 1 = V i (lower terms) where the lower terms are irreducible representations W < V i. In this statement we order the irreducible representations of G according to their respective maximal weights. 1

For any G-module M, we let Top(M) denote the smallest submodule of M having the same maximal weights, with the same multiplicities. In other words, Top(M) is the smallest G-submodule of M satisfying M = Top(M) (lower terms) (here we mean that any irreducible factor in the lower terms is dominated by at least one irreducible factor in Top(M)). With this terminology, theorem 1.1 can be reformulated thus: V i = Top(l i V 1 ). As another illustration of the Top notation, which is due to Adams, we point out that the traditional proof of Weyl s theorem (eg as in [1]) shows that any irreducible representation of G is of the form Top(V n1 1 V n2 2 Vr nr ). With this result in hand, one can show easily that l i V 1 may replace V i in the statement of Weyl s theorem (this is more or less trivial, but requires a certain amount of notation we shall come back to this in due time). Therefore, if we see R(G) as a λ-ring, the whole arm of the Dynkin diagram is accounted for by the single generator V 1. Adams (unlike Conway) had a proof of theorem 1.1 ready, but wanted to check if anyone else had previously obtained the result before publishing it. His intention seemed to include the proof in a book he had in preparation on exceptional Lie groups. This was sadly prevented by Adams s untimely death. The book has appeared [3], based on his lecture notes, but does not include this proof. In this quick note we present an elementary proof of the main theorem, and give a few applications. 2. Notations G is a 1-connected, compact Lie group, with maximal torus T and Weyl group W. We fix a W-invariant inner product on L(T), the real Lie algebra tangent to T, denoted by,. The roots of G are non-zero weights of the adjoint representation. They come in m pairs ±θ 1,±θ 2,...,±θ m, and we assume that the signs are arranged so that the fundamental Weyl chamber FWC = {v L(T) : θ i (v) > 0} is nonempty. We call θ 1,...,θ m the positive roots; we assume that the indices are so chosen that the simple roots are θ 1,θ 2,...,θ r (where r = dimt). Dually, the fundamental dual Weyl chamber is FDWC = {λ L(T) : λ,θ i > 0}. To each simple root θ i corresponds the fundamental weight ø i characterized by 2 ø i,θ j θ j,θ j = δ ij for any simple root θ j. We partially order weights by requiring λ 1 < λ 2 if and only if λ 1 (v) < λ 2 (v) for all v FWC. 2

For any root θ i, there is an element ϕ i W which induces on L(T) the orthogonal reflexion in kerθ i. For convenience, ϕ 0 will denote the identity. Very explicitly, for λ L(T) or λ L(T) : ϕ i (λ) = λ 2 λ,θ i θ i,θ i θ i. Finally, we assume, of course, that the Dynkin diagram of G has an arm of length k, as explained in the introduction. For simplicity, we arrange the notations so that θ 1,θ 2,...,θ k are the simple roots corresponding to the nodes v 1,v 2,...,v k. We write V i for the representation associated with ø i (or with v i, if you want). We will also write V for V 1 and ø for ø 1. It may be useful to recall that two simple roots θ i and θ j correspond to nodes with a simple bond between them precisely when θ i,θ i = θ j,θ j and 2 θ i,θ j θ i,θ i = 1. For example, we shall use repeatedly ϕ i θ i+1 = ϕ i+1 θ i = θ i +θ i+1 for i < k. 3. Main proof Our first objective is to show that the weight ø i occurs in l i V. Lemma 3.1 For 1 i < k, we have ø i+1 = ø+ϕ 1 ø+ϕ 2 ϕ 1 ø+...+ϕ i ϕ i 1 ϕ 1 ø. Proof. Let λ denote the right hand side. We shall compute λ,θ j. If j > i+1 then we have ϕ s ϕ 1 ø,θ j = ø,ϕ 1 ϕ s θ j = ø,θ j = 0 for s i and it follows that λ,θ j = 0. Ifnow j < i+1, the equationsabovestill hold when s < j 1. The two terms in λ,θ j obtained for s = j 1 and s = j cancel each other, for ϕ j θ j = θ j. Finally, for s j +1 we obtain ø,ϕ 1 ϕ s θ j = ø,ϕ 1 ϕ j+1 θ j = ø,θ j+1 = 0 where we use ϕ j ϕ j+1 (θ j ) = ϕ j (θ j + θ j+1 ) = θ j + θ j + θ j+1 = θ j+1. So λ,θ j = 0 in this case also. It remains to compute λ,θ i+1. Arguing as above, we see that all terms vanish except ø,ϕ 1 ϕ i θ i+1 = ø,θ 1 +θ 2 + +θ i+1 = ø,θ 1 and we obtain λ,θ i+1 = ø,θ 1 = 1 2 θ 1,θ 1 = 1 2 θ i+1,θ i+1. It follows that λ = ø i+1. As a corollary, the weights ϕ i ϕ i 1 ϕ 1 ø = ø i+1 ø i, for various values of i, are pairwise distinct. It follows that, if v V is a non-zero eigenvector under the action of T with weight ø, then v ϕ 1 v... ϕ i ϕ 1 v l i+1 V is a non-zero eigenvector with weight ø i+1. Let us prove another technical lemma before proceeding. 3

Lemma 3.2 Let i 1,i 2,...i t be integers, where 1 t < k. Then ϕ i1 ϕ i2 ϕ it ø = ϕ s ϕ s 1 ϕ 1 ø for some s t. More precisely, for s < k 1, the weight ϕ s ϕ s 1 ϕ 1 ø is fixed by any ϕ u unless u = s or u = s+1. Proof. Clearly the first equality follows by repeated use of the more precise statement. Let λ = ϕ s ϕ s 1 ϕ 1 ø = ø s+1 ø s. Then ϕ u λ,θ j = λ,ϕ u θ j = λ,θ j +c j λ,θ u for some constant c j. If u is neither equal to s nor s+1, then ø s+1 ø s,θ u = 0 and we conclude that ϕ u λ,θ j = λ,θ j for all j. It follows in particular that ϕ i ø = ø if i 2. An immediate induction then gives ϕ i ϕ i 1 ϕ 1 ø = ø (θ 1 +θ 2 +...+θ i ) which will be handy in the sequel. Wearenowreadytoprovethatø i ismaximalinl i V. Letusstartbyordering the weights in the W-orbit of ø. Lemma 3.3 Let v FWC. Then ø(v) > ϕ 1 ø(v) > ϕ 2 ϕ 1 ø(v) > > ϕ k 1 ϕ k 2 ϕ 1 ø(v) and ϕ k 1 ϕ k 2 ϕ 1 ø(v) > ϕø(v) for any other value of ϕø(v) with ϕ W. Proof. The first inequalities are trivial, and follow from the last expression given for ϕ i ϕ i 1 ϕ 1 ø. It is well-known ([1], 6.26) that ϕø(v) < ø(v) as soon as ϕø ø. Now, let ϕ W be such that ϕø(v) reaches its second highest value. Then ϕø ø so ϕø is not in ClFDWC, and therefore there is an u such that ϕø,θ u < 0. Compute then ϕ u ϕø(v) = ϕø(v) 2 ϕø,θ u θ u,θ u θ u(v) > ϕø(v). Bydefinition ofϕ, thisimplies ϕ u ϕø(v) = ø(v), sothat ϕ u ϕø = øandϕø = ϕ u ø. Bythelastlemma, ϕ u øcanonlybeϕ 1 ø,soweconcludethatϕ 1 ø(v)isthesecond highest value of ϕø(v) for ϕ W, and that it is only reached when ϕ 1 ϕ fixes ø. We would prove similarly that ϕ 2 ϕ 1 ø(v) is the third highest value of ϕø(v), which is only reached when ϕ 2 ϕ 1 ϕ fixes ø. The result follows by induction. In order to deal with the weights in V which are not in the W-orbit of ø, we shall need 1 the following. Lemma 3.4 Let M be an irreducible G-module with maximal weight λ. 1 such weights do exist, cf the 26-dimensional representation of F 4 which has two zero weights. 4

1. Let m M be an eigenvector for the weight α. For each positive root θ i, let X θi L(G) be an eigenvector for θ i. Then α = λ if and only if X θi m = 0 for each i. 2. All the weights occuring in M are of the form where n i is a nonnegative integer. λ n i θ i, Proof. Half of (1) is trivial: if m is an eigenvector for λ, then X θi m is an eigenvector for λ+θ i > λ, so it must be zero. We prove the converse and (2) at the same time. Supposethen thatx θi m = 0foreachpositiverootθ i. SinceM isirreducible, it is generated as an L(G)-module by m alone. Therefore M is spanned by elements of the form X s X s 1 X 1 m with X j L(G). Splitting L(G) according to the adjoint action of the maximal torus T, we may assume that each X j is an eigenvector associated to the root λ j, or to the zero weight. A vector as above is then an eigenvector in M with weight α + λ 1 +...λ s. We claim that M is in fact generated by elements as above for which each λ i is a negative root (or zero). Granting this, α is clearly maximal, so α = λ, and the other weights in M are as announced in (2). To prove the claim, we show by induction on s how to replace the generating vectors. The case s = 1 is our assumption, while the induction proceeds easily using X i X j = X j X i +[X i,x j ]. Theorem 1.1 now follows from this: Proposition 3.5 ø i+1 is the highest weight is l i+1 V (equivalently, it is the only maximal weight). Proof. By induction on i (the case i = 0 being given). So let λ = λ 1 +λ 2 +...+λ i+1 be a maximal weight occuring in l i+1 V, where each λ j is a weight in V. First case. Suppose to start with that for each λ j and each v FWC we have λ j (v) ϕ i ϕ i 1 ϕ 1 ø(v). By the preceding lemma, we have λ j = ø s n s θ s while ϕ i ϕ i 1 ϕ 1 ø = ø θ 1 θ 2... θ i. It follows that n s θ s (v) θ 1 (v)+...θ i (v) s for all v FWC, or even for v ClFWC. Applying this with v = ø s (and identifying L(G) with its dual) weconclude that eachn s is 0 or1when 1 s i (and zero otherwise). 5

So if λ j ø, let t denote the largest value of s such that n s = 1, and write λ j = ϕ t ϕ t 1 ϕ 1 ø+ s<tm s θ s, with m s = 1 n s. If there exists an eigenvector in V for this weight, then there is also one for the weight ( ) α = ø+ϕ 1 ϕ 2 ϕ t m s θ s. s<t Here ϕ 1 ϕ 2 ϕ t θ s = ϕ 1 ϕ s+1 θ s = θ s+1 (as already observed during the proof of 3.1). It follows that α > ø if any coefficient m s is nonzero. As this is impossible, we conclude that λ j = ϕ t ϕ t 1 ϕ 1 ø. It follows that t must be different for each j, and that λ = ø i+1. Second case. Suppose now that there is a v FWC and a j such that λ j (v) < ϕ i ϕ i 1 ϕ 1 ø(v), say j = i+1. We shall derive a contradiction (recall that λ is assumed to be maximal). Nowifλ i+1 isinthew-orbitofø,wedrawfrom3.3thatλ j (v) < ϕ i ϕ i 1 ϕ 1 ø(v) actually holds for all v FWC. Then by the induction hypothesis, we have λ 1 (v)+λ 2 (v)+...λ i (v) ø(v)+ϕ 1 ø(v)+...+ϕ i 1 ϕ i 2 ϕ 1 ø(v), so λ(v) < ø i+1 (v) and we are done. If on the other hand λ i+1 is not in the W-orbit of ø, we shall proceed quite differently. Let x V be an eigenvector associated to ø, and let y V be an eigenvector for λ i+1. Put m = x ϕ 1 x... ϕ i 1 ϕ i 2 ϕ 1 x y l i+1 V. Then m is non-zero, and is an eigenvector for a weight which is λ (using the induction hypothesis), so by maximality of λ, we see that m is an eigenvector for λ. Using lemma 3.4, we conclude that there exists a positive root θ j and an eigenvector X = X θj L(G) such that Xy 0. However by the same lemma, Xx = 0 and Xm = 0 (m being an eigenvector for the maximal weight of some irreducible submodule of l i+1 ). Write 0 = Xm = X(x ϕ 1 x... ϕ i 1 ϕ i 2 ϕ 1 x) y ±x ϕ 1 x... ϕ i 1 ϕ i 2 ϕ 1 x Xy The first term here is 0 by induction, so we conclude that Xy is in the linear span of x,ϕ 1 x,...,ϕ i 1 ϕ i 2 ϕ 1 x. Since Xy is a (nonzero) eigenvector for the action of T, it follows that Xy is in fact (a scalar multiple of) one of these. In conclusion, λ i+1 + θ j = ϕ s ϕ s 1 ϕ 1 ø for some s with 1 s i 1, or equivalently λ i+1 = (ø s+1 ø s ) θ j. Now, the contradiction will come from the fact that, by maximality of λ, we must have λ i+1 ClFDWC, so that ϕ j λ i+1 λ i+1. This amounts to ø s+1 ø s,θ j θ j,θ j 1. For this to happen, we would need a positive root θ j which, when written as a sum of positive simple roots θ j = n t θ t, has n s+1 n s 2. This never 6

happens. You may either inspect the existing root systems, or fill in the details ofthe followingsketch: let S be theset ofpositiverootssuchthat n s+1 n s 1 for all 1 s+1 k 1. Then S contains the positive simple roots, and is stable under each relexion ϕ u for u r. Hence S is stable under W and consequently, it is equal to the set of all positive roots. Corollary 3.6 We may replace V i by l i V in the statement of Weyl s theorem, for 1 i k. More generally, if Top(V i ) = V i for each i, then R(G) = Z[V 1,V 2,...,V r ] and moreover any irreducible representation of G is of the form Top(V 1 n 1 V n 2 V 2 r n r ). Proof. Run through a proof of Weyl s theorem and replace V i by V i. As this is short enough to write out, we give the details. For each weight λ ClFDWC, let S(λ) denote the elementary symmetric sum S(λ) = e 2iπα. α Wλ These form a Z-basis for R(G) = R(T) W (viewed as the character ring rather than the representation ring). Now, if λ = n 1 ø 1 +...+n r ø r, we clearly have χ(v 1 n 1 V r n r ) = S(λ)+ m i S(λ i ) λi<λ where χ means the character (use [1], 6.36). If we suppose by induction that each S(λ i ) can be written in the form χ(p(v 1,...,V r )) (with P a polynomial), then the same can be said of S(λ). As a result, V 1,V 2,...,V r generate R(G) as a ring. They are clearly algebraically independant(for example, consider the transcendance degree of R(G)). The rest is easy. 4. Applications All the information on exceptional Lie groups used below can be found in Adams s book[3]. The representation ring of F 4. This group has the following Dynkin diagram: 1 2 3 4 short short long long The groupf 4 may be defined as the groupof automorphisms of the (Jordan) algebra J of 3 3 hermitian matrices over the octonions, the multiplication being A,B 1 2 (AB +BA). The real dimension of J is 27, and F 4 acts on the subspace of matrices with trace 0. Complexifying, this affords a 26-dimensional 7

representationu off 4. This U turnsout to be irreducible andits highest weight corresponds to the node 1. The (complexified) adjoint representation Ad, on the other hand, corresponds to the node 4. It follows from the main theorem in this paper that The E family. R(F 4 ) = Z[Ad,l 2 Ad,U,l 2 U]. Adams originally considered the case of E 8 in [2]. This group has the following Dynkin diagram: α β Adams concludes that γ R(E 8 ) = Z[α,l 2 α,l 3 α,l 4 α,β,l 2 β,γ,δ] where δ can be taken to be either l 5 α, or l 3 β, or l 2 γ. It is easy to see that the representation α is the adjoint representation, and Adams in loc cit gives some concrete information on β and γ (eg, the way they sit in α α). I do not know a direct, concrete definition of β and γ. Similarly, for E 7 we have the following diagram: So that c a b R(E 7 ) = Z[a,l 2 a,l 3 a,b,l 2 b,c,d] where d can be taken to be either l 4 a, or l 3 b, or l 2 c. Here however, the adjoint representation does not correspond to any of a, b, c. We can identify nevertheless the representation a as the complexification of the 56-dimensional real representation W of E 7 which can be used to define E 7 as the group of linear automorphisms of W preserving both a quadratic form and a quartic form. The situation for E 6 is analogous. The Dynkin diagram has three branches, so R(E 6 ) is generated by 3 elements as a λ-ring. I am not aware of any explicit description of the three fundamental representations involved. They are neither given by the adjoint representation, nor by the representation W used to define E 6 as a group of maps. 8

References [1] J. F. Adams, Lectures on Lie groups, W. A. Benjamin, Inc., New York- Amsterdam, 1969. [2], The fundamental representations of E 8, Contemp. Math, 37 (1985). [3], Lectures on exceptional Lie groups, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, IL, 1996. [4] M. F. Atiyah and D. O. Tall, Group representations, λ-rings and the J-homomorphism, Topology, 8 (1969), pp. 253 297. 9