Lie Algebra Class
Lie Algebra Class
Brooks Roberts
University of Idaho
1 Basic concepts 1
1.1 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 The definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.4 Some important examples . . . . . . . . . . . . . . . . . . . . . . 3
1.5 The adjoint homomorphism . . . . . . . . . . . . . . . . . . . . . 5
5 Cartan’s criteria 33
5.1 The Jordan-Chevalley decomposition . . . . . . . . . . . . . . . . 33
5.2 Cartan’s first criterion: solvability . . . . . . . . . . . . . . . . . 34
5.3 Cartan’s second criterion: semi-simplicity . . . . . . . . . . . . . 40
5.4 Simple Lie algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.5 Jordan decomposition . . . . . . . . . . . . . . . . . . . . . . . . 45
6 Weyl’s theorem 51
6.1 The Casmir operator . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2 Proof of Weyl’s theorem . . . . . . . . . . . . . . . . . . . . . . . 55
6.3 An application to the Jordan decomposition . . . . . . . . . . . . 58
iii
iv CONTENTS
8 Root systems 83
8.1 The definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.2 Root systems from Lie algebras . . . . . . . . . . . . . . . . . . . 84
8.3 Basic theory of root systems . . . . . . . . . . . . . . . . . . . . . 85
8.4 Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
8.5 Weyl chambers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
8.6 More facts about roots . . . . . . . . . . . . . . . . . . . . . . . . 101
8.7 The Weyl group . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
8.8 Irreducible root systems . . . . . . . . . . . . . . . . . . . . . . . 114
Basic concepts
1.1 References
The main reference for this course is the book Introduction to Lie Algebras, by
Karin Erdmann and Mark J. Wildon; this is reference [4]. Another important
reference is the book [6], Introduction to Lie Algebras and Representation The-
ory, by James E. Humphreys. The best references for Lie theory are the three
volumes [1], Lie Groups and Lie Algebras, Chapters 1-3, [2], Lie Groups and Lie
Algebras, Chapters 4-6, and [3], Lie Groups and Lie Algebras, Chapters 7-9, all
by Nicolas Bourbaki.
1.2 Motivation
Briefly, Lie algebras have to do with the algebra of derivatives in settings where
there is a lot of symmetry. As a consequence, Lie algebras appear in various
parts of advanced mathematics. The nexus of these applications is the theory
of symmetric spaces. Symmetric spaces are rich objects whose theory has com-
ponents from geometry, analysis, algebra, and number theory. With these short
remarks in mind, in this course we will begin without any more motivation,
and start with the definition of a Lie algebra. For now, rather than be con-
cerned about advanced applications, the student should instead exercise critical
thinking as basic concepts are introduced.
[·, ·] : L × L −→ L
1
2 CHAPTER 1. BASIC CONCEPTS
2. [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0 for all x, y, z ∈ L.
The map [·, ·] is called the Lie bracket of L. The second property is called the
Jacobi identity.
0 = [x + y, x + y]
= [x, x] + [x, y] + [y, x] + [y, y]
= [x, y] + [y, x],
Proof. Let y ∈ L and x ∈ Z(L). If z ∈ L, then [[y, x], z] = −[[x, y], z] = 0. This
implies that [y, x] ∈ Z(L).
Proof. Let y ∈ ker(T ) and x ∈ L1 . Then T ([x, y]) = [T (x), T (y)] = [T (x), 0] =
0, so that [x, y] ∈ ker(T ).
1.4. SOME IMPORTANT EXAMPLES 3
[x, y] = xy − yx,
so that [x, y] is just the commutator of x and y. With this definition of a Lie
bracket, the F -vector space A is a Lie algebra.
Proof. It is easy to verify that [·, ·] is F -bilinear and that property 1 of the
definition of a Lie algebra is satisfied. We need to prove that the Jacobi identity
is satisfied. Let x, y, z ∈ A. Then
[x, [y, z]] + [y, [z, x]] + [z, [x, y]] = x(yz − zy) − (yz − zy)x
+ y(zx − xz) + (zx − xz)y
+ z(xy − yx) + (xy − yx)z
= xyz − xzy − yzx + zyx
+ yzx − yxz − zxy + xzy
+ zxy − zyx − xyz + yxz
= 0.
then
1
[x, [y, z]] = x(yz − zy) = xyz − xzy = xyz − = xyz
1
and
1 1
[[x, y], z] = (xy − yx)z = xyz − yxz = xyz − = xyz − .
1
We have:
[e, f ] = h, [e, h] = −2e, [f, h] = 2f.
Proposition 1.4.3. Let n be a non-negative integer, and let S ∈ gl(n, F ). Let
then we write
so(n, F ) = so(2` + 1, F ) = glS (n, F ).
Also, if n = 2` is even and
1`
S= ,
−1`
then we write
sp(n, F ) = sp(2`, F ) = glS (n, F ).
If the F -vector space V is actually an algebra R over F , then the Lie algebra
gl(R) admits a natural subalgebra. Note that in the next proposition we do not
assume that R is associative.
ad : L −→ gl(L)
by
ad(x) (y) = [x, y]
for x, y ∈ L. Then ad is a Lie algebra homomorphism. Moreover, the kernel of
ad is Z(L), and the image of ad lies in Der(L). We refer to ad as the adjoint
homomorphism.
6 CHAPTER 1. BASIC CONCEPTS
Also,
[ad(x1 ), ad(x2 )] (y) = ad(x1 ) ◦ ad(x2 ) (y) − ad(x2 ) ◦ ad(x1 ) (y)
= ad(x1 ) [x2 , y] − ad(x2 ) [x1 , y]
= [x1 , [x2 , y]] − [x2 , [x1 , y]].
It follows that
ad([x1 , x2 ]) (y) − [ad(x1 ), ad(x2 )] (y)
= [[x1 , x2 ], y] − [x1 , [x2 , y]] + [x2 , [x1 , y]]
= −[y, [x1 , x2 ]] − [x1 , [x2 , y]] − [x2 , [y, x1 ]]
=0
and
Therefore,
again by the Jacobi identity. This proves that the image of ad lies in Der(L).
The previous proposition shows that elements of a Lie algebra can always
be thought of as derivations of an algebra. It turns out that if L is a finite-
dimensional semi-simple Lie algebra over the complex numbers C, then the
image of the adjoint homomorphism is Der(L).
Chapter 2
2.1 Solvability
Proposition 2.1.1. Let L be a Lie algebra over F , and let I and J be ideals of
L. Define [I, J] to be the F -linear span of all the brackets [x, y] for x ∈ I and
y ∈ J. The F -vector subspace [I, J] of L is an ideal of L.
Proof. Let x ∈ L, y ∈ I and z ∈ J. We need to prove that [x, [y, z]] ∈ [I, J].
We have
Proof. This follows immediately from [e, f ] = h, [e, h] = −2e, [f, h] = 2f.
Proposition 2.1.3. Let L be a Lie algebra over F . The quotient algebra L/L0
is abelian.
Proof. This follows immediately from the definition of the derived algebra.
7
8 CHAPTER 2. SOLVABLE AND NILPOTENT LIE ALGEBRAS
Each term of the sequence is actually an ideal of L; also, the successive quotients
are abelian. To improve the notation, we will write
L(0) = L,
L(1) = L0 ,
L(2) = (L0 )0 ,
···
L (k+1)
= (Lk )0
··· .
We have then
L = L(0) ⊃ L(1) ⊃ L(2) ⊃ · · · .
This is called the derived series of L. We say that L is solvable if L(k) = 0
for some non-negative integer k.
Proposition 2.1.4. Let L be a Lie algebra over F . Then L is solvable if and
only if there exists a sequence I0 , I1 , I2 , . . . , Im of ideals of L such that
L = I0 ⊃ I1 ⊃ I2 ⊃ · · · ⊃ Im−1 ⊃ Im = 0
(k) (k)
= [T (L1 ), T (L1 )]
(k) (k)
= [L2 , L2 ]
(k+1)
= L2 .
Lemma 2.1.6. Let L be a Lie algebra over F . We have L(k+j) = (L(k) )(j) for
all non-negative integers k and j.
Proof. Fix a non-negative integer k. We will prove that L(k+j) = (L(k) )(j) by
induction on j. If j = 0, then L(k+j) = L(k) = (L(k) )(0) = (L(k) )(j) . Assume
that the statement holds for j; we will prove that it holds for j + 1. By the
induction hypothesis,
Also,
Lemma 2.1.7. Let L be a Lie algebra over F . Let I be an ideal of L. The Lie
algebra L is solvable if and only if I and L/I are solvable.
Proof. If L is solvable then I is solvable because I (k) ⊂ L(k) for all non-negative
integers; also, L/I is solvable by Lemma 2.1.5. Assume that I and L/I are
solvable. Since L/I is solvable, there exists a non-negative integer k such that
(L/I)(k) = 0. This implies that L(k) + I = I, so that L(k) ⊂ I. Since I is
solvable, there exists an non-negative integer j such that I (j) = 0. It follows
that (L(k) )(j) ⊂ I (j) = 0. Since L(k+j) = (L(k) )(j) by Lemma 2.1.6, we conclude
that L is solvable.
Lemma 2.1.8. Let L be a Lie algebra over F , and let I and J be solvable ideals
of L. Then I + J is solvable.
I + J ⊃ J ⊃ 0.
The following theorem will not be proven now, but is an important reduction
in the structure of Lie algebras.
so that
It follows that
a
x ∈ Z(gl(2, F )) = { : a ∈ F },
a
so that I ⊂ Z(gl(2, F )). Since Z(gl(2, F )) ⊂ I = rad(gl(2, F )), the proposition
is proven.
Proposition 2.1.14. Let b(2, F ) be the F -subspace of gl(2, F ) consisting of
upper triangular matrices. Then b(2, F ) is a Lie subalgebra of gl(2, F ), and
b(2, F ) is solvable.
Proof. Let
a1 b1 a2 b2
x1 = , x2 =
d1 d2
be in b(2, F ). Then
b1 d2 − b2 d1 + a1 b2 − a2 b1 ∗
[x1 , x2 ] = ∈ .
From this formula it follows that b(2, F ) is a Lie subalgebra of gl(2, F ). More-
over, it is clear that
∗
b(2, F )(1) = ,
b(2, F )(2) = 0,
More generally, one has the following theorem, the proof of which will be
omitted:
Theorem 2.1.16. Let b(n, F ) be the Lie algebra over F consisting of all upper
triangular n × n matrices with entries from F . Then b(n, F ) is solvable.
2.2 Nilpotency
There is a stronger property than solvability. Let L be a Lie algebra over F .
We define the lower central series of L to the sequence of ideals:
L0 = L, L1 = L0 , , Lk = [L, Lk−1 ], k ≥ 2.
···
∗
b(2, F )k = , k ≥ 1.
On the other hand, the Lie algebra n(2, F ) of strictly upper triangular 2 × 2
over F is nilpotent:
n(2, F )k = 0, k ≥ 1.
Proof. The first assertion is clear. Assume that L/Z(L) is nilpotent. We claim
that (L/Z(L))k = (Lk + Z(L))/Z(L) for all non-negative integers k. This
statement is clear if k = 0. Assume that the statement holds for k; we will
prove that it holds for k + 1. Now
Theorem 2.2.2. Let n(n, F ) be the Lie algebra over F consisting of all strictly
upper triangular n × n matrices with entries from F . Then n(n, F ) is nilpotent.
14 CHAPTER 2. SOLVABLE AND NILPOTENT LIE ALGEBRAS
Chapter 3
Theorem 3.1.2 (Lie’s Theorem). Assume that F has characteristic zero and is
algebraically closed. Let V be a finite-dimensional vector space over F . Suppose
that L is a solvable Lie subalgebra of gl(V ). Then there exists a basis for V such
that in this basis every element of L is an upper triangular matrix.
15
16 CHAPTER 3. THE THEOREMS OF ENGEL AND LIE
a(yw) = (ay)w
= ([a, y] + ya)w
= [a, y]w + yaw
= λ([a, y])w + λ(a)(yw).
Since w 6= 0, this calculation shows that we must prove that λ([a, y]) = 0.
To prove this, we consider the subspace U of V spanned by the vectors
w, yw, y 2 w, . . .
w, yw, y 2 w, . . . , y m w
are linearly independent. This set is a basis for U . We claim that for all z ∈ A
we have zU ⊂ U , and that moreover the matrix of z with respect to the basis
w, yw, y 2 w, . . . , y m w has the form
λ(z) ∗ ... ∗
λ(z) . . . ∗
.. .
. ..
.
λ(z)
This proves the claim for the second column. Assume that the claim has been
proven for the first k columns with k ≥ 2; we will prove it for the k + 1 column.
Let z ∈ A. Then
By the induction hypothesis, since [z, y] ∈ A, the vector u1 = [z, y](y k−1 w) is
in the span of w, yw, y 2 w, . . . , y k−1 w. Also, by the induction hypothesis, there
exists u2 in the span of w, yw, y 2 w, . . . , y k−2 w such that
It follows that
Since the vector u1 + yu2 is in the span of w, yw, y 2 w, . . . , y k−1 w, our claim
follows.
Now we can complete the proof. We recall that we are trying to prove that
λ([a, y]) = 0. Let z = [a, y]; then z ∈ A. By the last paragraph, z acts on U , and
we have that the trace of the action of z on U is (m + 1)λ(z) = (m + 1)λ([a, y]).
On the other hand, z = [a, y] = ay−ya, and a and y both act on U . This implies
that trace of the action of z on U is zero. We conclude that λ([a, y]) = 0.
Proof. Since our field is algebraically closed, it will suffice to prove that the only
eigenvalue of [x, y] is zero. Let c be an eigenvalue of [x, y].
Let
L = F x + F y + F [x, y].
Since [x, [x, y]] = [y, [x, y]] = 0, the vector space L is a Lie subalgebra of gl(V ).
Let
A = F [x, y].
Evidently, A is an ideal of L; in fact [z, a] = 0 for all z ∈ L. Let λ : A → F be
the linear functional such that λ([x, y]) = c. Then the weight space Vλ is
By the Lemma 3.2.1, the Invariance Lemma, Vλ is mapped by L into itself. Pick
a basis for Vλ , and write the action of x and y on Vλ in this basis as matrices X
and Y , respectively. On the one hand, we have tr[X, Y ] = 0, as usual. On the
other hand, [X, Y ] acts by c on Vλ , which implies that tr[X, Y ] = (dim Vλ )c. It
follows that c = 0.
ad(x)n+1 (y)
= ad(x)(ad(x)n (y))
n
X n
= [x, (−1)k xn−k yxk ]
k
k=0
n n
X n X n
= (−1)k xn−k+1 yxk − (−1)k xn−k yxk+1
k k
k=0 k=0
n n+1
X
X n n
= (−1)k xn−k+1 yxk + (−1)k xn−k+1 yxk
k k−1
k=0 k=1
n
X n n
= xn+1 y + (−1)n+1 yxn+1 + ( + )(−1)k xn−k+1 yxk
k k−1
k=1
n
n+1 n+1 n+1
X n+1
=x y + (−1) yx + (−1)k xn−k+1 yxk
k
k=1
n+1
X
n+1
(−1)k xn+1−k yxk .
k
k=0
xv = (a + cy)v = av + cyv = 0 + 0 = 0.
This proves that the assertion of the lemma holds for L. By induction, the
lemma is proven.
Proof of Theorem 3.1.1, Engel’s Theorem. We prove this theorem by induction
on dim V . If dim V = 0, then there is nothing to prove. Assume that dim V ≥ 1,
and that the theorem holds for all Lie algebras satisfying the hypothesis of the
theorem that have dimension strictly less than dim V .
By Lemma 3.3.2, there exists a non-zero vector v ∈ V such that xv = 0 for
all x ∈ L. Let U = F v. Define V̄ = V /U . We consider the natural map
ϕ : L −→ gl(V̄ )
v1 + U, . . . , vn−1 + U
of V̄ such that the elements of ϕ(L) are strictly upper triangular in this basis.
The vectors
v, v1 , . . . , vn−1
form an ordered basis for V . It is evident that the elements of L are strictly
upper triangular in this basis.
Vλ = {w ∈ V : aw = λ(a)w for a ∈ A}
is non-zero. By the Invariance Lemma, Lemma 3.2.1, the Lie algebra L maps
the weight space Vλ to itself. Since dim A = dim L − 1, there exists z ∈ L such
that L = A + F z. Consider the action of z on Vλ . Since F is algebraically
closed, there exists a non-zero vector w ∈ Vλ that is eigenvector for z; let d ∈ F
be the eigenvalue. We claim that w is an eigenvector for every element of L.
Let x ∈ L, and write x = a + cz for some a ∈ A and c ∈ F . Then
4.1 Representations
Let L be a Lie algebra over F . A representation consists of a pair (ϕ, V ), where
V is a vector space over F and ϕ : L → gl(V ) is a Lie algebra homomorphism.
Evidently, if V is a vector space over F , and ϕ : L → gl(V ) is a linear map,
then the pair (ϕ, V ) is a representation of L if and only if
ad(x)y = [x, y]
for x, y ∈ L.
23
24 CHAPTER 4. SOME REPRESENTATION THEORY
This proves our claim. Applying Theorem 4.2.2, Schur’s Lemma, to ϕ(z), we
see that there exists a unique c ∈ F such that ϕ(z)v = cv for v ∈ V . We now
define λ(z) = c. It is straightforward to verify that λ is a linear map.
4.3. REPRESENTATIONS OF SL(2) 25
Lemma 4.3.1. Let V be a vector space over F , and let ϕ : sl(2, F ) → gl(V ) be
an F -linear map. Define
∂p
Ep = X ,
∂Y
∂p
Fp = Y ,
∂X
∂p ∂p
Hp = X −Y .
∂X ∂Y
26 CHAPTER 4. SOME REPRESENTATION THEORY
pk = X d−k Y k .
To summarize:
Proposition 4.3.3. Let the notation be as in Lemma 4.3.2. The linear map
ϕ : sl(2, F ) → gl(Vd ) determined by setting ϕ(e) = E, ϕ(f ) = F , and ϕ(h) = H
is a Lie algebra homomorphism, so that (ϕ, Vd ) is a representation of sl(2, F ).
H · pk = (d − 2k)pk .
with distinct roots. In particular, H|W has an eigenvector. This implies that
for some k ∈ {0, . . . , d} we have pk ∈ W . By applying powers of E and F we
find that all the vectors v0 , . . . , vd are contained in W . Hence, W = Vd and Vd
is irreducible.
Lemma 4.3.5. Let V be a representation of sl(2, F ). Assume that v is an
eigenvector for h with eigenvalue λ ∈ F . Either ev = 0, or ev is non-zero and
ev is an eigenvector for h such that
h(ev) = (λ + 2)ev.
u, eu, e2 u, ....
v, f v, f 2 v, ....
v, f v, f 2 v, ..., f dv
e(f j v) = ef (f j−1 v)
= (f e + [e, f ])(f j−1 v)
= (f e + h)(f j−1 v)
= f e(f j−1 v) + h(f j−1 v).
The vector f e(f j−1 v) is in W by the induction hypothesis, and the vector
h(f j−1 v) is in W because W is invariant under h. This proves our claim by
induction, so that W is an sl(2, F )-subspace of V . Since V is irreducible and W is
non-zero, we obtain V = W . In particular, we see that dim V = dim W = d + 1.
Next, we will prove that λ = d. To prove this, consider the matrix of h with
respect to the basis
v, f v, f 2 v, . . . , f d v
of V = W . The matrix of h with respect to this basis is:
λ
λ−2
λ − 4 .
. ..
λ − 2d
30 CHAPTER 4. SOME REPRESENTATION THEORY
It follows that
trace(h) = trace([e, f ])
= trace(ef − f e)
= trace(ef ) − trace(f e)
= trace(ef ) − trace(ef )
= 0.
T (f k v) = F k X d
for k ∈ {0, . . . , d}. This map is evidently an isomorphism. To complete the proof
we need to prove that T is an sl(2, F )-map. First we prove that T (f w) = F T (w)
for w ∈ V . To prove this it suffices to prove that this holds for w = f k v for
k ∈ {0, . . . , d}. If k ∈ {0, . . . , d − 1}, then
T (f (f k v)) = T (f k+1 v)
= F k+1 X d
= F T (f k v).
If k = d, then
T (f (f d v)) = T (0)
=0
= F d+1 X d
= F T (f d v).
Next we prove that T (hw) = HT (w) for w ∈ V . Again, it suffices to prove that
this holds for w = f k v for k ∈ {0, . . . , d}. Let k ∈ {0, . . . , d}. Then
Finally, we need to prove that T (ew) = ET (w) for w = f k v for k ∈ {0, . . . , d}.
We will prove this by induction on k. If k = 0, this clear because T (ef 0 v) =
T (0) = 0 = EX d = ET (f 0 v). Assume that k ∈ {1, . . . , d} and T (e(f j v)) =
ET (f j v) for j ∈ {0, . . . , k − 1}; we will prove that T (e(f k v)) = ET (f k v). Now
Cartan’s criteria
xs ys = xs sy (y)
= sy (y)xs
= ys xs .
33
34 CHAPTER 5. CARTAN’S CRITERIA
It follows that ad(xs ) is diagonalizable. To see that ad(xs ) and ad(xn ) commute,
let y ∈ gl(V ). Then
ad(xs )ad(xn ) (y) = ad(xs ) ad(xn )(y)
= ad(xs ) [xn , y]
= [xs , [xn , y]]
= [xs , xn y − yxn ]
= xs (xn y − yxn ) − (xn y − yxn )xs
= xs xn y − xs yxn − xn yxs + yxn xs
= xn xs y − xs yxn − xn yxs + yxs xn
= xn (xs y − yxs ) − (xs y − yxs )xn
= [xn , xs y − yxs ]
= [xn , [xs , y]]
= ad(xn )ad(xs ) (y).
Let r(X) ∈ F [X] be the Langrange interpolation polynomial for this set. Then
r(X) does not have a contant term because r(0) = 0. Also,
r(λi − λj ) = f (λi − λj )
r(ad(s)) = ad(y).
36 CHAPTER 5. CARTAN’S CRITERIA
By Lemma 5.1.3 we have ad(s) = ad(x)s . Hence, by Theorem 5.1.1, there exists
a polynomial p(X) ∈ F [X] with no constant term such that
ad(s) = p(ad(x)).
We now have
ad(y) = p(r(ad(x)).
Now, because x ∈ M , we have ad(x)(B) ⊂ A. We claim that this implies that
ad(x)k (B) ⊂ A for all positive integers k. We prove this claim by induction on
k. The claim holds for k = 1. Assume it holds for k. Then
This proves the claim. Since ad(y) is a polynomial in ad(x) with constant term
we conclude that ad(y)(B) ⊂ A. This implies that y ∈ M , by definition. By
our assumption on x we have tr(xy) = 0. This means that:
This completes the proof by induction. It follows that L(k) = 0 if and only if
(k)
LK = 0. Hence, L is solvable if and only if LK is solvable.
Similarly, L is nilpotent if and only if LK is nilpotent.
Proof. Assume that tr(xy) = 0 for all x ∈ L0 and y ∈ L. We need to prove that
L is solvable.
We will first prove that we may assume that F is algebraically closed. Let
K = F̄ , the algebraic closure of F . Define VK = K ⊗F V . Then VK is a
K-vector space, and dimK VK = dimF V . There is a natural inclusion
id
K −−−−→ K
∼
38 CHAPTER 5. CARTAN’S CRITERIA
0 ⊂ L0 ⊂ L.
m
X
=− tr([y, zi ]xi ).
i=1
K ⊗ gl(L) ,→ gl(LK )
And
ad(a ⊗ x)ad(b ⊗ y) (c ⊗ z) = ad(a ⊗ x) ad(b ⊗ y)(c ⊗ z)
= ad(a ⊗ x) [b ⊗ y, c ⊗ z]
= [a ⊗ x, [b ⊗ y, c ⊗ z]]
= [a ⊗ x, bc ⊗ [y, z]]
= abc ⊗ [x, [y, z]].
It follows that
ab ⊗ ad(x)ad(y) = ad(a ⊗ x)ad(b ⊗ y).
The diagram
∼
K ⊗ gl(L) −−−−→ gl(LK )
id⊗try try
id
K −−−−→ K
commutes. Hence, we obtain
ab · tr(ad(x)ad(y)) = tr ad(a ⊗ x)ad(b ⊗ y) .
40 CHAPTER 5. CARTAN’S CRITERIA
Since ad(x1 ) and ad(x2 ) are upper triangular, a calculation shows that the up-
per triangular matrix [ad(x1 ), ad(x2 )] is strictly upper triangular. This implies
that all the elements of ad(L0 ) are strictly upper triangular matrices. Another
calculation now shows that ad(x)ad(y) is strictly upper triangular for x ∈ L0
and y ∈ L; therefore, tr(ad(x)ad(y)) = 0 for x ∈ L0 and y ∈ L.
Now assume that tr(ad(x)ad(y)) = 0 for x ∈ L0 and y ∈ L. Consider ad(L).
By Lemma 2.1.5, ad(L0 ) = ad(L)0 . Therefore, our hypothesis and Lemma 5.2.3
imply that ad(L) is solvable. Now ad(L) ∼ = L/Z(L) as Lie algebras. Hence,
L/Z(L) is solvable. Since Z(L) is solvable, we conclude from Lemma 2.1.7 that
L is solvable.
κ : L × L −→ F
by
κ(x, y) = tr(ad(x)ad(y))
for x, y ∈ L. We refer to κ as the Killing form on L.
for x, y, z ∈ L.
Proof. The linearity of ad and tr imply that kappa is bilinear. The Killing form
is symmetric because in general tr(AB) = tr(BA) for A and B linear operators
on a finite-dimensional vector space. Finally, let x, y, z ∈ L. Then
= tr(ad(x)[ad(y), ad(z)])
= tr(ad(x)ad([y, z]))
= κ(x, [y, z]).
Proof. Fix a F -vector space basis for I, and extend this to a basis for L. Let
x ∈ I. Then because I is an ideal, we have ad(x)L ⊂ I. It follows that the
matrix of ad(x) in our basis for L has the form
M (x) ∗
ad(x) =
0 0
where M (x) is the matrix of ad(x)|I in our chosen basis for I. Let y ∈ I. Then
(ad(a)ad(x)ad(a))(y) = (ad(a)(ad(x)ad(a))(y))
= [a, (ad(x)ad(a))(y)]
= [a, [x, ad(a)(y)]]
= [a, [x, [a, y]]].
Since A is an ideal of L we have [a, y] ∈ A, and hence also [x, [a, y]] ∈ A. Since A
is abelian, this implies that [a, [x, [a, y]]] = 0. It follows that ad(a)ad(x)ad(a) =
0 and thus (ad(x)ad(a))2 = 0. Since nilpotent operators have trivial traces, we
obtain tr(ad(a)ad(x)) = 0. Thus, κ(a, x) = 0. Because x ∈ L was arbitrary, we
have a ∈ L⊥ = 0. Thus, A = 0, a contradiction.
If b is non-degenerate, then
V −→ V ∨
Proof. By Lemma 5.4.1 and Lemma 5.4.2, to prove that L = I ⊕I ⊥ it will suffice
to prove that I ∩ I ⊥ = 0. Let J = I ∩ I ⊥ . Then J is an ideal of L. By Lemma
5.3.2, we have κJ (J, J) = 0. In particular, κJ (J, J 0 ) = 0. By Theorem 5.2.4,
Cartan’s first criterion, the Lie algebra J is solvable. Since L is semi-simple, we
get J = 0, as desired.
By Theorem 5.3.4, Cartan’s second criterion, to prove that I is semi-simple,
it will suffice to prove that if x ∈ I and κI (x, y) = 0 for all y ∈ I, then x = 0.
Assume that x ∈ I is such that κI (x, y) = 0 for all y ∈ I. By Lemma 5.3.2,
κ(x, y) = 0 for all y ∈ I. Let z ∈ L. By the first paragraph, we may write
z = z1 + z2 with z1 ∈ I and z2 ∈ I ⊥ . We have κ(x, z) = κ(x, z1 ) + κ(x, z2 ). Now
κ(x, z1 ) = 0 because z1 ∈ I and the assumption on x, and κ(x, z2 ) = 0 because
x ∈ I and z2 ∈ I ⊥ . It follows that κ(x, z) = 0. Since z ∈ L was arbitrary, we
obtain x ∈ L⊥ . By Theorem 5.3.4, Cartan’s second criterion, L⊥ = 0. Hence,
x = 0.
I = I1 ⊕ · · · ⊕ It .
Proof. Via induction on dim L, we will prove the assertion that if L is semi-
simple, then there exist simple ideals of L as in the theorem. The assertion is
trivially true when dim L = 0, because in this case L cannot be semi-simple.
Assume that the assertion holds for all Lie algebras over F with dimension less
than dim L; we will prove the assertion for L. Assume that L is semi-simple.
Let I be an ideal of L with the smallest possible non-zero dimension. Assume
that dim I = dim L, i.e., I = L. Then certainly L has no ideals other than 0 and
L. Moreover, L is not abelian because rad(L) = 0. It follows that L is simple.
Assume that dim I < dim L. By Lemma 5.4.3 we have L = I ⊕I ⊥ , and I and I ⊥
are semi-simple Lie algebras over F with dim I < dim L and dim I ⊥ < dim L.
By induction, there exist simple ideals I1 , . . . , Ir of I and simple ideals J1 , . . . , Js
of I ⊥ such that
I = I1 ⊕ · · · ⊕ Ir and I ⊥ = J1 ⊕ · · · ⊕ Js .
We have
L = I1 ⊕ · · · ⊕ Ir ⊕ J1 ⊕ · · · ⊕ Js
as F -vector spaces. It is easy to check that I1 , . . . , Ir , J1 , . . . , Js are ideals of L.
The assertion follows now by induction.
Next, assume that there exist simple ideals of L as in the statement of the
theorem. Let x, y, z ∈ L. Write x = x1 + · · · + xt , y = y1 + · · · + yt , and
z = z1 + · · · + zt with xi , yi , zi ∈ Ii for i ∈ {1, . . . , t}. We have
t
X
= (ad(xi )ad(yi ))(zi ).
i=1
It follows that
ad(x1 )ad(y1 )
ad(x)ad(y) =
.. .
.
ad(xt )ad(yt )
for a, b ∈ R.
Proposition 5.5.1. Let F be a field of characteristic zero. Let L be a semi-
simple finite-dimensional Lie algebra over F . Then the ad homomorphism is an
isomorphism of L onto Der(L):
∼
ad : L −→ Der(L).
where the last equality follows from Lemma 5.3.2. Since K is semi-simple we
must have x = 0 by Theorem 5.3.4, Cartan’s second criterion. Therefore, K ∩
K ⊥ = 0. Now since K and K ⊥ are both ideals of Der(L) we have [K, K ⊥ ] ⊂ K
and [K, K ⊥ ] ⊂ K ⊥ , so that [K, K ⊥ ] ⊂ K ∩ K ⊥ . Thus, [K, K ⊥ ] = 0. Let
D ∈ K ⊥ and x ∈ L. Then [D, ad(x)]
= 0. From above, we also have [D, ad(x)] =
ad D(x) . Therefore, ad D(x) = 0. Since ad is injective, we get D(x) = 0.
Since x ∈ L was arbitrary, we obtain D = 0. Thus, K ⊥ = 0. Now by Lemma
5.4.1 we have dim K + dim K ⊥ ≥ dim Der(L); therefore, dim K = dim Der(L)
so that K = Der(L).
We recall the following theorem from linear algebra.
Theorem 5.5.2 (Generalized eigenvalue decomposition). Assume that F has
characteristic zero and is algebraically closed. Let V be a finite-dimensional
vector space and let T ∈ gl(V ). If λ ∈ F , then define Vλ (T ) to be the subset of
v ∈ V such that there exists a non-negative integer such that (T − λ1V )k v = 0.
For λ ∈ F , Vλ (T ) is an F -subspace of V that is mapped to itself by T . We have
M
V = Vλ (T ).
λ∈F
(X − λ1 )n1 · · · (X − λt )nt
where the λi ∈ F are pairwise distinct for i ∈ {1, . . . , t}, and n1 , . . . , nt are
positive integers such that n1 + · · · + nt = dim V . Define E(T ) = {λ1 , . . . , λt },
the set of eigenvalues of T . For λ ∈ F we have Vλ (T ) 6= 0 if and only if
λ ∈ E(T ), and dim Vλi = ni for i ∈ {1, . . . , t}. Let T = s + n be the Jordan-
Chevalley decomposition of T , with s diagonalizable and n nilpotent. The set of
eigenvalues for T is the same as the set of eigenvalues for s, and Vλ (s) = Vλ (T )
for λ ∈ E(T ) = E(s). Moreover, for every λ ∈ E(T ) = E(s), Vλ (s) is the usual
λ-eigenspace for s.
5.5. JORDAN DECOMPOSITION 47
n+1
X n + 1
(D − λ1L )k x, (D − µ1L )n+1−k y .
=
k
k=0
To prove this, let x ∈ Lλ (D) and y ∈ Lµ (D). Let n be a positive even integer
such that (D − λ1L )n/2 x = 0 and (D − µ1L )n/2 y = 0. By Lemma 5.5.3 we have
n
n X n
(D − λ1L )k x, (D − µ1L )n−k y .
D − (λ + µ)1L ([x, y]) =
k
k=0
Weyl’s theorem
51
52 CHAPTER 6. WEYL’S THEOREM
βV (x, [y, z]) = 0. This proves that [x, y] ∈ I, so that I is an ideal of L. Since L
is semi-simple, to prove that I = 0 it will now suffice to prove that I is solvable.
Consider J = ϕ(I). Since ϕ is faithful, I ∼ = J; thus, it suffices to prove that J
is solvable. Now by the definition of I we have tr(xy) = 0 for all x ∈ J and
y ∈ ϕ(L); in particular, we have tr(xy) = 0 for all x, y ∈ J. By Lemma 5.2.3,
the Lie algebra J is solvable.
Let the notation be as in the statement of Lemma 6.1.1. Since the symmetric
bilinear form βV is non-degenerate, if x1 , . . . , xn is an ordered basis for L, then
there exists a unique ordered basis x01 , . . . , x0n for L such that
for i, j ∈ {1, . . . , n}. We refer to x01 , . . . , x0n as the basis dual to x1 , . . . , xn with
respect to βV .
Lemma 6.1.2. Assume that F has characteristic zero. Let L be a semi-simple
finite-dimensional Lie algebra over F , let V be a finite-dimensional F -vector
space, and let ϕ : L → gl(V ) be a faithful representation. Let x1 , . . . , xn be an
ordered basis for L, with dual basis x01 , . . . , x0n defined with respect to βV . Define
n
X
C= ϕ(xi )ϕ(x0i ).
i=1
Then C ∈ gl(V ), the definition of C does not depend on the choice of ordered
basis for L, and Cϕ(x) = ϕ(x)C for x ∈ L. Moroever, tr(C) = dim L. We refer
to C as the Casmir operator for ϕ.
Proof. To show that the definition of C does not depend on the choice of basis,
let y1 , . . . , yn be another ordered basis for L. Let (mij ) ∈ GL(n, F ) be the
matrix such that
n
X
yi = mij xj
j=1
n
X n
X
= nlj βV ( mik xk , x0l )
l=1 k=1
n X
X n
= nlj mik βV (xk , x0l )
l=1 k=1
n X
X n
= nlj mik δkl
l=1 k=1
Xn X n
= nlj mil
l=1 k=1
= δij .
It follows that
n
X
yj0 = nlj x0l
l=1
This proves that the definition of C does not depend on the choice of ordered
basis for L.
Next, let x ∈ L. We need to prove that Cϕ(x) = ϕ(x)C. Let (ajk ) ∈ M(n, F )
be such that
n
X
[xj , x] = ajk xk
k=1
= 0.
Finally, we have
Xn
tr(C) = tr( ϕ(xi )ϕ(x0i ))
i=1
n
X
= tr(ϕ(xi )ϕ(x0i ))
i=1
n
X
= βV (xi , x0i )
i=1
n
X
= 1
i=1
= dim L.
This completes the proof.
6.2. PROOF OF WEYL’S THEOREM 55
Proof. By Lemma 5.4.3, I ⊥ is also a Lie algebra over F , I and I ⊥ are semi-
simple as Lie algebras over F , and L = I⊕I ⊥ as Lie algebras. We have L/I ∼
= I⊥
as Lie algebras; it follows that L/I is semi-simple.
Lemma 6.2.3. Let L be a Lie algebra over F , and let V and W be L-modules.
Let
M = Hom(V, W )
be the F -vector space of all F -linear maps from V to W . For x ∈ L and T ∈ M
define x · T : V → W by
(x · T )(v) = x · T (v) − T (x · v)
M1 = {T ∈ Hom(V, W ) : f |W is a constant}
and
M0 = {T ∈ Hom(V, W ) : f |W = 0}
are L subspaces of M with M0 ⊂ M1 and the action of L maps M1 into
M0 .
and
x(yT ) − y(xT ) (v)
56 CHAPTER 6. WEYL’S THEOREM
= x(yT ) (v) − y(xT ) (v)
= x((yT )(v)) − (yT )(xv) − y((xT )(v)) + (xT )(yv)
= x(yT (v) − T (yv)) − y(T (xv)) + T (y(xv))
− y(xT (v) − T (xv)) + xT (yv) − T (x(yv))
= x(yT (v)) − xT (yv) − y(T (xv)) + T (y(xv))
− y(xT (v)) − yT (xv)) + xT (yv) − T (x(yv))
= x(yT (v)) + T (y(xv)) − y(xT (v)) − T (x(yv)).
It follows that
[x, y] · T = x(yT ) − y(xT )
so that with the above definition Hom(V, W ) is an L-module.
The assertion 1 of the lemma is clear.
To prove the assertion 2, let T ∈ M1 and let a ∈ F be such that T (w) = aw
for w ∈ W . Let x ∈ L. Let w ∈ W . Then
Theorem 6.2.4 (Weyl’s Theorem). Let F be algebraically closed and have char-
acteristic zero. Let L be a finite-dimensional semi-simple Lie algebra over F .
If (ϕ, V ) is a finite-dimensional representation of L, then V is a direct sum of
irreducible representations of L.
Proof. By induction, to prove the theorem it will suffice to prove that if W is a
proper, non-zero L-subspace of V , then W has a complement, i.e., there exists
an L-subspace W 0 of V such that V = W ⊕ W 0 . Let W be a proper, non-zero
L-subspace of V .
We first claim that W has a complement in the case that dim W = dim V −1.
Assume that dim W = dim V − 1.
We will first prove our claim when W is irreducible; assume that W is irre-
ducible. The kernel ker(ϕ) of ϕ : L → gl(V ) is an ideal of L. By Lemma 6.2.1
the Lie algebra L/ ker(ϕ) is semi-simple. By replacing ϕ : L → gl(V ) by the
representation ϕ : L/ ker(ϕ) → gl(V ), we may assume that ϕ is faithful. Con-
sider the quotient V /W . By assumption, this is a one-dimensional L-module.
Since [L, L] acts by zero on any one-dimensional L-module, and since L = [L, L]
by Lemma 6.2.2, it follows that L acts by zero on V /W . This implies that
ϕ(L)V ⊂ W . In particular, if C is the Casmir operator for ϕ, then CV ⊂ W .
By Lemma 6.1.2, C is an L-map. Hence, ker(C) is an L-submodule of V ; we will
prove that V = W ⊕ ker(C), so that ker(C) is a complement to W . To prove
that ker(C) is a complement to W it will suffice to prove that W ∩ ker(C) = 0
and dim ker(C) = 1. Consider the restriction C|W of C to W . This is an L-map
6.2. PROOF OF WEYL’S THEOREM 57
It follows that tr(C) = (dim W )a. On the other hand, by Lemma 6.1.2, we have
tr(C) = dim L. It follows that (dim W )a = dim L, and in particular, a 6= 0.
Thus, C is injective on W and maps onto W . Therefore, W ∩ ker(C) = 0, and
dim ker(C) = dim V − dim im(C) = dim V − dim W = 1. This proves our claim
in the case that W is irreducible.
We will now prove our claim by induction on dim V . We cannot have
dim V = 0 or 1 because W is non-zero and proper by assumption. Suppose
that dim V = 2. Then dim W = 1, so that W is irreducible, and the claim
follows from the previous paragraph. Assume now that dim V ≥ 3, and that
for all L-modules A with dim A < dim V , if B is an L-submodule of A of co-
dimension one, then B has a complement. If W is irreducible, then W has a
complement by the previous paragraph. Assume that W is not irreducible, and
let W1 be a L-submodule of W such that 0 < dim W1 < dim W . Consider
the L-submodule W/W1 of V /W1 . This L-submodule has co-dimension one in
V /W1 , and dim V /W1 < dim V . By the induction hypothesis, there exists an
L-submodule U of V /W1 such that
V /W1 = U ⊕ W/W1 .
Since dim M = 1 + dim W1 < 1 + dim W ≤ dim V , we can apply the induction
hypothesis again: let W2 be an L-submodule of M that is a complement to W1
in M , i.e.,
M = W1 ⊕ W2 .
We assert that W2 is a complement to W in V , i.e., V = W ⊕ W2 . Since
dim W2 = 1, to prove this it suffices to prove that W ∩ W2 = 0. Assume that
w ∈ W ∩ W2 . Then
w + W1 ∈ W/W1 ∩ M/W1 = 0.
Using the claim, we will now prove that W has a complement. Set
M = Hom(V, W ),
M1 = {T ∈ Hom(V, W ) : f |W is multiplication by some constant},
M0 = {T ∈ Hom(V, W ) : f |W = 0}.
M1 −→ F w
M1 = M0 ⊕ M00 .
V = W ⊕ ker(T ).
Proof. We will first prove that [xs , L] ⊂ L and [xn , L] ⊂ L. To see this, consider
adgl(V ) (x) : gl(V ) → gl(V ). This linear map has a Jordan-Chevalley decompo-
sition adgl(V ) (x) = adgl(V ) (x)s + adgl(V ) (x)n . Because x ∈ L, the linear map
adgl(V ) (x) maps L into L (i.e., [x, L] ⊂ L). Because adgl(V ) (x)s and adgl(V ) (x)n
are polynomials in adgl(V ) (x), these linear maps also map L into L. Now by
Lemma 5.1.3 we have adgl(V ) (x)s = adgl(V ) (xs ) and adgl(V ) (x)n = adgl(V ) (xn ).
It follows that adgl(V ) (xs ) and adgl(V ) (xn ) map L into L, i.e., [xs , L] ⊂ L and
[xn , L] ⊂ L.
6.3. AN APPLICATION TO THE JORDAN DECOMPOSITION 59
Define
N = {y ∈ gl(V ) : [y, L] ⊂ L}.
Evidently, L ⊂ N ; also, we just proved that xs , xn ∈ N . Moreover, we claim
that N is a Lie subalgebra of gl(V ), and that L is an ideal of N . To see that N
is a Lie subalgebra of gl(V ), let y1 , y2 ∈ N . Let z ∈ L. Then
ad(z + I)(y + I) = [z + I, y + I]
= [z, y] + I
= ad(z)(y) + I.
minimal polynomial of ad(s+I) has no repeated roots; this implies that ad(s+I)
is diagonalizable. Similarly, since ad(n) is nilpotent, we see that ad(n + I) is
nilpotent. Finally, we have [s + I, n + I] = [s, n] + I = 0 + I = I.
Theorem 6.3.4. Let F have characteristic zero and be algebraically closed.
Let L be a semi-simple finite-dimensional Lie algebra over F . Let V be a finite-
dimensional F -vector space, and let θ : L → gl(V ) be a homomorphism. Let
x ∈ L. Let x = s + n be the abstract Jordan decomposition of x as in Theorem
5.5.5. Then the Jordan-Chevalley decomposition of θ(x) ∈ gl(V ) is given by
θ(x) = θ(s) + θ(n), with θ(s) diagonalizable and θ(n) nilpotent.
Proof. Set J = θ(L); this is a Lie subalgebra of gl(V ). Since we have an
isomorphism of Lie algebras
∼
θ : L/ ker(θ) −→ J
for x, y ∈ L. This means that ad(x)m = 0. Thus, for every x ∈ L, the linear
map ad(x) is nilpotent. Conversely, assume that for every x ∈ L, the linear
map ad(x) ∈ gl(L) is nilpotent. Consider the Lie subalgebra ad(L) of gl(L). By
Theorem 3.1.1, the original version of Engel’s Theorem, there exists a basis for L
in which all the elements of ad(L) are strictly upper triangular; this implies that
ad(L) is a nilpotent Lie algebra. By Proposition 2.2.1, since ad(L) ∼ = L/Z(L) is
nilpotent, the Lie algebra L is also nilpotent.
Lemma 7.0.2. Let F have characteristic zero and be algebraically closed. Let
L be a semi-simple finite-dimensional Lie algebra over F . Then L has a Cartan
subalgebra.
Proof. It will suffice to prove that L contains a non-zero abelian subalgebra
consisting of semi-simple elements; to prove this, it will suffice to prove that L
63
64 CHAPTER 7. THE ROOT SPACE DECOMPOSITION
Moreover:
1. If α, β ∈ H ∨ , then
[Lα , Lβ ] ⊂ Lα+β .
2. If α, β ∈ H ∨ and α + β 6= 0, then
κ(Lα , Lβ ) = 0,
α(h)κ(x, y) = κ(α(h)x, y)
65
= κ([h, x], y)
= −κ([x, h], y)
= −κ(x, [h, y])
= −κ(x, β(h)y)
= −β(h)κ(x, y).
It follows that (α+β)(h)κ(x, y) = 0. Since this holds for all h ∈ H and α+β 6= 0,
it follows that κ(x, y) = 0. That is, κ(Lα , Lβ ) = 0.
To prove 3, let x ∈ L0 . Assume that κ(x, y) = 0 for all y ∈ L0 . By 2, we
have then κ(x, L) = 0. Since κ is non-degenerate, we must have x = 0.
Lemma 7.0.4. Let F have characteristic zero and be algebraically closed. Let
L be a semi-simple finite-dimensional Lie algebra over F . Let H be a Cartan
subalgebra of L. Let h ∈ H be such that dim CL (h) is minimal. Then CL (h) =
CL (H).
Proof. We first claim that for all s ∈ H, we have CL (h) ⊂ CL (s). Let s ∈ H.
There are filtrations of F -vector spaces:
x1 , . . . , x k
be any basis for CL (h) ∩ CL (s). Next, consider the restrictions of ad(h) and
ad(s) to CL (s). Since [s, CL (s)] = 0, the restriction of ad(s) to CL (s) is zero.
We claim that ad(h) maps CL (s) to itself. To see this, let x ∈ CL (s). We
calculate:
itself, since ad(s) and ad(h) commute, and since both ad(s) and ad(h) are diag-
onalizable, the restrictions of ad(s) and ad(h) to CL (s) can be simultaneously
diagonalized, so that there exist elements y1 , . . . , y` in CL (s) so that
x1 , . . . , xk , y1 , . . . , y`
is a basis for CL (s), and each element is an eigenvector for ad(s) and ad(h) (the
elements x1 , . . . , xk are already in the 0-eigenspaces for the restrictions of ad(h)
and ad(s) to CL (s)). Since ad(s) is zero on CL (s), the elements y1 , . . . , y` are in
the 0-eigenspace for ad(s). Similarly, there exist elements z1 , . . . , zm in CL (h)
such that
x1 , . . . , xk , z1 , . . . , zm
is a basis for CL (h) and each element is an eigenvector for ad(s) and ad(h);
note that since ad(h) is zero on CL (h), the elements z1 , . . . , zm are in the 0
eigenspace for ad(h). We claim that
x1 , . . . , xk , y1 , . . . , y` , z1 , . . . , zm
form a basis for CL (h) + CL (s). It is evident that these vectors span CL (h) +
CL (s). Now
It follows that this is a basis for CL (s) + CL (h). Finally, there exist elements
w1 , . . . , wn in L such that
x1 , . . . , xk , y1 , . . . , y` , z1 , . . . , zm , w1 , . . . , wn
is a basis for L and w1 , . . . , wn are eigenvectors for ad(s) and ad(h). Since
w1 , . . . , wn are not in CL (s), it follows that the eigenvalues of ad(s) on these
elements do not include zero; similarly, the eigenvalues of ad(h) on w1 , . . . , wn
do not include zero. Let α, . . . , αn in F and β1 , . . . , βn be such that
ad(s)wi = αi wi , ad(h)wi = βi wi
We have:
x1 , . . . , xk , y1 , . . . , y` , z1 , . . . , zm , w1 , . . . , wn
CL (s + c · h) ⊂ CL (s) ∩ CL (h).
CL (s + c · h) = CL (s) ∩ CL (h).
The operator ad(m) : L → L is nilpotent; it follows that adCL (h) (y) is also
nilpotent. Hence, CL (h) is a nilpotent Lie algebra.
Now we prove that the n from the first paragraph is zero. Since CL (h) is a
nilpotent Lie algebra, it is a solvable Lie algebra. Consider the Lie subalgebra
ad(CL (h)) of gl(L). Since L is semi-simple, ad is injective (see Proposition
5.5.1). It follows that ad(CL (h)) is a solvable Lie subalgebra of gl(L). By Lie’s
Theorem, Theorem 3.1.2, there exists a basis for L in which all the elements of
ad(CL (h)) are upper-triangular. The element ad(n) is a nilpotent element of
gl(L), and is hence strictly upper triangular. Let z ∈ CL (h). Then
κ(n, z) = tr(ad(n)ad(z)) = 0
Corollary 7.0.6. Let F have characteristic zero and be algebraically closed. Let
L be a semi-simple finite-dimensional Lie algebra over F . Let H be a Cartan
subalgebra of L. Then L0 = H.
Lemma 7.0.7. Let F have characteristic zero and be algebraically closed. Let
L be a semi-simple finite-dimensional Lie algebra over F . Let H be a Cartan
subalgebra of L, and let the notation be as in Proposition 7.0.3. If α ∈ Φ, then
−α ∈ Φ. Let α ∈ Φ, and let x ∈ Lα be non-zero. There exists y ∈ L−α such
that F x + F y + F [x, y] is a Lie subalgebra of L isomorphic to sl(2, F ).
Since κ(x, z) 6= 0, this implies that there exists β ∈ Φ such that α + β = 0, i.e,
−α ∈ Φ. Also, we have proven that there exists y ∈ L−α such that κ(x, y) 6= 0.
By 1 of Proposition 7.0.3 and Corollary 7.0.6 we have [x, y] ∈ L0 = H.
Let c ∈ F × . We claim that S(cy) = F x + F y + F [x, y] is a Lie subalgebra of
L. To prove this it suffices to check that [[x, y], x], [[x, y], y] ∈ S(cy). Now since
[x, y] ∈ H, we have by the definition of Lα ,
e = x, f = cy, h = [e, f ].
We first claim that h is non-zero for all c ∈ F × . We will prove the stronger
statement that α([x, y]) 6= 0. Assume that α([x, y]) = 0; we will obtain a
contradiction. From above, we have that [x, y] commutes with x and y. This
implies that ad([x, y]) = [ad(x), ad(y)] commutes with ad(x) and ad(y); these
are elements of gl(L). By Corollary 3.2.2, the element ad([x, y]) is a nilpotent
element of gl(L). However, by the definition of a Cartan subalgebra, ad([x, y])
is semi-simple. It follows that [x, y] = 0. Since α 6= 0, there exists t ∈ H such
that α(t) 6= 0. Now
[e, h] = −[h, x]
= −α(h)x
= −α([x, cy])x
= −cα([x, y])x
= −cα([x, y])e
and
[f, h] = −[h, f ]
= −[[x, cy], cy]
= −c[[x, y], cy]
= −c(−α([x, y]))f
= cα([x, y])f
sl(α) = F eα + F fα + F hα .
We note that
α(hα ) = α((2/α([x, y]))[x, y]) = 2.
Consider the action of sl(α) on L. By Weyl’s Theorem, Theorem 6.2.4, L can be
written as a direct sum of irreducible sl(α) representations. By Theorem 4.3.7
every one of these irreducible representations is of the form Vd for some integer
d ≥ 0. Moreover, the explicit description of the representations Vd shows that
Vd is a direct sum of hα eigenspaces, and each eigenvalue is an integer. It follows
that L is a direct sum of hα eigenspaces, and that each eigenvalue is an integer.
As every subspace Lβ for β ∈ Φ is obviously contained in the β(hα )-eigenspace
for hα , this implies that for all β ∈ Φ we have that β(hα ) is an integer.
Proposition 7.0.8. Let F have characteristic zero and be algebraically closed.
Let L be a semi-simple finite-dimensional Lie algebra over F . Let H be a Cartan
subalgebra of L, and let the notation be as in Proposition 7.0.3. Let β ∈ Φ. The
space Lβ is one-dimensional, and the only F -multiples of β contained in Φ are
β and −β.
71
X(β) = {c ∈ F : cβ ∈ Φ}.
[hβ , x] = (cβ)(hβ )x
= cβ(hβ )x
= 2cx.
and
X− (β) = {c ∈ F : cβ ∈ Φ and c < 0}.
We have
X(β) = X− (β) t X+ (β).
To prove the proposition it will suffice to prove that
α = c0 β.
here, we have used 1 of Proposition 7.0.3. This implies that [sl(α), Lcα ] ⊂ M .
Thus, sl(α) acts on M . The subspace M contains several subspaces. Evidently,
sl(α) ⊂ H ⊕ Lα ⊕ L−α ⊂ M.
K = ker(α) ⊂ H.
We claim that
K ∩ sl(α) = 0.
To see this, let k ∈ K ∩sl(α). Since K ⊂ H, we have k ∈ H ∩sl(α) = F hα ; write
k = ahα for some a ∈ F . By the definition of K, α(k) = 0. Since α(hα ) = 2,
we get a = 0 so that k = 0. Now let
N = K ⊕ sl(α).
H ⊂ N.
0 = [hα , v]
X
= [hα , h] + [hα , vcα ]
c∈X(α)
X
=0+ cα(hα )vcα
c∈X(α)
X
= 2cvcα .
c∈X(α)
and this sum is direct, we must have vcα = 0 for all c ∈ X(α). Hence, v = h ∈
H ⊂ N . On the other hand, v ∈ W . Therefore, v ∈ N ∩ W = 0, so that v = 0;
this is a contradiction. It follows that the Vd that occur in the decomposition
of W are such that d is odd.
Let d be an odd integer with d ≥ 1 and such that Vd occurs in W . By the
explicit description of Vd , there exists a vector v in Vd such that hα v = v, i.e,
[hα , v] = v. Again write M
v =h⊕ vcα
c∈X(α)
v = [hα , v]
X
= [hα , h] + [hα , vcα ]
c∈X(α)
X
=0+ cα(hα )vcα
c∈X(α)
X
= 2cvcα .
c∈{c∈X(α)
Therefore, M M
h⊕ vcα = 2cvcα
c∈X(α) c∈X(α)
Since v 6= 0, this implies that for some c ∈ X(α) we have 2c = 1, i.e, c = 1/2 ∈
X(α). This contradicts the fact that 1/2 ∈ / X(α). It follows that W = 0.
Since W = 0, we have N = M . This implies that #X+ (α) = 1 and dim Lα =
1. Hence, #X+ (β) = 1. Since 1 ∈ X+ (β), we obtain X+ (β) = {1}, so that
c0 = 1. This implies that in fact β = α, so that dim Lβ = 1. The proof is
complete.
74 CHAPTER 7. THE ROOT SPACE DECOMPOSITION
{k ∈ Z : β + kα ∈ Φ} = {k ∈ Z : −r ≤ k ≤ q}.
Moreover, r − q = β(hα ).
3. If α + β ∈ Φ, then [eα , eβ ] is a non-zero multiple of eα+β .
4. We have β − β(hα )α ∈ Φ.
Proof. Proof of 1. Consider the action of sl(α) on L. By Weyl’s Theorem,
Theorem 6.2.4, L is a direct sum of irreducible representations of sl(α). By
Theorem 4.3.7, each of these representations is of the form Vd for some integer
d ≥ 0. Each Vd is a direct sum of eigenspaces for hα , and each eigenvalue for hα
is an integer. It follows that L is a direct sum of eigenspaces for hα , with each
eigenvalue being an integer. Let x ∈ Lβ be non-zero. Then [hα , x] = β(hα )x,
so that β(hα ) is an eigenvalue for hα . It follows that β(hα ) is an integer.
Proof of 2. Let M
M= Lβ+kα .
k∈Z
We claim that there does not exist a k ∈ Z such that β + kα = 0. For suppose
such a k exists; we will obtain a contradiction. We have β = −kα. Hence,
−kα ∈ Φ. By Proposition 7.0.8 we must have −k = ±1. Thus, β = ±α; this
contradicts our hypothesis that β 6= ±α and proving our claim. It follows that
for every k ∈ Z either β + kα ∈ Φ or Lβ+kα = 0. Next, we assert that M is an
sl(α) module. Let k ∈ Z and x ∈ Lβ+kα . Then
Here we have used 1 of Proposition 7.0.3 and the fact that α(hα ) = 2. These
formulas show that M is an sl(α) module. We also see from the last formula
that M is the direct sum of hα eigenspaces because hα acts on the zero or
one-dimensional F -subspace Lβ+kα by β(hα ) + 2k for k ∈ Z; moreover, every
eigenvalue for hα is an integer, and all the eigenvalues for hα have the same
parity. As in the proof of 1, M is a direct sum of irreducible representations
of the form Vd for d a non-negative integer. The explicit description of the
representations of the form Vd for d a non-negative integer implies that if more
than one such representation Vd occurs in the decomposition of M , then either
some hα eigenspace is at least two-dimensional, or the hα eigenvalues do not
75
all have the same parity. It follows that M is irreducible, and there exists a
non-negative integer such that M ∼= Vd . The explicit description of Vd implies
that
Md
M= M (d − 2n)
n=0
where
M (n) = {x ∈ M : hα x = nx}
for n ∈ {0, . . . , d}, and that each of the hα eigenspaces M (d − 2n) for n ∈
{0, . . . , d} is one-dimensional. Now consider the set
{k ∈ Z : β + kα ∈ Φ}.
k ∈ {k ∈ Z : β + kα ∈ Φ}
q = (d − β(hα ))/2
is an integer; since k may assume the value 0, we also see that q is non-negative.
Continuing, we have
d ≥ n ≥ 0,
−d ≤ −n ≤ 0,
q−d≤q−n≤q
−(d − q) ≤ k ≤ q,
−r ≤ k ≤ q,
{k ∈ Z : β + kα ∈ Φ} ⊂ {k ∈ Z : −r ≤ k ≤ q}.
Now
#{k ∈ Z : β + kα ∈ Φ} = dim M = dim Vd = d + 1.
Also,
#{k ∈ Z : −r ≤ k ≤ q} = q − (−r) + 1
=q+r+1
=q+d−q+1
= d + 1.
76 CHAPTER 7. THE ROOT SPACE DECOMPOSITION
It follows that
{k ∈ Z : β + kα ∈ Φ} = {k ∈ Z : −r ≤ k ≤ q},
as desired. Finally,
r − q = d − q − q = d − 2q = d − (d − β(hα )) = β(hα ).
1∈
/ {k ∈ Z : β + kα ∈ Φ}.
−r ≤ q − r ≤ q,
−r ≤ −(r − q) ≤ q,
−r ≤ −β(hα ) ≤ q.
H −→ H ∨ (7.1)
defined by h 7→ κ(·, h). By 3 of Proposition 7.0.3 and Corollary 7.0.6, this map
is injective, i.e., the restriction of the Killing form to H is non-degenerate; since
both F -vector spaces have the same dimension, it is an isomorphism. There is
thus a natural isomorphism between H and H ∨ . In particular, for every root
α ∈ Φ there exists tα ∈ H such that
α(x) = κ(x, tα )
for x ∈ H.
Lemma 7.0.11. Let F have characteristic zero and be algebraically closed. Let
L be a semi-simple finite-dimensional Lie algebra over F . Let H be a Cartan
subalgebra of L, and let the notation be as in Proposition 7.0.3. Let α ∈ Φ.
1. For x ∈ Lα and y ∈ L−α we have
In particular,
hα = [eα , fα ] = κ(eα , fα )tα .
2. We have
2
hα = tα .
κ(tα , tα )
and
κ(tα , tα )κ(hα , hα ) = 4.
3. If β ∈ Φ, then
2(α, β)
= β(hα ).
(α, α)
κ(h, [x, y] − κ(x, y)tα ) = κ(h, [x, y]) − κ(h, κ(x, y)tα )
= κ([h, x], y) − κ(x, y)κ(h, tα )
= κ(α(h)x, y) − κ(x, y)α(h)
= α(h)κ(x, y) − κ(x, y)α(h)
= 0.
78 CHAPTER 7. THE ROOT SPACE DECOMPOSITION
Since this holds for all h ∈ H, and since the restriction of the Killing form to
H is non-degenerate, we obtain [x, y] − κ(x, y)tα = 0. This proves the first and
second assertions.
2. We first note that
2 = α(hα )
= κ(hα , tα )
= κ(κ(eα , fα )tα , tα )
2 = κ(eα , fα )κ(tα , tα )
2
= κ(eα , fα ).
κ(tα , tα )
The first claim of 2 now follows now from 1 by substitution. Next, we have:
2 2
κ(hα , hα ) = κ( tα , tα )
κ(tα , tα ) κ(tα , tα )
22
= κ(tα , tα )
κ(tα , tα )2
4
= .
κ(tα , tα )
3. Using the definition of (·, ·) and tα and tβ , we have
2(α, β) 2κ(tα , tβ )
=
(α, α) κ(tα , tα )
2
= · κ(tα , tβ )
κ(tα , tα )
= κ(eα , fα ) · κ(tα , tβ )
= κ(κ(eα , fα ) · tα , tβ )
= κ(hα , tβ )
= β(hα ).
β = c1 α1 + · · · + c` α`
for c1 , . . . , c` ∈ F . Then c1 , . . . , c` ∈ Q.
Proof. Let i ∈ {1, . . . , `}. Then
It follows that
(α1 , β) c1
.. ..
. =S .
(α` , β) c`
where
(α1 , α1 ) · · · (α1 , α` )
S= .. ..
.
. .
(α` , α1 ) ··· (α` , α` )
Since (·, ·) is a non-degenerate symmetric bilinear form the matrix S is invertible.
Therefore,
(α1 , β) c1
S −1 ... = ... .
(α` , β) c`
By the remark preceding the proposition the entries of all the matrices on the
left are in Q; hence, c1 , . . . , c` ∈ Q.
Proposition 7.0.14. Let F have characteristic zero and be algebraically closed.
Let L be a semi-simple finite-dimensional Lie algebra over F . Let H be a Cartan
subalgebra of L, and let the notation be as in Proposition 7.0.3. As a Lie algebra,
L is generated by the root spaces Lα for α ∈ Φ.
Proof. By the decomposition
M
L=H⊕ Lα
α∈Φ
that follows from Proposition 7.0.3 and Corollary 7.0.6 it suffices to prove that
H is contained in the F -span of the F -subspaces [Lα , L−α ] for α ∈ Φ. By the
discussion preceding Proposition 7.0.8, the elements hα for α ∈ Φ are contained
in this F -span. By Lemma 7.0.11, this F -span therefore contains the elements tα
for α ∈ Φ. By Lemma 7.0.10, the linear forms α ∈ Φ span H ∨ ; this implies that
the elements tα for α ∈ Φ span H. The F -span of the F -subspaces [Lα , L−α ]
for α ∈ Φ therefore contains H.
7.1. AN ASSOCIATED INNER PRODUCT SPACE 81
(a ⊗ v, b ⊗ w) = ab(v, w)0
x = a1 (1 ⊗ v1 ) + · · · + an (1 ⊗ vn ) = a1 ⊗ v1 + · · · + an ⊗ vn .
We have
n
X
(x, x) = (ai ⊗ vi , aj ⊗ vj )
i,j=1
Xn
= ai aj (vi , vj )0
i,j=1
X n
= a2i (vi , vi )0 .
i=1
Since (·, ·)0 is positive-definite, (vi , vi )0 > 0 for i ∈ {1, . . . , n}. It follows that if
(x, x) = 0, then a1 = · · · = an = 0, so that x = 0.
(y, y) = κ(h, h)
= tr(ad(h) ◦ ad(h))
X
= α(h)2
α∈Φ
X
= κ(tα , h)2
α∈Φ
X
= (α, y)2 .
α∈Φ
Root systems
det(sv ) = −1.
We will write
2(x, y)
hx, yi =
(y, y)
for x, y ∈ V . We note that the function h·, ·i : V × V → R is linear in the first
variable; however, this function is not linear in the second variable. We have
for x ∈ V .
Let R be a subset of V . We say that R is a root system if R satisfies the
following axioms:
83
84 CHAPTER 8. ROOT SYSTEMS
(R2) If α ∈ R, then α and −α are the only scalar multiples of α that are
contained in R.
In particular,
L0 = {x ∈ L : [h, x] = 0 for all h ∈ H}.
Here, Φ is the subset of α in
H ∨ = HomF (H, F )
such that Lα 6= 0. The elements of Φ are called the roots of L with respect to
H. By Corollary 7.0.6 we have L0 = H so that in fact
M
L=H⊕ Lα .
α∈Φ
V0 = Q span of Φ in H ∨ .
and the restriction (·, ·)0 of the symmetric bilinear form on H ∨ to V0 is an inner
product, i.e., is positive definite, and is Q valued. Let
V = R ⊗Q V 0 ,
so that V is an R vector space, and define an R symmetric bilinear form (·, ·) on
V by declaring (a ⊗ v, b ⊗ w) = ab(v, w)0 for a, b ∈ R and v, w ∈ V0 . By Lemma
7.1.1, we have that (·, ·) is positive-definite.
Proposition 8.2.1. Let the notation be as in the discussion preceding the propo-
sition. The subset Φ of the inner product space V is a root system.
Proof. It is clear that (R1) is satisfied. (R2) is satisfied by Proposition 7.0.8.
To see that (R3) is satisfied, let α, β ∈ Φ. Then by 3 of Lemma 7.0.11,
2(β, α)
sα (β) = β − α = β − β(hα )α.
(α, α)
By 4 of Proposition 7.0.9 we have β − β(hα )α ∈ Φ. It follows that sα (β) ∈ Φ,
so that (R3) is satisfied. To prove that (R4) holds, again let α, β ∈ Φ. We have
2(α, β)
hα, βi = .
(β, β)
By 3 of Lemma 7.0.11 we have
2(α, β)
= α(hβ ).
(β, β)
Finally, by 1 of Proposition 7.0.9, this quantity is an integer. This proves
(R4).
θ
y
z x
θ
y
ty
Then we have
x = z + ty.
Taking the inner product with y, we get
ktyk
cos θ =
kxk
kyk
cos θ = t
kxk
kxk
t = cos θ.
kyk
If we equate the two formulas for t we get (x, y) = kxkkyk cos θ. We say that
two vectors are orthogonal if (x, y) = 0; this is equivalent to the angle between
x and y being π/2. If (x, y) > 0, then we will say that x and y form an acute
angle; this is equivalent to 0 < θ < π/2. If (x, y) < 0, then we will say that x
and y form an obtuse angle; this is equivalent to π/2 < θ ≤ π.
Non-zero vectors also define some useful geometric objects. Let v ∈ V be
non-zero. We may consider three sets that partition V :
The first set consists of the vectors that form an acute angle with v, the middle
set is the hyperplane P orthogonal to Rv, and the last set consists of the vectors
that form an obtuse angle with v. We refer to the first and last sets as the half-
spaces defined by P . Of course, v lies in the first half-space. The formula for
the reflection sv shows that
for x in V , so that S sends one half-space into the other half-space. Also, S
acts by the identity on P . Multiplication by −1 also sends one half-space into
the other half-space; however, while multiplication by −1 preserves P , it is not
the identity on P .
Lemma 8.3.1. Let V be a vector space over R with an inner product (·, ·).
Let x, y ∈ V and assume that x and y are both non-zero. The following are
equivalent:
Assume that (x, y)2 = (x, x)(y, y) = kxk2 kyk2 . Then (x, y)2 = kxk2 kyk2 6= 0,
and cos2 θ = 1, so that cos θ = ±1. This implies that θ = 0 or θ = π/2.
3 =⇒ 2. Assume that the angle θ between x and y is 0 or π. Then
cos2 θ = 1. Hence, (x, y)2 = kxk2 kyk2 .
2 =⇒ 1. Suppose that (x, y)2 = (x, x)(y, y). We have
(x,y)
It follows that y − (x,x) x = 0, so that x and y are linearly dependent.
88 CHAPTER 8. ROOT SYSTEMS
kβk
angle type θ cos θ hα, βi hβ, αi kαk
√ √
π/6 = 30◦ 3/2 1 3 3
√ √
strictly acute π/4 = 45◦ 2/2 1 2 2
Proof. By the assumption kβk ≥ kαk we have (β, β) = kβk2 ≥ (α, α) = kαk2 ,
so that
2|(β, α)| 2|(α, β)|
|hβ, αi| = ≥ = |hα, βi|.
(α, α) (β, β)
8.3. BASIC THEORY OF ROOT SYSTEMS 89
By (R4) we have that hα, βi and hβ, αi are integers, and by Lemma 8.3.2 we
have hα, βihβ, αi ∈ {0, 1, 2, 3}. These facts imply that the possibilities for hα, βi
and hβ, αi are as in the table.
Assume first that hβ, αi = hα, βi = 0. From above, hα, βihβ, αi = 4 cos2 θ.
It follows that cos θ = 0, so that θ = π/2 = 90◦ .
Assume next that hβ, αi = 6 0. Now
hβ, αi 2(β, α) (β, β) (β, β)
= = ,
hα, βi (α, α) 2(α, β) (α, α)
hβ,αi
so that hα,βi is positive and
s
hβ, αi kβk
= ,
hα, βi kαk
so that
1 kβk
cos θ = hα, βi.
2 kαk
This gives the cos θ column.
Lemma 8.3.4. Let V be a finite-dimensional vector space over R equipt with
an inner product (·, ·), and let R be a root system in V . Let α, β ∈ R. Assume
that α 6= ±β and kβk ≥ kαk.
1. Assume that the angle θ between α and β is strictly obtuse, so that by
Lemma 8.3.3 we have θ = 2π/3 = 120◦ , θ = 3π/4 = 135◦ , or θ = 5π/6 =
150◦ . Then α + β ∈ R. Moreover,
θ = 3π/4 = 135◦ =⇒ 2α + β ∈ R,
θ = 5π/6 = 150◦ =⇒ 3α + β ∈ R.
2. Assume that the angle between α and β is strictly acute, so that by Lemma
8.3.3 we have θ = π/6 = 30◦ , θ = π/4 = 45◦ , or θ = π/3 = 60◦ . Then
−α + β ∈ R. Moreover,
Proof. 1. By (R3), we have sβ (α) = α − hα, βiβ ∈ R. Since the angle between α
and β is strictly obtuse, by Lemma 8.3.3 we have that hα, βi = −1. Therefore,
α + β ∈ R. Assume that θ = 3π/4 = 135◦ . By Lemma 8.3.3 we hβ, αi = −2.
Hence, sα (β) = β − hβ, αiα = β + 2α ∈ R. The case when θ = 5π/6 = 150◦ is
similar.
2. By (R3), we have sβ (α) = α − hα, βiβ ∈ R. Since the angle between α
and β is strictly acute, by Lemma 8.3.3 we have that hα, βi = 1. Therefore,
α − β ∈ R. Hence, −α + β ∈ R. Assume that θ = π/4 = 45◦ . By Lemma
8.3.3 we hβ, αi = 2. Hence, sα (β) = β − hβ, αiα = β − 2α ∈ R. The case
θ = π/3 = 60◦ is similar.
Proposition 8.3.5. Let V = R2 equipt with the usual inner product (·, ·), and
let R be a root system in V . Let ` be the length of the shortest root in R. Let S
be the set of pairs (α, β) of non-colinear roots such that kαk = ` and the angle
θ between α and β is obtuse, and β is to the left of α. The set S is non-empty.
Fix a pair (α, β) in S such that θ is maximal. Then
1. (A2 root system) If θ = 120◦ (so that kαk = kβk by Proposition 8.3.3),
then R, α, and β are as follows:
β α+β
60◦
60◦ 60◦
−α α
◦ ◦
60 60
60◦
−α − β −β
√
2. (B2 root system) If θ = 135◦ (so that kβk = 2kαk by Proposition 8.3.3),
then R, α, and β are as follows:
β α+β 2α + β
45◦ 45◦
◦
45 45◦
−α α
45◦ 45◦
45◦ 45◦
−2α − β −α − β −β
8.3. BASIC THEORY OF ROOT SYSTEMS 91
√
3. (G2 root system) If θ = 150◦ (so that kβk = 3kαk by Proposition 8.3.3),
then R, α, and β are as follows:
3α + 2β
β α+β 2α + β 3α + β
30◦ 30◦
30 ◦ 30◦
30◦ 30◦
−α α
30◦ 30◦
30◦ 30◦
30◦ 30◦
−3α − β −2α − β −α − β −β
−3α − 2β
4. (A1 × A1 root system) If θ = 90◦ (so that the relationship between kβk
and kαk is not determined by Proposition 8.3.3), then R, α, and β are as
follows:
90◦ 90◦
−α α
90◦ 90◦
−β
Proof. Let (α, β) be a pair of non-colinear roots in R such that kαk = `; such
a pair must exist because R contains a basis which includes α. If the angle
between α and β is acute, then the angle between α and −β is obtuse. Thus,
there exists a pair of roots (α, β) in R such that kαk = ` and the angle between
α and β is obtuse. If β is the right of α, then −β forms an acute angle with
92 CHAPTER 8. ROOT SYSTEMS
α and is to the left of α; in this case, sα (β) forms an obtuse angle with α and
sα (β) is to the left of β. It follows that S is non-empty.
Assume that θ = 120◦ , so that kαk = kβk by Lemma 8.3.3. By Lemma
8.3.4, α + β ∈ R. It follows that α, β, α + β, −α, −β, −α − β ∈ R. By geometry,
kα + βk = kαk = kβk. It follows that R contains the vectors in 1. Assume
that R contains a root γ other than α, β, α + β, −α, −β, −α − β. By Lemma
8.3.3 we see that γ must lie halfway between two adjacent roots from α, β, α +
β, −α, −β, −α − β. This implies that θ is not √ maximal, a contradiction.
Assume that θ = 135◦ , so that kβk = 2kαk by Lemma 8.3.3. By Lemma
8.3.4, we have α + β, 2α + β ∈ R. It follows that R contains α, β, α + β, 2α +
β, −α, −β, −α − β, −2α − β, so that R contains the vectors in 2. Assume that
R contains a root γ other than α, β, α + β, 2α + β, −α, −β, −α − β, −2α − β.
Then γ must make an angle strictly less than 30◦ with one of α, β, α + β, 2α +
β, −α, −β, −α − β, −2α − β. This is impossible√ by Lemma 8.3.3.
Assume that θ = 150◦ , so that kβk = 3kαk by Lemma 8.3.3. By Lemma
8.3.4 we have α + β, 3α + β ∈ R. By geometry, the angle between α and 3α + β
is 30◦ . By Lemma 8.3.3, −α + (3α + β) = 2α + β ∈ R. By geometry, the angle
between β and 3α + β is 120◦ . By Lemma 8.3.3, β + 3α + β = 3α + 2β ∈ R. It
now follows that R contains the vectors in 3. Assume that R contains a vector
γ other than α, β, α + β, 2α + β, 3α + β, 3α + 2β, −α, −β, −α − β, −2α − β, −3α −
β, −3α −2β. Then Then γ must make an angle strictly less than 30◦ with one of
α, β, α + β, 2α + β, 3α + β, 3α + 2β, −α, −β, −α − β, −2α − β, −3α − β, −3α − 2β.
This is impossible by Lemma 8.3.3.
Finally, assume that θ = 90◦ . Assume that R contains a root γ other than
α, β, −α, −β. Arguing as in the first paragraph, one can show that the set S
contains a pair with θ larger than 90◦ ; this is a contradiction. Thus, R is as in
4.
8.4 Bases
Let V be a finite-dimensional vector space over R equipt with an inner product
(·, ·), and let R be a root system in V . Let B be a subset of R. We say that B
is a base for R if
(B1) B is a basis for the R vector space V .
(B2) Every element α ∈ R can be written in the form
X
α= c(β)β
β∈B
where the coefficients c(β) for β ∈ B are all integers of the same sign (i.e.,
either all greater than or equal to zero, or all less than or equal to zero).
Assume that B is a base for R. We define
α is a linear combination of β ∈ B
R+ = α ∈ R : ,
with non-negative coefficients
8.4. BASES 93
α is a linear combination of β ∈ B
R− = α∈R: .
with non-positive coefficients
We have
R = R+ t R− .
We refer to R+ as the set of positive roots with respect to B and R− as the
set of negative roots with respect to B. If α ∈ R is written as in (B2), then
we define the height of α to be the integer
X
ht(α) = c(β).
β∈B
Pα = {x ∈ V : (x, α) = 0}
v ∈ V − ∪α∈R Pα .
We denote by Vreg the set of all vectors in V that are regular with respect to R,
so that
Vreg (R) = V − ∪α∈R Pα .
Evidently, Vreg (R) is an open subset of V ; however, it is not entirely obvious
that Vreg (R) is non-empty.
Proof. Assume that Ui is a proper subset of U for all i ∈ {1, . . . , n}. Since Ui is
a proper subset of U for all i ∈ {1, . . . , n} we must have n ≥ 2. After replacing
the collection of Ui for i ∈ {1, . . . , n} with a subcollection, we may assume that
Ui * Uj and Uj * Ui for i, j ∈ {1, . . . , n}, i 6= j. We will prove that U is not a
subspace for collections of proper subspaces U1 , . . . , Un with n ≥ 2 and such that
that Ui * Uj and Uj * Ui for i, j ∈ {1, . . . , n} by induction on n. Assume that
n = 2 and that U = U1 ∪ U2 is a subspace; we will obtain a contradiction. Since
U1 * U2 and U2 * U1 , there exist u2 ∈ U2 such that u2 ∈ / U1 and u1 ∈ U1 such
that u1 ∈/ U2 . Since U is a subspace we have u1 + u2 ∈ U . Hence, u1 + u2 ∈ U1
or u1 + u2 ∈ U2 . If u1 + u2 ∈ U1 , then u2 ∈ U1 , a contradiction; similary, if
u1 + u2 ∈ U2 , then u1 ∈ U2 , a contradiction. Thus, the claim holds if n = 2.
Suppose that n ≥ 3 and that the claim holds for n − 1; we will prove that
the claim holds for n. We argue by contradiction; assume that U is a subspace.
94 CHAPTER 8. ROOT SYSTEMS
u1 + λ1 u2 , u1 + λ2 u2 , ..., u1 + λn u2
are all contained in U , and hence must each lie in some Ui with i ∈ {1, . . . , n}.
However, no such vector can be in U1 because otherwise u2 ∈ U1 ; similarly, no
such vector can be in U2 . By the pigeonhole principle, this means that there exist
distinct j, k ∈ {2, . . . , n} and i ∈ {3, . . . , n} such that u1 + λj u2 , u1 + λk u2 ∈ Ui .
It follows that (λj − λk )u2 ∈ Ui , so that u2 ∈ Ui . This is a contradiction.
Proof. Assume that there exists no v ∈ V such that v is regular with respect
to R; we will obtain a contradiction. Since no regular v ∈ V exists, we have
V = ∪α∈R Pα . Since dim V ≥ 2, and since R contains a basis for V over R, it
follows that #R ≥ 2. Also, dim Pα = dim V − 1 for all α ∈ R. We now have a
contradiction by Lemma 8.4.1.
We will write
Evidently,
R = R+ (v) t R− (v).
Let α ∈ R+ (v). We will say that α is decomposable if α = β1 + β2 for some
β1 , β2 ∈ R+ (v). If α is not decomposable we will say that α is indecomposable.
We define
B(v) = {α ∈ R+ (v) : α is indecomposable}.
By the definition of R+ (v), the real numbers (v, α), (v, α1 ), and (v, α2 ) are all
positive. It follows that we must have (v, α) > (v, α1 ). This contradicts the
definition of α.
Proof. Assume that the angle between α and β is strictly acute. With out loss
of generality, we may assume that kαk ≤ kβk. Since (v, α) > 0 and (v, β) > 0 we
must have α 6= −β. By Lemma 8.3.4 we have γ = −α + β ∈ R. Since γ ∈ R, we
also have −γ ∈ R. Since R = R+ (v)tR− (v), we have γ ∈ R+ (v) or −γ ∈ R+ (v).
Assume that γ ∈ R+ (v). We have γ+α = β with γ, α ∈ R+ (v). This contradicts
the fact that β is indecomposable. Similarly, the assumption that −γ ∈ R+ (v)
implies that α = γ + β, contradicting the fact that α is indecomposable. It
follows that the angle between α and β is obtuse, i.e., (α, β) ≤ 0.
Proof. Assume that c(α) for α ∈ B are real numbers such that
X
0= c(α)α.
α∈B
We need to prove that c(α) = 0 for all α ∈ B. Suppose that c(α) 6= 0 for some
α ∈ B; we will obtain a contradiction. Since c(α) 6= 0 for some α ∈ B, we may
assume that, after possibly multiplying by −1, that there exists α ∈ B such
that c(α) > 0. Define X
x= c(α)α.
α∈B, c(α)>0
We also have X
x= (−c(β))β.
β∈B, c(β)<0
Therefore,
X X
(x, x) = ( c(α)α, (−c(β))β)
α∈B, c(α)>0 β∈B, c(β)<0
96 CHAPTER 8. ROOT SYSTEMS
X
(x, x) = c(α) · (−c(β))(α, β).
α∈B, c(α)>0
β∈B, c(β)<0
By the definition of B we have (v, α) > 0 for all α ∈ B. The last displayed
equation now yields a contradiction since the set of α ∈ B such that c(α) > 0
is non-empty.
Proof. We will begin by proving that (B2) holds. Evidently, since R− (v) =
−R+ (v), to prove that (B2) holds it suffices to prove that every β ∈ R+ (v) can
be written as X
β= c(α)α, c(α) ∈ Z≥0 . (8.1)
α∈B(v)
Let S be the set of β ∈ R+ (v) for which (8.1) does not hold. We need to prove
that S is empty. Suppose that S is not empty; we will obtain a contradic-
tion. Let β ∈ S be such that (v, β) is minimal. Clearly, β ∈ / B(v), i.e., β is
decomposable. Let β1 , β2 ∈ R+ (v) be such that β = β1 + β2 . We have
By the definition of R+ (v), the real numbers (v, β), (v, β1 ), and (v, β2 ) are all
positive. It follows that we must have (v, β) > (v, β1 ) and (v, β) > (v, β2 ).
The definition of β implies that β1 ∈ / S and β2 ∈ / S. Hence, β1 and β2 have
expressions as in (8.1). It follows that β = β1 + β2 has an expression as in (8.1).
This contradiction implies that (B2) holds.
Now we prove that B(v) satisfies (B1). Since R spans V , and since every
element of R is a linear combination of elements of B(v) because B(v) satisfies
(B2), it follows that B(v) spans V . Finally, B(v) is linearly independent by
Lemma 8.4.4 and Lemma 8.4.5.
Proof. By Lemma 8.4.7 there exists a vector v ∈ V such that (v, α) > 0 for
α ∈ B. We claim that v is regular with respect to R. Let β ∈ R, and write
X
β= c(α)α,
α∈B
where the coefficients c(α) for α ∈ B are integers of the same sign. We have
X
(v, β) = (v, c(α)α)
α∈B
X
= c(α)(v, α).
α∈B
Since all the coefficients c(α), α ∈ B, have the same sign, and since (v, α) > 0
for α ∈ B, it follows that (v, β) > 0 or (v, β) < 0. Thus, v is regular with respect
to R. Next, since (v, α) > 0 for α ∈ B, we have R+ ⊂ R+ (v) and R− ⊂ R− (v).
Since R = R+ t R− and R = R+ (v) t R− (v) we now have R+ = R+ (v) and
R− = R− (v). We now have B = B(v) by Lemma 8.4.8.
We have
X
(f (t), β) = (1 − t)v + tw, c(α)α
α∈B(v)
X X
= (1 − t) c(α) v, α + t c(α) w, α .
α∈B(v) α∈B(v)
Since (v, α), (w, α) > 0 for α ∈ B(v) it follows that (f (t), β) > 0; thus, the image
of f is indeed in Vreg (R), so that f is well-defined. Evidently, f is continuous,
and f (0) = v and f (1) = w. It follows that every element of X is path connected
in Vreg (R) to v.
Finally, we prove that if u ∈ Vreg (R) and u ∈ / X(v), then u is not path
connected in Vreg (R) to v. Suppose that u ∈ Vreg (R), u ∈ / X(v), and that u is
path connected in Vreg (R) to v; we will obtain a contradiction. Since u is path
connected in Vreg (R) to v there exists a continuous function g : [0, 1] → Vreg (R)
such that g(0) = v and g(1) = u. Since u ∈ / X(v), there exists α ∈ B(v) such
that (u, α) < 0. Define F : [0, 1] → R by F (t) = (g(t), α) for t ∈ [0, 1]. We have
F (0) > 0 and F (1) < 0. Since F is continuous, there exists a t ∈ (0, 1) such
that F (t) = 0. This means that (g(t), α) = 0. However, this is a contradiction
since g(t) is regular with respect to R.
Weyl chambers in V ∼
−→ Bases for R
with respect to R
Then
X
(v2 , β) = c(α)(v2 , α).
α∈B(v1 )
Since (v2 , α) > 0 for all α ∈ B(v1 ) we must have (v2 , β) > 0. Thus, R+ (v1 ) ⊂
R+ (v2 ). Similarly, R+ (v2 ) ⊂ R+ (v1 ), so that R+ (v1 ) = R+ (v2 ). We now obtain
B(v1 ) = B(v2 ) by Lemma 8.4.8.
To see that the map is injective, suppose that C1 and C2 are Weyl chambers
that map to the same base for R. Let v1 ∈ C1 and v2 ∈ C2 . By assumption, we
have B(v1 ) = B(v2 ). Since B(v1 ) = B(v2 ) we have X(v1 ) = X(v2 ). By Lemma
8.5.1, this implies that C1 = C2 .
Finally, the map is surjective by Proposition 8.4.9.
by M
f (w) = (w, α) ⊕ (|(w, β)| − |(w, α)|).
β∈R,β6=±α
102 CHAPTER 8. ROOT SYSTEMS
with |(v, β)| > 0 for β ∈ R, β 6= ±α. Fix > 0 be such that |(v, β)| > > 0 for
β ∈ R, β 6= ±α. Since f is continuous, there exists an open set A containing v
such that M
f (A) ⊂ (−, ) ⊕ (|(v, β)| − , |(v, β)| + ).
β∈R,β6=±α
0 < |(v, β)| − < |(w, β)| − |(w, α)| = |(w, β)| − (w, α)
so that
(w, α) < |(w, β)|.
Consider now the base B(w). We claim that α ∈ B(w). We have (w, α) > 0, so
that α ∈ R+ (w). Assume that α = β1 + β2 for some β1 , β2 ∈ R+ (w); we obtain
a contradiction, proving that α ∈ B(w). We must have β1 6= ±α1 and β2 6= ±α;
otherwise, 0 ∈ R or 2α ∈ R, a contradiction. Now
Since (w, β1 ) > 0 and (w, β2 ) > 0 we must have (w, α) > (w, β1 ). This contra-
dicts (w, α) < |(w, β1 )| = (w, β1 ).
Lemma 8.6.3. Let V be a finite-dimensional vector space over R equipt with
an inner product (·, ·), and let R be a root system in V . Let B be a base for R.
Let α be a positive root with respect to B such that α ∈ / B. Then there exists
β ∈ B such that (α, β) > 0 and α − β is a positive root.
Proof. By Proposition 8.4.9 there exists v ∈ Vreg (R) such that B = B(v). Since
α and the elements of B are all in R+ = R+ (v) (see Proposition 8.4.6) we have
(v, α) > 0 and (v, β) > 0 for β ∈ B. If (α, β) ≤ 0 for all β ∈ B, then by Lemma
8.4.4 Lemma 8.4.5, the set B t {α} is linearly independent, contradicting the
fact that B is a basis for the R vector space V . It follows that there exists β ∈ B
such that (α, β) > 0. By Lemma 8.3.4 we have α − β ∈ R. Since α is positive
we can write X
α = c(β)β + c(γ)γ
γ∈B,γ6=β
α1 ,
α1 + α2 ,
α2 + α2 + α3 ,
···
α1 + α2 + α3 + · · · + αt
α1 ,
α1 + α2 ,
α2 + α2 + α3 ,
···
α1 + α2 + α3 + · · · + αt
with c(γ) ∈ Z≥0 for γ ∈ B. We claim that c(γ0 ) > 0 for some γ0 ∈ B with
γ0 6= α. Suppose this is false, so that β = c(α)α; we will obtain a contradiction.
By (R2), we have c(α) = ±1. By hypothesis, α 6= β; hence, c(α) = −1, so that
β = −α. This contradicts the fact that β is positive, proving our claim. Now
This is the expression of the root sα (β) in terms of the base B. Since c(γ0 ) > 0,
we see that sα (β) is a positive root and that sα (β) 6= α, i.e, sα (β) ∈ R+ −
{α}.
Lemma 8.6.6. Let V be a finite-dimensional vector space over R equipt with
an inner product (·, ·), and let R be a root system in V . Let B be a base for R.
Set
1 X
δ= β.
2 +
β∈R
If α ∈ B, then
sα (δ) = δ − α.
Proof. We have
1 1
sα (δ) = sα ( α) + sα (δ − α)
2 2
1 1 X
=− α+ sα (β)
2 2 +
β∈R −{α}
1 1 X
=− α+ β
2 2
β∈R+ −{α}
1 1 1 X
=− α− α+ β
2 2 2 + β∈R
= −α + δ.
V = Rα ⊕ Pα .
It follows that the image of P under the projection map V → V /Rα is all of
V /Rα; similarly, the image of Pα under V → V /Rα is all of V /Rα. Since s fixes
P pointwise, it follows that the endomorphism of V /Rα induced by s is the
identity. Similarly, the endomorphism of V /Rα induced by sα is the identity.
Therefore, the endomorphism of V /Rα induced by t = ssα is also the identity.
Let v ∈ V . We then have t(v) = v + aα for some a ∈ R. Applying t again, we
obtain t2 (v) = t(v)+aα. Solving this last equation for aα gives aα = t2 (v)−t(v).
Substituting into the first equation yields:
That is, p(t) = 0 for p(z) = z 2 − 2z + 1 = (z − 1)2 . It follows that the minimal
polynomial of t divides (z − 1)2 . On the other hand, s and sα both send X into
X, so that t also sends X into X. Let β ∈ X, and consider the sequence
= s(−α)
= −s(α).
Equating, we conclude that hβ, s(α))i = hs−1 (β), αi. Since this holds for all
α, β ∈ R, this implies that hs(α), s(β)i = hα, βi for all α, β ∈ R (substitute s(α)
for β and β for α).
8.7. THE WEYL GROUP 107
We have
s1 (β1 ) = β0 ,
s2 (β2 ) = β1 ,
s3 (β3 ) = β2 ,
···
st−1 (βt−1 ) = βt−2 .
We also have that β0 is negative, and βt−1 is positive. Let k be the smallest
integer in {1, . . . , t − 1} such that βk is positive. Consider sk (βk ) = βk−1 . By
the choice of k, sk (βk ) = βk−1 must be negative. Recalling that sk = sαk , by
Lemma 8.6.5 we must have βk = αk . This means that
By Lemma 8.7.3,
s = sα1 · · · sαt
Proof. If t = 1 then s = sα1 , and s(α1 ) = −α1 is negative. We may thus assume
that t ≥ 2. Assume that s(αt ) is positive; we will obtain a contradiction. Now
Weyl chambers in V ∼
i: −→ Bases for R
with respect to R
from Proposition 8.5.2. These actions are transitive. If B is a base for R, then
the Weyl group W is generated by the reflections sα for α ∈ B. The stabilizer
of any point is trivial.
Proof. Let s ∈ W. If B is a base for R, then it is clear that s(B) is a base for
R. Let C be a Weyl chamber of V with respect to R. Let v ∈ C. By Lemma
8.5.1, we have
It follows that
Since
Hence,
Let s ∈ W 0 be such that (s(w), δ) is maximal. We claim that (s(w), α) > 0 for
all α ∈ B. To see this, let α ∈ B. Since sα s is also in W 0 we have, by the
maximality of (s(w), δ),
That is,
(s(w), δ) ≥ (s(w), δ) − (s(w), α).
This implies that (s(w), α) ≥ 0. If (s(w), α) = 0, then (w, s−1 (α)) = 0; this
is impossible since s−1 (α) is a root and w is regular. Thus, (s(w), α) > 0.
Since (s(w), α) > 0 for all α ∈ B it follows that s(w) ∈ X(v). This implies
110 CHAPTER 8. ROOT SYSTEMS
that s(C) = X(v), so that W 0 , and hence W, acts transitively the set of Weyl
chambers of V with respect to R. Since the bijection i is compatible with the
actions, the subgroup W 0 , and hence W, also acts transitively on the set of bases
of R.
Let B be a base for R, and as above, let W 0 be the subgroup of W generated
by the sα for α ∈ B. To prove that W = W 0 it suffices to prove that if α ∈ R,
then sα ∈ W 0 . Let α ∈ R. By Lemma 8.6.2, there exists a base B 0 for R such
that α ∈ B 0 . By what we have already proven, there exists s ∈ W 0 such that
s(B 0 ) = B. In particular, s(α) = β for some β ∈ B. Now by Lemma 8.7.3,
sβ = ss(α) = ssα s−1 ,
which implies that sα = s−1 sβ s. Since s−1 sβ s ∈ W 0 , we get sα ∈ W 0 , as desired.
Finally, suppose that B is a base for R and that s ∈ W is such that s(B) = B.
Assume that s 6= 1; we will obtain a contradiction. Write s = sα1 · · · sαt with
α1 , . . . , αt ∈ B and t ≥ 1 minimal. By Proposition 8.7.5, s(αt ) is negative with
respect to B. This contradicts s(αt ) ∈ B.
Let V be a finite-dimensional vector space over R equipt with an inner
product (·, ·), let R be a root system in V , and let W be the Weyl group of R.
Let s ∈ W with s 6= 1, and write
s = sα1 · · · sαt
with α1 , . . . , αt ∈ B and t minimal. We refer to such an expression for s as
reduced, and define the length of s to be the positive integer `(s) = t. We
define `(1) = 0.
Proposition 8.7.7. Let V be a finite-dimensional vector space over R equipt
with an inner product (·, ·), let R be a root system in V , and let W be the Weyl
group of R. Let s ∈ W. The length `(s) is equal to the number of positive roots
α such that s(α) is negative.
Proof. For r ∈ W let n(r) be the number of positive roots α such that r(α) is
negative. We need to prove that `(s) = n(s). We will prove this by induction on
`(s). Assume first that `(s) = 0. Then necessarily s = 1. Clearly, n(1) = 0. We
thus have `(s) = n(s). Assume now that `(s) > 0 and that `(r) = n(r) for all
r ∈ W with `(r) < `(s). We need to prove that `(s) = n(s). Let s = sα1 · · · sαt
be a reduced expression for s. Set s0 = ssαt . Evidently, `(s0 ) = `(s) − 1. By
Lemma 8.6.5,
s(R+ − {αt }) = s0 (sαt (R+ − {αt }))
= s0 (R+ − {αt }).
Also, by Proposition 8.7.5, s(αt ) is negative. Since
s(αt ) = s0 (sαt (αt ))
= −s0 (αt )
we see that s0 (αt ) is positive. It follows that n(s0 ) = n(s) − 1. By the induction
hypothesis, `(s0 ) = n(s0 ). This implies now that `(s) = n(s), as desired.
8.7. THE WEYL GROUP 111
3α + 2β
v
β α+β 2α + β 3α + β
30◦ 30◦ ◦
30◦ 30
30◦ 30◦
−α α
30◦ 30◦
30◦ 30◦
30◦ 30◦
−3α − β −2α − β −α − β −β
−3α − 2β
We consider bases, Weyl chambers, and the Weyl group for the root system
G2 , which appears in the above diagram. Define the vector v as in the diagram.
Then v ∈ Vreg (G2 ). By definition, R+ (v) consists of the roots that form a
strictly acute angle with v, i.e.,
By definition, R− (v) consists of the roots that form a strictly obtuse angle with
v, that is:
s1 = sα , s2 = sβ .
We know that W is generated by the two elements s1 and s2 which each have
order two. This means that W is a dihedral group (the definition of a dihedral
112 CHAPTER 8. ROOT SYSTEMS
and
(s1 s2 )(β) = sα sβ (β)
= −sα (β)
= −3α − β.
s1 s2 s1 s2 s1 s2 s1 s2 s1 s2 s1 s2 = 1.
s1 rs−1 −1
1 = s1 (s1 s2 )s1
= s2 s1
= s−1 −1
2 s1
= (s1 s2 )−1
= (s1 s2 )5
= r5
= r−1 .
We have
W = hs1 s2 i o hs1 i = hri o hs1 i
The elements of W are:
1, s1 ,
r = s1 s2 , s2 ,
r2 = s1 s2 s1 s2 , s2 s1 s2 ,
r3 = s1 s2 s1 s2 s1 s2 , s2 s1 s2 s1 s2 ,
r4 = s2 s1 s2 s1 , s1 s2 s1 s2 s1 ,
r5 = s2 s1 , s1 s2 s1 .
8.7. THE WEYL GROUP 113
In the ordered basis α, β the linear maps s1 , s2 and r have the matrices
−1 3 1 0 −1 3 1 0 2 −3
s1 = , s2 = , r = s1 s2 = = .
0 1 1 −1 0 1 1 −1 1 −1
Using these matrices, it is easy to calculate that:
α 7→ α,
α 7→ −α,
3α + β →
7 3α + β, 3α + β 7→ β,
2α + β 7→ 2α + β, 2α + β 7→ α + β,
1: s1 :
3α + 2β →
7 3α + 2β,
3α + 2β 7→ 3α + 2β,
α + β →
7 α + β, α + β 7→ 2α + β,
β 7→ β, β 7→ 3α + β,
α 7→ 2α + β,
α 7→ α + β,
3α + β →
7 3α + β, 3α + β 7→ 3α + 2β,
2α + β 7→ α + β, 2α + β 7→ 2α + β,
r: s1 r = s2 :
3α + 2β →
7 β,
3α + 2β 7→ 3α + β,
α + β →
7 −α, α + β 7→ α,
β 7→ −3α − β, β 7→ −β,
α 7→ α + β,
α 7→ 2α + β,
3α + β →
7 β, 3α + β 7→ 3α + β,
2α + β 7→ −α, 2α + β 7→ α,
2 2
r : s1 r :
3α + 2β →
7 −3α − β,
3α + 2β 7→ −β,
α + β →
7 −2α − β, α + β 7→ −α − β,
β 7→ −3α − 2β, β 7→ −3α − 2β,
α 7→ −α,
α 7→ α,
3α + β →
7 −3α − β, 3α + β 7→ −β,
2α + β 7→ −2α − β, 2α + β 7→ −α − β,
3 3
r : s1 r :
3α + 2β →
7 −3α − 2β,
3α + 2β 7→ −3α − 2β,
α + β →
7 −α − β, α + β 7→ −2α − β,
β 7→ −β, β 7→ −3α − β,
α
7→ −2α − β, α
7→ −α − β,
3α + β →
7 −3α − 2β, 3α + β 7→ −3α − 2β,
2α + β 7→ −α − β, 2α + β 7→ −2α − β,
4 4
r : s1 r :
3α + 2β →
7 −β,
3α + 2β 7→ −3α − β,
α + β →
7 α, α + β 7→ −α,
β 7→ 3α + β, β 7→ β,
α
7→ −α − β, α
7→ −2α − β,
3α + β →
7 −β, 3α + β 7→ −3α − β,
2α + β 7→ α, 2α + β 7→ −α,
5 5
r : s1 r :
3α + 2β →
7 3α + β,
3α + 2β 7→ β,
α + β →
7 2α + β, α + β 7→ α + β,
β 7→ 3α + 2β, β 7→ 3α + 2β.
Using this and that Proposition 8.7.7, we can calculate the length of each ele-
ment of W. We see that the expressions of the elements of W in the list from
114 CHAPTER 8. ROOT SYSTEMS
By Lemma 8.6.2 and Theorem 8.7.6, for every α ∈ R there exists s ∈ W such
that s(α) ∈ B. It follows that R = R1 ∪ R2 .
To prove (R1 , R2 ) = 0 we need to introduce some subgroups of W. Let W1
be the subgroup of W generated by the sα with α ∈ B1 , and let W2 be the
subgroup of W generated by the sα with α ∈ B2 . We claim that the elements
of W1 commute with the elements of W2 . To prove this, it suffices to verify
that sα1 sα2 = sα2 sα1 for α1 ∈ B1 and α2 ∈ B2 . Let α1 ∈ B1 and α2 ∈ B2 . Let
α ∈ B1 . Then
And
Thus, (sα1 sα2 )(α) = (sα2 sα1 )(α). A similar argument also shows that this
equality holds for α ∈ B2 . Since B = B1 ∪ B2 and B is a vector space basis
for V , we have sα1 sα2 = sα2 sα1 as claimed. By Theorem 8.7.6 the group W is
generated by the subgroups W1 and W2 , and by the commutativity property
that we have just proven, if s ∈ W, then there exist s1 ∈ W1 and s2 ∈ W2 such
that s = s1 s2 = s2 s1 . Now let α ∈ R1 . By definition, there exists s ∈ W and
α1 ∈ R1 such that α = s(α1 ). Write s = s1 s2 with s1 ∈ W1 and s2 ∈ W2 .
116 CHAPTER 8. ROOT SYSTEMS
Here, we use that B is also a vector space basis for V . We define a relation
on R by
v1 v2
if and only if
c1 (γ) ≥ c2 (γ) for all γ ∈ B.
The relation is a partial order on V . Evidently,
R+ = {α ∈ R : α 0} and R− = {α ∈ R : α ≺ 0}.
8.8.2 we must have (B1 , B2 ) 6= 0. Proposition 8.4.9 and Lemma 8.4.4 imply that
(α1 , α2 ) ≤ 0 for all α1 ∈ B1 and α ∈ B2 . For α2 ∈ B2 we have
X X
(β, α2 ) = b(α)(α, α2 ) = b(α1 )(α1 , α2 )
α∈B α1 ∈B1
where each term is less than or equal to zero. Since (B1 , B2 ) 6= 0, there exist
α10 ∈ B1 and α20 ∈ B2 such that (α10 , α20 ) 6= 0, so that (α10 , α20 ) < 0. This implies
that (β, α20 ) < 0. By Lemma 8.3.4, either β = ±α20 or β + α20 is a root. Assume
that β = α20 . Then (β, α20 ) = (β, β) > 0, contradicting (β, α20 ) < 0. Assume
that β = −α20 . Then b(α20 ) = −1 < 0, a contradiction. It follows that β + α20 is
a root. Now β + α20 β. Since β is maximal, we have β + α20 = β. This means
that α20 = 0, a contradiction. It follows that B2 is empty, so that b(α) > 0 for all
α ∈ B. Arguing similarly, we also see that (β, α) ≥ 0 for all α ∈ B (if (β, α) < 0
for some α ∈ B, then β + α is a root, which contradicts the maximality of β).
Since B is a basis for V we cannot have (β, B) = 0; hence, there exists α0 ∈ B
such that (β, α0 ) > 0.
Now suppose that β 0 is another maximal root. Write
X
β0 = b0 (α)α.
α∈B
0
As in the last paragraph, b (α) > 0 for all α ∈ B. Now
X
(β, β 0 ) = b0 (α0 )(β, α).
α∈B
0
As (β, α) ≥ 0 and b (α) > 0 for all α ∈ B, and (β, α0 ) > 0, we see that
(β, β 0 ) > 0. By Lemma 8.3.4, either β = β 0 , β = −β 0 or β − β 0 is a root.
Assume that β = −β 0 . Then b(α) = −b0 (α) for α ∈ B; this contradicts the fact
that b(α) and b(α0 ) are positive for all α ∈ B. Assume that β − β 0 is a root.
Then either β − β 0 0 or β − β 0 ≺ 0. Assume that β − β 0 0. Then β β 0 ,
which implies β = β 0 by the maximality of β 0 . Therefore, β − β 0 = 0; this is not
a root, and hence a contradiction. Similarly, the assumption that β − β 0 ≺ 0
leads to a contradiction. We conclude that β = β 0 .
Lemma 8.8.4. Let V be a finite-dimensional vector space over R equipt with
an inner product (·, ·), and let R ⊂ V be a root system. Assume that R is
irreducible. Let B be a base for R. Let β be the maximal root of R with respect
to B. We have β α for all α ∈ R, α 6= β. Also, if α ∈ B, then (β, α) ≥ 0.
Proof. Let α ∈ R with α 6= β. Since α 6= β, α is not maximal by Lemma 8.8.3.
It follows that there exists γ1 ∈ R such that γ1 α and γ1 6= α. If γ1 = β, then
β α. Assume γ1 6= β. Since γ1 6= β, γ1 is not maximal by Lemma 8.8.3. It
follows that there exists γ2 ∈ R such that γ2 γ1 and γ2 6= γ1 . If γ2 = β, then
β γ1 α, so that β α. If γ2 6= β, we continue to argue in the same fashion.
Since R is finite, we eventually conclude that β α.
Let α ∈ B. Assume that (α, β) < 0. Then certainly α 6= β. Also, we cannot
have α = −β because β is a positive root with respect to B by Lemma 8.8.3.
By Lemma 8.3.4, α + β is a root. This contradicts the maximality of β.
118 CHAPTER 8. ROOT SYSTEMS
R1 = {α ∈ R : α ∈ U } and R2 = {α ∈ R : α ∈ U ⊥ }.
Proof. Suppose that there exist α1 , α2 , α3 ∈ R such that kα1 k < kα2 k < kα3 k;
we will obtain a contradiction.
We first assert that there exist roots α10 , α20 , α30 ∈ R such that
and
(α10 , α20 ) 6= 0, (α20 , α30 ) 6= 0, (α10 , α30 ) 6= 0.
To see this we note that by Lemma 8.8.5, the vectors s(α2 ) for s ∈ W span V ;
it follows that there exists s ∈ W such that (α1 , s(α2 )) 6= 0. Similarly, there
exists r ∈ W such that (s(α2 ), r(α3 )) 6= 0. If (α1 , r(α3 )) 6= 0, we define
and these vectors have the desired properties. Assume that (α1 , r(α3 )) = 0. In
this case we define
We have
And
kα20 k kα30 k
1< < .
kα10 k kα10 k
Applying Lemma 8.3.3 to the pair α10 and α20 , and the pair α10 and α30 , and
taking note of the above inequalities, we must have
kα20 k √ kα30 k √
= 2 and = 3.
kα10 k kα10 k
This implies that √
kα30 k 3
= √ .
kα20 k 2
√ √
However, Lemma 8.3.3 applied to the pair α20 and α30 implies that 3/ 2 is not
an allowable value for kα30 k/kα20 k. This is a contradiction.
Assume that α, β ∈ R have the same length. Arguing as in the last para-
graph, there exists s ∈ W such that (s(α), β) 6= 0. If s(α) = β, then s is
the desired element of W. If s(α) = −β, then (sβ s)(α) = β, and sβ s is the
desired element. Assume that s(α) 6= ±β. Since s(α) and β have the same
length, we have by Lemma 8.3.3 that hs(α), βi = hβ, s(α)i = ±1. Assume that
hs(α), βi = 1. We have
Proof. Let α ∈ R. We need to prove that (β, β) ≥ (α, α). By Proposition 8.4.9
there exists v ∈ Vreg (R) such that B = B(v). Let C be the Weyl chamber
containing v. By Lemma 8.5.1 we have
and
C̄ 0 = {w ∈ V : (w, α) ≥ 0 for all α ∈ B 0 }.
By Theorem 8.7.6 there exists s in the Weyl group of R such that s(C 0 ) = C
and s(B 0 ) = B. It follows that s(C̄ 0 ) = C̄. Replacing α with s(α) (which has
the same length as α), we may assume that α ∈ C̄. By Lemma 8.8.4 we also
have β ∈ C̄. Next, by Lemma 8.8.4, we have β α. This means that
X
β−α= c(γ)γ
γ∈B
(α, β − α) ≥ 0, (β, β − α) ≥ 0.
1. φ(R1 ) = R2 .
123
124 CHAPTER 9. CARTAN MATRICES AND DYNKIN DIAGRAMS
= sφ(α) (α0 ).
It follows that φ◦sα ◦φ−1 = sφ(α) is contained in W2 , so that the map W1 → W2
is well-defined. This map is evidently a homomorphism of groups. The map
W2 → W1 defined by s0 7→ φ−1 ◦ s0 ◦ φ is also a well-defined homomorphism and
is the inverse of W1 → W2 .
Let V be a finite-dimensional vector space over R equipt with an inner
product (·, ·), and let R ⊂ V be a root system. If φ : V → V is an isomorpism
from R to R then we say that φ is an automorphism of R.
Lemma 9.1.2. Let V be a finite-dimensional vector space over R equipt with
an inner product (·, ·), and let R ⊂ V be a root system. A function φ : V → V is
an automorphism of R if and only if φ is an R vector space isomorphism from
V to V , and φ(R) = R. The set of automorphisms of R forms a group Aut(R)
under composition of functions. The Weyl group W of R is a normal subgroup
of Aut(R).
Proof. Let φ : V → V be a function. If φ is an automorphism of R, then φ is
a vector space isomorphism from V to V and φ(R) = R by definition. Assume
that φ is a vector space isomorphism from V to V and φ(R) = R. By Lemma
8.7.3 we have hφ(α), φ(β)i = hα, βi for all α, β ∈ R. It follows that φ is an
automorphism of R. It is clear that Aut(R) is a group under composition of
functions, and that W is a subgroup of Aut(R). To see that W is normal in
Aut(R), let α, β ∈ R and φ ∈ Aut(R). Then
(φ ◦ sα ◦ φ−1 )(β) = φ sα (φ−1 (β)
= sφ(α) (β).
Since R contains a basis for V this implies that φ ◦ sα ◦ φ−1 = sφ(α) . It follows
that W is normal in Aut(R).
Proof. By Theorem 8.7.6 there exists an element s in the Weyl group of R such
that B 0 = s(B). Since B 0 = s(B), there exists a t × t permutation matrix P
such that P −1 · C(α10 , . . . , αt0 ) · P = C(s(α1 ), . . . , s(αt )). Now
Proof. Assume that R and R0 have the same Cartan matrices. Then V and
V 0 have the same dimension t, and there exists bases B = {α1 , . . . , αt } and
B 0 = {α10 , . . . , αt0 } for R1 and R2 , respectively, such that C(α1 , . . . , αt ) =
C(α10 , . . . , αt0 ). Define φ : V1 → V2 by φ(αi ) = αi0 for i ∈ {1, . . . , t}. We
need to prove that φ(R) = R0 and that hφ(α), φ(β)i = hα, βi for α, β ∈ R. Let
α, β ∈ B. Since C(α1 , . . . , αt ) = C(α10 , . . . , αt0 ) we have hφ(β), φ(α)i = hβ, αi.
Therefore,
s = sδ1 · · · sδn .
Let β ∈ R. Repeatedly using the identity we have already proved, we find that:
φ(s(β)) = φ (sδ1 · · · sδn )(β)
= sφ(δ1 ) φ (sδ2 · · · sδn )(β)
= sφ(δ1 ) sφ(δ2 ) φ (sδ3 · · · sδn )(β)
···
φ(s(β)) = sφ(δ1 ) · · · sφ(δn ) φ(β) .
Again let β ∈ R. By Lemma 8.6.2 and Theorem 8.7.6, there exists s in the Weyl
0
group of R such that s(β) ∈ B. We have φ(s(β)) ∈ B . Write s as a0 product,
as above. Then φ(s(β)) = s φ(δ1 ) · · · sφ(δn ) φ(β) . Since φ(s(β)) ∈ B , we have
0
sφ(δ1 ) · · · sφ(δn ) φ(β) ∈ B . Applying the inverse of sφ(δ1 ) · · · sφ(δn ) , we see that
φ(β) ∈ R0 . Thus, φ(R) ⊂ R0 . A similar argument implies that φ(R0 ) ⊂ R, so
that φ(R) = R0 .
We still need to prove that hφ(α), φ(β)i = hα, βi for α, β ∈ R. By the
definition of φ, and since C(α1 , . . . , αt ) = C(α10 , . . . , αt0 ), we have hφ(α), φ(β)i =
hα, βi for α, β ∈ B. Since this formula is linear in α, the formula holds for all
α ∈ R and β ∈ B. Let β be an arbitrary element of R. As before, there
exists s in the Weyl group of R such that s(β) ∈ B, and δ1 , . . . , δn such that
δ1 , . . . , δn ∈ B and s = sδ1 · · · sδn . Let α ∈ R. Then
60◦
◦
60 60◦
−α α
60◦ 60◦
60◦
−α − β −β
hα, αi hα, βi 2 −1
Cartan matrix: = .
hβ, αi hβ, βi −1 2
45◦ 45◦
45◦ 45◦
−α α
◦ ◦
45 45
45◦ 45◦
−2α − β −α − β −β
hα, αi hα, βi 2 −1
Cartan matrix: = .
hβ, αi hβ, βi −2 2
128 CHAPTER 9. CARTAN MATRICES AND DYNKIN DIAGRAMS
3α + 2β
β α+β 2α + β 3α + β
30◦ 30◦ ◦
30◦ 30
30◦ 30◦
−α α
30◦ 30◦
30◦ 30◦
30◦ 30◦
−3α − β −2α − β −α − β −β
−3α − 2β
hα, αi hα, βi 2 −1
Cartan matrix: = .
hβ, αi hβ, βi −3 2
90◦ 90◦
−α α
◦ ◦
90 90
−β
hα, αi hα, βi 2 0
Cartan matrix: = .
hβ, αi hβ, βi 0 2
9.3. DYNKIN DIAGRAMS 129
lines; recall that in Lemma 8.3.2 we proved that dαβ is in {0, 1, 2, 3}, and that
dαβ was computed in more detail in Lemma 8.3.3. By Lemma 8.3.3, if dαβ > 1,
then α and β have different lengths; in this case, we draw an arrow pointing
to the shorter root. We will also sometimes consider another graph associated
to R. This is called the Coxeter graph, and consists of the Dynkin diagram
without the arrows pointing to shorter roots.
We have the following of examples of Dynkin diagrams:
Proof. Assume that R and R0 have the same directed Dynkin diagrams. Since
R and R0 have same directed Dynkin diagrams it follows that R and R0 have
bases B = {α1 , . . . , αt } and B 0 = {α10 , . . . , αt0 }, respectively, such that for i, j ∈
{1, . . . , t},
dij = hαi , αj ihαj , αi i = hαi0 , αj0 ihαj0 , αi0 i
and if dij > 1, then kαj k > kαi k and kαj0 k > kαi0 k (note that if i, j ∈ {1, . . . , t},
then hαi , αj i = hαj , αi i = hαi0 , αj0 i = hαj0 , αi0 i = 2). Let i, j ∈ {1, . . . , t}. We
claim that hαi , αj i = hαi0 , αj0 i and hαj , αi i = hαj0 , αi0 i. If i = j, then this is
clear by the previous comment. Assume that i 6= j. By Lemma 8.4.4, the angle
130 CHAPTER 9. CARTAN MATRICES AND DYNKIN DIAGRAMS
between αi and αj , and the angle between αi0 and αj0 , are obtuse. By Lemma
8.3.2 we have dij = 0, 1, 2 or 3. Assume that dij = 0. By Lemma 8.3.3 we
have hαi , αj i = hαj , αi i = hαi0 , αj0 i = hαj0 , αi0 i = 0. Assume that dij = 1. By
Lemma 8.3.3 we have hαi , αj i = hαi0 , αj0 i = −1 and hαj , αi i = hαj0 , αi0 i = −1.
Assume that dij = 2. By Lemma 8.3.3 we have hαi , αj i = hαi0 , αj0 i = −1 and
hαj , αi i = hαj0 , αi0 i = −2. Assume that dij = 3. By Lemma 8.3.3 we have
hαi , αj i = hαi0 , αj0 i = −1 and hαj , αi i = hαj0 , αi0 i = −3. Our claim follows. We
now have an equality of Cartan matrices:
which implies
n
X
n> −2(vi , vj ).
i,j=1, i<j
Let N be the number of pairs {vi , vj }, i, j ∈ {1, . . . , n}, i 6= j, that are joined
by at least one edge, i.e., for which dij ≥ 1. We have
n
X p
dij ≥ N.
i,j=1, i<j
1. k = 1 and
v v1
2. k = 1 and
v v1
3. k = 1 and
v v1
4. k = 2 and
v1
v v2
5. k = 2 and
v1
v v2
6. k = 3 and
v1
v2
v v3
9.4. ADMISSIBLE SYSTEMS 133
Proof. By Lemma 9.4.2, ΓA does not contain a cycle; this implies that (vi , vj ) =
0 for i, j ∈ {1, . . . , k} with i 6= j. Consider the subspace U of V spanned by the
linearly independent vectors v1 , . . . , vk , v. There exists a vector v0 ∈ U such that
v0 , v1 , . . . , vk is a basis for U , (v0 , v0 ) = 1, and (v0 , vi ) = 0 for i ∈ {1, . . . , k}. It
follows that v0 , v1 , . . . , vk is an orthonormal basis for U . Now
k
X
v= (v, vi )vi .
i=0
It follows that
k
X k
X
(v, v) = ( (v, vi )vi , (v, vj )vj )
i=0 j=0
k X
X k
= (v, vi )(v, vj )(vi , vj )
i=0 j=0
k
X
= (v, vi )2 .
i=0
Then v ∈
/ A. Define
A0 = A − {v1 , . . . , vk } ∪ {v}.
Proof. Since the set A is linearly independent and since k ≥ 2, we must have
v ∈/ A. Similarly, the set A0 is linearly independent. To show that property
2 of the definition of an admissible system is satisfied by A0 it will suffice to
prove that (v, v) = 1. Now by assumption we have that 4(vi , vi+1 )2 = 1 for
i ∈ {1, . . . , k − 1}, or equivalently, (vi , vi+1 ) = −1/2 for i ∈ {1, . . . , k − 1}. Also,
by assumption, (vi , vj ) = 0 for i, j ∈ {1, . . . , k} i < j and j 6= i + 1. We obtain:
k
X k
X
(v, v) = ( vi , vj )
i=1 j=1
k X
X k
= (vi , vj )
i=1 j=1
k
X k−1
X
= (vi , vi ) + 2 (vi , vi+1 )
i=1 i=1
k
X k−1
X
= 1+2 (−1/2)
i=1 i=1
= k − (k − 1)
= 1.
that (w, vj ) = 0 for all j ∈ {1, . . . , k} with j 6= i. We now have (w, v) = (w, vi ),
so that 4(w, v)2 = 4(w, vi )2 ∈ {0, 1, 2, 3}, as desired.
Finally, consider ΓA0 . To see that ΓA0 is obtained from ΓA by shrinking the
above line to the single vertex v it suffices to see that, for all i ∈ {1, . . . , k}, if
there is an edge in ΓA between vi and a vertex w with w ∈ / {v1 , . . . , vk }, then
w is not incident to vj for all j ∈ {1, . . . , k} with i 6= j; this was proven in the
last paragraph.
Proof. By Lemma 9.4.4 we may assume that ΓA does not contain a triple edge.
Proof of 1. Assume that ΓA has at least two double edges; we will obtain
a contradiction. Since ΓA is connected, for every pair of double edges there
exists at least one path joining a vertex of one double edge to a vertex of the
other double edge; moreover, any such joining path must have at least one edge
by Lemma 9.4.3. Chose a pair such that the length of the joining path is the
shortest among all joining paths between pairs of double edges. Let v1 , . . . , vk
be the vertices on this shortest path, with v1 on the first double edge, vk on
the second double edge, and vi joined to vi+1 for i ∈ {1, . . . , k − 1} by at least
one edge. Since this is the shortest path we cannot have vi and vj joined by
an edge for some i, j ∈ {1, . . . , k}, i < j, and j 6= i + 1. Also, as this is the
shortest choice, it is not the case that vi is joined to vi+1 by a double edge for
i ∈ {1, . . . , k − 1}. Let A0 be as in Lemma 9.4.5; by Lemma 9.4.5, A0 is an
admissible system. It follows that
Then
k(k + 1)
(v, v) = .
2
Proof. Since the number of edges between vi and vi+1 is one for i ∈ {1, . . . , k −
1} it follows that 4(vi , vi+1 )2 = 1, so that (vi , vi+1 ) = −1/2 (recall that by
the definition of an admissible system we have (vi , vi+1 ) ≤ 0). Also, we have
(vi , vj ) = 0 for i, j ∈ {1, . . . , k} with i < j and j 6= i + 1. It follows that
k
X k
X
(v, v) = ( i · vi , j · vj )
i=1 j=1
k
X k−1
X
= i2 (vi , vi ) + 2 i(i + 1)(vi , vi+1 )
i=1 i=1
k
X k−1
X
= i2 + 2(−1/2) (i2 + i)
i=1 i=1
k−1
X k−1
X k−1
X
2 2 2
=k + i − i − i
i=1 i=1 i=1
k−1
X
= k2 − i
i=1
(k − 1)k
= k2 −
2
2k 2 − k 2 + k
=
2
k(k + 1)
= .
2
This completes the calculation.
,
,
,
···
Proof. By Lemma 9.4.6, since ΓA has a double edge, ΓA has exactly one double
edge, ΓA has no triple edge, and ΓA does not contain a branch vertex. It follows
that ΓA has the form
... ...
v1 v2 vk wj wj−1 w1
with no other edges between the shown vertices; here k ≥ 1 and j ≥ 1. Without
loss of generality we may assume that k ≥ j. Define
k
X j
X
v= i · vi , w= i · wi .
i=1 i=1
k(k + 1) j(j + 1)
(v, v) = , (w, w) = .
2 2
We have 4(vk , wj )2 = 2 since there is a double edge joining vk and vj , and
(vi , w` ) = 0 since no edge joins vi and w` for all i ∈ {1, . . . , k} and ` ∈ {1, . . . , j}
with i 6= k or ` 6= j. It follows that
Xk j
X
(v, w) = ( i · vi , ` · w` )
i=1 `=1
= kj(vk , wj ),
so that
k2 j 2
(v, w)2 = k 2 j 2 (vk , wj )2 = .
2
By the Cauchy-Schwarz inequality we have
Note that v and w are linearly independent, so that the inequality is strict.
Substituting, we obtain:
k2 j 2 k(k + 1) j(j + 1)
< ,
2 2 2
2 2
2k j < k(k + 1)j(j + 1),
138 CHAPTER 9. CARTAN MATRICES AND DYNKIN DIAGRAMS
2k 2 j 2 < k 2 j 2 + jk 2 + j 2 k + jk,
2kj < kj + k + j + 1,
kj < k + j + 1,
kj − k − j < 1,
kj − k − j + 1 < 2,
(k − 1)(j − 1) < 2.
D` , ` ≥ 4 : ...
or
E6 :
or
E7 :
or
E8 : .
Proof. By Lemma 9.4.4 and Lemma 9.4.6, since ΓA is connected and contains
a double edge, ΓA contains exactly one branch vertex, no double edges, and no
triple edges. It follows that ΓA has the form
9.4. ADMISSIBLE SYSTEMS 139
u1
..
.
u`
... ...
v1 vk z w1 wj
with k ≥ j ≥ `. We define
k
X j
X `
X
v= i · vi , w= i · wi , u= i · ui .
i=1 i=1 i=1
Since there are no edges between the vertices in {v1 , . . . , vk } and the vertices
in {w1 , . . . , vj }, the vectors v and w are orthogonal. Similarly, v and u are
orthogonal, and w and u are orthogonal. Define
v w u
v0 = , w0 = , u0 = .
kvk kwk kuk
The vectors v 0 , w0 and u0 are also mutually orthogonal, and have norm one.
Let U be the subspace of V spanned by v 0 , w0 , u0 and z. This space is four-
dimensional as these vectors are linearly independent. The orthonormal vectors
v 0 , w0 , u0 can be extended to an orthonormal basis v 0 , w0 , u0 , z 0 for U . We have
so that
1 = (z, z) = (z, v 0 )2 + (z, w0 )2 + (z, u0 )2 + (z, z 0 )2 .
The vector z 0 cannot be orthogonal to z; otherwise, (z 0 , U ) = 0, a contradiction.
Since (z, z 0 )2 > 0, we obtain
Now
(z, v)2
(z, v 0 )2 =
(v, v)
Pk
2(z, i=1 ivi )2
=
k(k + 1)
2k (z, vk )2
2
=
k(k + 1)
k
= .
2(k + 1)
140 CHAPTER 9. CARTAN MATRICES AND DYNKIN DIAGRAMS
Similarly,
j `
(z, w0 )2 = and (z, u0 )2 = .
2(j + 1) 2(` + 1)
Substituting, we get:
k j `
+ + < 1,
2(k + 1) 2(j + 1) 2(` + 1)
k+1 1 j+1 1 `+1 1
− + − + − < 1,
2(k + 1) 2(k + 1) 2(j + 1) 2(j + 1) 2(` + 1) 2(` + 1)
1 1 1 1 1 1
− + − + − < 1,
2 2(k + 1) 2 2(j + 1) 2 2(` + 1)
3 1 1 1
− − − < 1,
2 2(k + 1) 2(j + 1) 2(` + 1)
1 1 1
3− − − < 2,
k+1 j+1 `+1
1 1 1
+ + > 1.
k+1 j+1 `+1
Now k ≥ j ≥ ` ≥ 1. Hence,
k+1≥j+1≥`+1≥2
and thus
1 1 1 1
≤ ≤ ≤ .
k+1 j+1 `+1 2
It follows that
1 1 1
+ + > 1,
k+1 j+1 `+1
1 1 1
+ + > 1,
`+1 `+1 `+1
3
> 1,
`+1
3 > ` + 1,
2 > `.
2 1
> ,
j+1 2
3 > j.
1. (` vertices, ` ≥ 1) ...
2. (` vertices, ` ≥ 2) ...
3. (` vertices, ` ≥ 3) ...
4.
5.
6.
7.
8. .
142 CHAPTER 9. CARTAN MATRICES AND DYNKIN DIAGRAMS
A` , ` ≥ 1: ...
B` , ` ≥ 2: ... i
C` , ` ≥ 3: ... h
D` , ` ≥ 4: ...
G2 : i
F4 : i
E6 :
E7 :
9.5. POSSIBLE DYNKIN DIAGRAMS 143
E8 : .
Proof. Let B a base for R. Let A be the admissible system associated to R and
B as at the beginning of Section 9.4. Let C be the Coxeter graph of R; this is
the same as ΓA , the graph associated to A. By Theorem 9.4.10, ΓA = C must
be one of the graphs listed in this theorem. This implies the result.
144 CHAPTER 9. CARTAN MATRICES AND DYNKIN DIAGRAMS
Chapter 10
Let F have characteristic zero and be algebraically closed. The classical Lie
algebras over F are sl(` + 1, F ), so(2` + 1, F ), sp(2`, F ), and so(2`, F ) for ` a
positive integer. In this chapter we will prove that these Lie algebras are simple
(with the exception of so(2`, F ) when ` = 1 or ` = 2) We will also determine
the root systems associated to these classical Lie algebras.
10.1 Definitions
sl(` + 1, F )
Let F have characteristic zero and be algebraically closed, and let ` be a positive
integer. We define sl(` + 1, F ) to be the F -subspace of g ∈ gl(` + 1, F ) such that
tr(g) = 0. The bracket on sl(`+1, F ) is inherited from gl(`+1, F ), and is defined
by [X, Y ] = XY − Y X for X, Y ∈ sl(` + 1, F ). Note that [X, Y ] ∈ sl(` + 1, F )
for X, Y ∈ sl(` + 1, F ) because tr([X, Y ]) = tr(XY ) − tr(Y X) = XY − XY = 0.
The bracket on sl(` + 1, F ) satisfies 1 and 2 of the definition of Lie algebra
from Section 1.3 because the bracket on gl(` + 1, F ) satisfies these properties by
Proposition 1.4.1.
Lemma 10.1.1. Let n be a positive integer. Let S ∈ gl(n, F ). Let L be the
F -subspace of X ∈ gl(n, F ) such that
t
XS + SX = 0.
145
146 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
= (t Y t X − t X t Y )S + SXY − SY X
= t Y t XS − t X t Y S + SXY − SY X
= −t Y SX + t XSY + SXY − SY X
= SY X − SXY + SXY − SY X
= 0.
so(2` + 1, F )
Let F have characteristic zero and be algebraically closed, and let ` be a positive
integer. Let S ∈ gl(2` + 1, F ) be the matrix
1
S= 1` .
1`
sp(2`, F )
Let F have characteristic zero and be algebraically closed, and let ` be a positive
integer. Let S ∈ gl(2`, F ) be the matrix
1`
S= .
−1`
Here, 1` is the `×` identity matrix. We define sp(2`, F ) to be the Lie subalgebra
of gl(2`, F ) defined by S as in Lemma 10.1.1. By Lemma 10.1.1, since S is
invertible, we have sp(2`, F ) ⊂ sl(2`, F ).
10.2. A CRITERION FOR SEMI-SIMPLICITY 147
so(2`, F )
Let F have characteristic zero and be algebraically closed, and let ` be a positive
integer. Let S ∈ gl(2` + 1, F ) be the matrix
1`
S= .
1`
Here, 1` is the `×` identity matrix. We define so(2`, F ) to be the Lie subalgebra
of gl(2`, F ) defined by S as in Lemma 10.1.1. By Lemma 10.1.1, since S is
invertible, we have so(2`, F ) ⊂ sl(2`, F ).
L = [L, L] ⊕ Z(L)
Since now dim[L, L] = dim L−dim Z(L) = dim M , we conclude that [L, L] = M .
Hence, L = [L, L]⊕Z(L) as Lie algebras. Since L = [L, L]⊕Z(L) as Lie algebras
148 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
of Lemma 7.0.11,
2(α, β)
hα, βi = = α(hβ ).
(β, β)
Also, by the definition of Lα ,
α(hβ )eα = [hβ , eα ].
Consider [hβ , eα ]. On the one hand, since eα ∈ Lα ⊂ I, and since I is an ideal
of L, we have [hβ , eα ] ∈ I. On the other hand, hβ = [eβ , fβ ]; since fβ ∈ I ⊥ ,
and I ⊥ ; we must have hβ ∈ I ⊥ . Using again that I ⊥ is an ideal, we see that
[hβ , eα ] ∈ I ⊥ . Now we have [hβ , eα ] ∈ I ∩ I ⊥ = 0, proving that [hβ , eα ] = 0. It
follows from above that α(hβ ) = 0, and hence that hα, βi = 0, as claimed. This
contradicts the irreducibility of Φ.
{x ∈ W : ad(h)(x) = [h, x] = 0, h ∈ H} = 0.
L = H ⊕ W.
The operators ad(h) for h ∈ H leave the subspace W invariant; since ad(h) is
diagonalizable, it follows that ad(h)|W is diagonalizable for h ∈ H. For a linear
functional β : H → F , let
Since the subspaces Wβ for β ∈ B form a direct sum, we must have β(h)xβ = 0
for all β ∈ B and h ∈ H. Since every β ∈ B is non-zero, we must have xβ = 0
for all β ∈ B. This implies that x = x0 ∈ H, as desired.
t(x, y) = tr(xy).
And
so that
(π ∨ (x)π ∨ (y) − π ∨ (y)π ∨ (x))λ (v)
= − π ∨ (y)λ (π(x)v) + π ∨ (x)λ (π(y)v)
= λ π(y)π(x)v − λ π(x)π(y)v) .
It follows that
r1 (ad(x)v)(w) = t1 (ad(x)v, w)
= t1 ([x, v], w)
= t1 (−[v, x], w)
= t1 (v, −[x, w])
= t1 (v, −ad(x)w)
= r1 (v)(−ad(x)w)
= ad∨ (x)(r1 (v)) (w).
w`+1
156 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
Since the eij for i, j ∈ {1, . . . , ` + 1} are linearly independent, we get wij (hii −
hjj ) = 0 for all i, j ∈ {1, . . . , ` + 1} with i 6= j and all h ∈ H. Let i, j ∈
{1, . . . , ` + 1} with i 6= j. Set h = eii − ejj . Then h ∈ H, and we have
wij (hii −hjj ) = 2wij . Since F has characteristic zero, we conclude that wij = 0.
Thus, w = 0.
Lemma 10.7.4. Assume that the characteristic of F is zero and F is alge-
braically closed. Let H be the Cartan subalgebra of L = sl(` + 1, F ) consisting
of diagonal matrices in sl(` + 1, F ), as in Lemma 10.7.3. Then Φ consists of
the linear forms
αij : H −→ F
defined by
αij (h) = hii − hjj
10.7. THE LIE ALGEBRA sl(` + 1) 157
Lαij = F eij
the inclusion must be an equality. This implies that Φ and Lαij for 1 ≤ i, j ≤
` + 1 with i 6= j are as claimed.
Lemma 10.7.5. Let F have characteristic zero and be algebraically closed. Let
` be a positive integer. Let H be the subalgebra of sl(` + 1, F ) consisting of
diagonal matrices; by Lemma 10.7.3, H is a Cartan subalgebra of sl(` + 1, F ).
Let Φ be the set of roots of sl(`+1, F ) defined with respect to H. Let V = R⊗Q V0 ,
where V0 is the Q subspace of H ∨ = HomF (H, F ) spanned by the elements of
Φ; by Proposition 8.2.1, Φ is a root system in V . Let i ∈ {1, . . . , `}, and define
βi : H −→ F
by
βi (h) = hii − hi+1,i+1
for h ∈ H. The set B = {β1 , . . . , β` } is a base for Φ. The positive roots in Φ
are the αij with i < j, and if i < j, then
Proof. It was proven in Lemma 10.7.4 that the linear functionals αij : H → F
defined by αij (h) = hii − hjj for h ∈ H and i, j ∈ {1, . . . , ` + 1}, i 6= j,
constitute the set of roots Φ of sl(` + 1, C) with respect to H. Evidently, B ⊂ Φ.
Also, it is clear that B is linearly independent; since B has ` elements and the
dimension of V is ` (by Proposition 7.1.2), it follows that B is a basis for V .
Let i, j ∈ {1, . . . , ` + 1}, i 6= j. Assume that i < j. Then
It follows that B is a base for Φ and the positive roots in Φ are as described.
158 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
h11 β1 = α12 α13 α14 α15
−α h22 β2 = α23 α24 α25
12
−α −α23 h33 β3 = α34 α35
13
−α14 −α24 −α34 h44 β4 = α45
−α15 −α25 −α35 −α45 h55
Figure 10.1: The root spaces in sl(5, F ). For this example, ` = 3. The positions are labeled
with the corresponding root. Note that the diagonal is our chosen Cartan subalgebra. The
positive roots with respect to our chosen base {β1 , β2 , β3 , β4 } are boxed, while the colored
roots form our chosen base. The linear functionals αij are defined in Proposition 10.7.4.
κ : sl(` + 1, F ) × sl(` + 1, F ) −→ F
is given by
κ(h, h0 ) = (2` + 2) · tr(hh0 )
for h, h0 ∈ H. Here, H is the subalgebra of diagonal matrices in sl(` + 1, F ); H
is a Cartan subalgebra of sl(` + 1, F ) by Lemma 10.7.3.
κ(h, h0 )
= tr(ad(h) ◦ ad(h0 ))
X
= α(h)α(h0 )
α∈Φ
X
= (hii − hjj )(h0ii − h0jj )
i,j∈{1,...,`+1},
i6=j
X X
= hii h0ii − hii h0jj
i,j∈{1,...,`+1}, i,j∈{1,...,`+1},
i6=j i6=j
X X
− hjj h0ii + hjj h0jj
i,j∈{1,...,`+1}, i,j∈{1,...,`+1},
i6=j i6=j
X X
= 2` hii h0ii − 2 hii h0jj
i∈{1,...,`+1} i,j∈{1,...,`+1},
i6=j
10.7. THE LIE ALGEBRA sl(` + 1) 159
X X
= 2` · tr(hh0 ) − 2 hii h0jj + 2 hii h0ii
i,j∈{1,...,`+1} i∈{1,...,`+1}
= (2` + 2) · tr(hh0 ) − 2 · 0 · 0
= (2` + 2) · tr(hh0 ),
Also,
1 1 1
κ h, (eii − ei+1,i+1 ) = κ(h, eii ) − κ(h, ei+1,i+1 )
2` + 2 2` + 2 2` + 2
2` + 2 2` + 2
= · tr(heii ) − · tr(hei+1,i+1 )
2` + 2 2` + 2
= tr(heii ) − tr(hei+1,i+1 )
= hii − hi+1,i+1 .
By definition, tβi is the unique element of H such that βi (h) = κ(h, tβi ) for all
h ∈ H. The last two equalities imply that
1
tβi = (eii − ei+1,i+1 ).
2` + 2
Let i, j ∈ {1, . . . , `}. By the definition of the inner product on V and Lemma
10.7.6 we have
1
= tr(eii ejj ) − tr(eii ej+1,j+1 )
2` + 2
− tr(ei+1,i+1 ejj ) + tr(ei+1,i+1 ej+1,j+1 )
1
= δij − δi,j+1 − δi+1,j + δi+1,j+1 .
2` + 2
The formula for (βi , βj ) follows.
Lemma 10.7.8. Let F have characteristic zero and be algebraically closed. The
Dynkin diagram of sl(` + 1, F ) is
A` : ...
Proof. By Lemma 10.5.3, there exists c ∈ F × such that κ(x, y) = ctr(xy) for
x, y ∈ sl(` + 1, F ). Let H be the subalgebra of diagonal matrices in sl(` + 1, F );
H is a Cartan subalgebra of sl(` + 1, F ) by Lemma 10.7.3. By Lemma 10.7.6 we
have κ(h, h0 ) = (2`+2)·tr(hh0 ) for h, h0 ∈ H. Hence, ctr(hh0 ) = (2`+2)·tr(hh0 )
for h, h0 ∈ H. Since there exist h, h0 ∈ H such that tr(hh0 ) 6= 0 we conclude
that c = 2` + 2.
Lemma 10.7.10. Let the notation as in Lemma 10.7.4 and Lemma 10.7.5. Let
1
i, j ∈ {1, . . . , ` + 1} with i 6= j. The length of every root is √`+1 .
Proof. Let α ∈ Φ+ . We know that α1 , . . . , α` is an ordered basis for V . By
Lemma 10.7.7 the matrix of the inner product (·, ·) in this basis is
2 −1
−1 2 −1
1 −1 2 −1
M= .
2` + 2
.. .. ..
. . .
−1 2 −1
−1 2
The coordinate vector of α in this basis has the form
0
..
.
0
1
c = ... .
1
0
.
..
0
2 1
A calculation shows that (α, α) = t cM c = 2`+2 = `+1 ; hence the length of α is
√1 .
`+1
t
c tg th 1` t
c th tg
And:
1 a b c −a −b −c
−Sx = −
1`
B
f = −C
g −G −h.
1` C G h −B −f −g
It follows that x ∈ so(2` + 1, F ) if and only if:
a = 0,
B = −t c,
C = −t b,
G = −t G,
h = −t f,
g = −t g.
Evidently, e, f and h form a vector space basis for so(3, F ), and calculations
prove that [e, f ] = h, [h, e] = 2e and [h, f ] = −2f .
10.8. THE LIE ALGEBRA so(2` + 1) 163
our claim holds in this case. We may thus assume that y 6= 0 or z 6= 0. Assume
that z 6= 0. Let g ∈ M`,` (F ) be such that −t g = g and gz 6= 0; such a g exists
by Lemma 10.8.3. Since
0 0 0 0 0 0 x 0
0 0 g v = 0 0 g y = gz ∈ V,
0 0 0 0 0 0 z 0
our claim holds in this case. We may now assume that z = 0 and y 6= 0 so that
v has the form
x
v = y .
0
164 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
proving our claim in this final case. Thus, our claim holds; that is, V contains
a vector
0
w = y
0
with y 6= 0. If f ∈ gl(`, F ), then
0 0 0 0 0 0 0 0
0 f 0 w = 0 f 0 y = f y ∈ V.
0 0 −t f 0 0 −t f 0 0
Lemma 10.8.6. Let F have characteristic zero and be algebraically closed. The
set H of diagonal matrices in so(2`+1, F ) is a Cartan subalgebra of so(2`+1, F ).
and
0 0 0 0 b c 0 0 0
hw = 0 d 0 −t c f g = −dt c df dg .
0 0 −d −t b G −t f dt b −dG dt f
It follows that
bd = 0,
cd = 0,
f d = df,
gd = −dg,
Gd = −dG.
Since these equations hold for all diagonal matrices d ∈ gl(`, F ), it follows that
b = 0 and c = 0. Also, by Lemma 10.8.5, f is a diagonal matrix and g = 0 and
G = 0. Since, by assumption, w has zero entries on the main diagonal, we see
that f = 0. Thus, w = 0.
Lemma 10.8.7. Let ` be an integer with ` ≥ 2. Let F have characteristic zero
and be algebraically closed. Let ` be a positive integer. Let H be the subalgebra
of so(2`+1, F ) consisting of diagonal matrices; by Lemma 10.8.6, H is a Cartan
subalgebra of so(2` + 1, F ). Let Φ be the set of roots of so(2` + 1, F ) defined with
respect to H. Let V = R⊗Q V0 , where V0 is the Q subspace of H ∨ = HomF (H, F )
spanned by the elements of Φ; by Proposition 8.2.1, Φ is a root system in V .
For j ∈ {1, . . . , `}, define a linear functional
αj : H −→ F
by
0 0 0
αj (0 h 0 ) = hjj
0 0 −h
for h ∈ gl(`, F ) and h diagonal. The set Φ consists of the following 2`2 linear
functionals on H:
α1 , . . . , αn ,
−α1 , . . . , −αn ,
αi − αj , i, j ∈ {1, . . . , `}, i 6= j,
αi + αj , i, j ∈ {1, . . . , `}, i < j,
−(αi + αj ), i, j ∈ {1, . . . , `}, i < j.
The set
B = {β1 = α1 − α2 , β2 = α2 − α3 , . . . , β`−1 = α`−1 − α` , β` = α` }
is a base for Φ, and the positive roots with respect to B are
α1 , . . . , αn ,
10.8. THE LIE ALGEBRA so(2` + 1) 167
And
0 0 0 0 0 0
[0 h 0 , 0 0 eij − eji ]
0 0 −h 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0
= 0 h 0 0 0 eij − eji − 0 0 eij − eji 0 h 0
0 0 −h 0 0 0 0 0 0 0 0 −h
0 0 0 0 0 0
= 0 0 hii eij − hjj eji − 0 0 −hjj eij + hii eji
0 0 0 0 0 0
0 0 0
= 0 0 hii eij − hjj eji + hjj eij − hii eji
0 0 0
0 0 0
= 0 0 (hii + hjj )eij − (hii + hjj )eji
0 0 0
0 0 0
= (hii + hjj ) · 0 0 eij − eji .
0 0 0
And
0 0 0 0 0 0
[0 h 0 , 0 eij 0 ]
0 0 −h 0 0 −eji
0 0 0 0 0 0 0 0 0 0 0 0
= 0 h 0 0 eij 0 − 0 eij 0 0 h 0
0 0 −h 0 0 −eji 0 0 −eji 0 0 −h
0 0 0 0 0 0
= 0 hii eij 0 − 0 hjj eij 0
0 0 hjj eji 0 0 hii eji
0 0 0
= 0 (hii − hjj )eij 0
0 0 (hii − hjj )(−eji )
0 0 0
= (hii − hjj ) · 0 eij 0 .
0 0 −eji
These calculations show that the linear functionals from the statement of the
lemma are indeed roots, and that the root spaces of these roots are as stated
(recall that any root space is one-dimensional by Proposition 7.0.8). Since the
span of H and the stated root spaces is so(2` + 1, F ) it follows that these roots
are all the roots of so(2` + 1, F ) with respect to H. It is straightforward to
verify that B is a base for Φ, and that the positive roots of Φ with respect to
B are as stated. Note that the dimension of V is ` (by Proposition 7.1.2).
0 −α1 −α2 −α3 α1 α2 β3 = α3
∗ β1 = α1 − α2 α1 − α3
h11 0 α1 + α2 α1 + α3
∗ α2 − α1 h22 β2 = α2 − α3 ∗ 0 α2 + α3
∗ α3 − α1 α3 − α2 h33 ∗ ∗ 0
∗ 0 −(α1 + α2 ) −(α1 + α3 ) −h11 ∗ ∗
∗ ∗ 0 −(α2 + α3 ) ∗ −h22 ∗
∗ ∗ ∗ 0 ∗ ∗ −h33
Finally,
(β` , β` ) = (α` , α` )
1
= .
4` − 2
This completes the proof.
Lemma 10.8.10. Let ` be an integer such that ` ≥ 2. Let F have characteristic
zero and be algebraically closed. Let ` be a positive integer. The Dynkin diagram
of so(2` + 1, F ) is
B` : ... i
10.8. THE LIE ALGEBRA so(2` + 1) 173
(βi , βj )
hβi , βj i = 2
(βj , βj )
−2 if i and j are consecutive and j = `,
= −1 if i and j are consecutive and j 6= `,
0 if i and j are not consecutive.
Hence,
(βi , βj )2
hβi , βj ihβj , βi i = 4
(βi , βi )(βj , βj )
2 if i and j are consecutive and j = ` or i = `,
= 1 if i and j are consecutive and i 6= ` and j 6= `,
0 if i and j are not consecutive.
It follows that the Dynkin diagram of so(2` + 1, F ) is B` , and the Cartan matrix
of so(2` + 1, F ) is as stated. Since B` is connected, so(2` + 1, F ) is simple by
Lemma 9.3.2 and Proposition 10.3.2.
Lemma 10.8.11. Assume that the characteristic of F is zero and F is alge-
braically closed. Let ` be a positive integer. The Killing form
κ : so(2` + 1, F ) × so(2` + 1, F ) −→ F
is given by
κ(x, y) = (2` − 1) · tr(xy).
for x, y ∈ so(2` + 1, F ).
Proof. By Lemma 10.5.3, there exists c ∈ F × such that κ(x, y) = ctr(xy) for
x, y ∈ so(2`+1, F ). Let H be the subalgebra of diagonal matrices in so(2`+1, F );
H is a Cartan subalgebra of so(2`+1, F ) by Lemma 10.8.6. By Lemma 10.8.8 we
have κ(h, h0 ) = (2`−1)·tr(hh0 ) for h, h0 ∈ H. Hence, ctr(hh0 ) = (2`−1)·tr(hh0 )
for h, h0 ∈ H. Since there exist h, h0 ∈ H such that tr(hh0 ) 6= 0 we conclude
that c = 2` − 1.
174 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
Proof. Let
a b
x=
c d
t
with a, b, c, d ∈ gl(`, F ). Then, by definition, x ∈ sp(2`, F ) if and only if xS =
−Sx where
0 1`
S= .
−1` 0
Thus,
x ∈ sp(2`, F )
t
⇐⇒ xS = −Sx
t
a b 0 1` 0 1` a b
⇐⇒ =−
c d −1` 0 −1` 0 c d
t t
a c 0 1` −c −d
⇐⇒ t t =
b d −1` 0 a b
t t
− c a −c −d
⇐⇒ t t = .
− d b a b
This is the first assertion of the lemma. Using this result it is straightforward
to see that dimF sp(2`, F ) = 2`2 + `.
Now
0 0
1` 0
is contained in sp(2`, F ) and
0
0 0 x 0
= 0
1` 0 0 x
for x ∈ M`,1 (F ). It follows that W contains all the vectors of the form
0
.
∗
2α1 , . . . , 2αn ,
αi + αj , i, j ∈ {1, . . . , `}, i < j.
This equation proves that αi − αj is a root and that Lαi −αj is as stated. Next,
let h ∈ gl(`, F ) be a diagonal matrix, and let i, j ∈ {1, . . . , `}. Then
h 0 0 eij + eji h 0 0 eij + eji 0 eij + eji h 0
[ , ]= −
0 −h 0 0 0 −h 0 0 0 0 0 −h
0 hii eij + hjj eji 0 −hjj eij − hii eji
= −
0 0 0 0
0 eij + eji
= (hii + hjj ) .
0 0
This proves that 2αi is a root for i ∈ {1, . . . , `} and that αi + αj is a root for
i, j ∈ {1, . . . , `} with i < j; also the root spaces of these roots are as stated.
Again let h ∈ gl(`, F ) be a diagonal matrix, and let i, j ∈ {1, . . . , `}. Taking
tranposes of the last equation, we obtain:
t
t h 0 0 eij + eji 0 eij + eji
[ , ] = (hii + hjj )
0 −h 0 0 0 0
0 0 h 0 0 0
[ , ] = (hii + hjj )
eij + eji 0 0 −h eij + eji 0
178 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
h 0 0 0 0 0
[ , ] = −(hii + hjj ) .
0 −h eij + eji 0 eij + eji 0
This proves that −2αi is a root for i ∈ {1, . . . , `} and that −(αi + αj ) is a
root for i, j ∈ {1, . . . , `} with i < j; also the root spaces of these roots are as
described.
To see that B is a base for Φ we note first that dimF V = `, and that the
elements of B are evidently linearly independent; it follows that B is a basis for
the F -vector space V . Since B is the disjoint union of P and {−λ : λ ∈ P }, to
prove that B is a base for Φ it will now suffice to prove that every element of P is
a linear combination of elements from B with non-negative integer coefficients.
Let i, j ∈ {1, . . . , `} with i < j. Then
αi − αj = βi+1 + · · · + βj .
Also, we have
2α` = β` ,
2α`−1 = 2(α`−1 − α` ) + 2α` = 2β`−1 + β` ,
2α`−2 = 2(α`−2 − α`−1 ) + 2α`−1 = 2β`−2 + 2β`−1 + β` ,
···
2α1 = 2β1 + · · · 2β`−1 + β` .
h11 β1 = α1 − α2 α1 − α3 2α1 α1 + α2 α1 + α3
α2 − α1 h22 β2 = α2 − α3 ∗ 2α2 α2 + α3
α3 − α1 α3 − α2 h33 ∗ ∗ β3 = 2α3
−2α1 −(α1 + α2 ) −(α1 + α3 ) −h11 ∗ ∗
∗ −2α2 −(α2 + α3 ) ∗ −h22 ∗
∗ ∗ −2α3 ∗ ∗ −h33
Figure 10.3: The decomposition of sp(6, F ). For this example, ` = 3. The positions are
labeled with the corresponding root. Note that the diagonal is our chosen Cartan subalgebra.
The positive roots with respect to our chosen base {β1 , β2 , β3 } are boxed, while the colored
roots form our chosen base. Positions labeled with ∗ are determined by other entries. The
linear functionals α1 , α2 and α3 are defined in Proposition 10.9.4.
10.9. THE LIE ALGEBRA sp(2`) 179
κ : sp(2`, F ) × sp(2`, F ) −→ F
is given by
κ(h, h0 ) = (2` + 2) · tr(hh0 )
for h, h0 ∈ H. Here, H is the subalgebra of diagonal matrices in sp(2`, F ); H is
a Cartan subalgebra of sp(2`, F ) by Lemma 10.9.3.
Finally,
(β` , β` ) = 4(α` , α` )
4
= .
4` + 4
This completes the proof.
Lemma 10.9.7. Let ` be an integer such that ` ≥ 2. Let F have characteristic
zero and be algebraically closed. Let ` be a positive integer. The Dynkin diagram
of sp(2`, F ) is
C` : ... h
(βi , βj )
hβi , βj i = 2
(βj , βj )
−1 if i, j ∈ {1, . . . , ` − 1} and i and j are consecutive,
−1 if
i = ` − 1 and j = `,
=
−2 if i = ` and j = ` − 1,
0
if none of the above conditions hold.
Hence,
(βi , βj )2
hβi , βj ihβj , βi i = 4
(βi , βi )(βj , βj )
1 if i, j ∈ {1, . . . , ` − 1} and i and j are consecutive,
= 2 if i = ` − 1 and j = `,
0 if none of the above conditions hold.
It follows that the Dynkin diagram of sp(2`, F ) is C` , and the Cartan matrix
of sp(2`, F ) is as stated. Since C` is connected, sp(2`, F ) is simple by Lemma
9.3.2 and Proposition 10.3.2.
Hence
x ∈ so(2`, F )
t
⇐⇒ xS = −Sx,
t t
a c 0 1` 0 1` a b
⇐⇒ t t =−
b d 1` 0 1` 0 c d
t t
c a −c −d
⇐⇒ t t = .
d b −a −b
This non-zero. An argument as above shows that W contains all the vectors of
the form
0
.
∗
We conclude that, in the current case, W = V . If x = 0 and y 6= 0, then a
similar argument shows that W = V . Assume that x 6= 0 and y 6= 0. By Lemma
184 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
t
10.8.3 there exists b ∈ gl(`, F ) such that − b = b and by 6= 0. Since by 6= 0 and
x 6= 0, there exists a ∈ GL(`, F ) such that ax = −by. Now
a b
t
0 − a
and
d 0 a b da db
hw = t = t .
0 −d c − a −dc d a
It follows that
ad = da,
bd = −db,
cd = −dc,
t t
ad = d a.
Lemma 10.8.5 implies that b = c = 0 and that a is diagonal. Since a has zeros
on the main diagonal by assumption, we also get a = 0. Hence, w = 0.
10.10. THE LIE ALGEBRA so(2`) 185
αi : H −→ F
by
h 0
αi ( ) = hii
0 −h
for h ∈ gl(`, F ) and h diagonal. The set Φ consists of the following 2`2 − 2`
linear functionals on H:
αi − αj , i, j ∈ {1, . . . , `}, i 6= j,
αi + αj , i, j ∈ {1, . . . , `}, i < j,
−(αi + αj ), i, j ∈ {1, . . . , `}, i < j.
The set
is a base for Φ, and the positive roots with respect to B are the set P , where P
consists of the following roots:
This proves that that αi +αj is a root for i, j ∈ {1, . . . , `} with i < j; also the root
spaces of these roots are as stated. Again let h ∈ gl(`, F ) be a diagonal matrix,
and let i, j ∈ {1, . . . , `} with i < j. Taking tranposes of the last equation, we
obtain:
t
t h 0 0 eij − eji 0 eij − eji
[ , ] = (hii + hjj )
0 −h 0 0 0 0
0 0 h 0 0 0
[ , ] = (hii + hjj )
eji − eij 0 0 −h eji − eij 0
h 0 0 0 0 0
[ , ] = −(hii + hjj ) .
0 −h eij − eji 0 eij − eji 0
This proves that that −(αi + αj ) is a root for i, j ∈ {1, . . . , `} with i < j; also
the root spaces of these roots are as described.
To see that B is a base for Φ we note first that dimF V = `, and that the
elements of B are evidently linearly independent; it follows that B is a basis for
the F -vector space V . Since B is the disjoint union of P and {−λ : λ ∈ P }, to
prove that B is a base for Φ it will now suffice to prove that every element of P is
a linear combination of elements from B with non-negative integer coefficients.
Let i, j ∈ {1, . . . , `} with i < j. Then
j−1
X
αi − αj = (αk − αk+1 )
k=i
j−1
X
= βk .
k=i
Also, we have
h11 β1 = α1 − α2 α1 − α3 0 α1 + α2 α1 + α3
α −α h22 β2 = α2 − α3 ∗ 0 β3 = α2 + α3
2 1
α3 − α1 α3 − α2 h33 ∗ ∗ 0
0 −(α1 + α2 ) −(α1 + α3 ) −h11 ∗ ∗
∗ 0 −(α2 + α3 ) ∗ −h22 ∗
∗ ∗ 0 ∗ ∗ −h33
Figure 10.4: The decomposition of so(6, F ). For this example, ` = 3. The positions are
labeled with the corresponding root. Note that the diagonal is our chosen Cartan subalgebra.
The positive roots with respect to our chosen base {β1 , β2 , β3 } are boxed, while the colored
roots form our chosen base. Positions labeled with ∗ are determined by other entries. The
linear functionals α1 , α2 and α3 are defined in Proposition 10.10.4.
= hii
h 0
= αi ( ).
0 −h
(βi , β` ) = (β1 , β2 )
= (α1 − α2 , α1 + α2 )
= (α1 , α1 ) + (α1 , α2 ) − (α2 , α1 ) − (α2 , α2 )
= 0.
190 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
Finally,
(β` , β` ) = (α`−1 + α` , α`−1 + α` )
= (α`−1 , α`−1 ) + (α`−1 , α` ) + (α` , α`−1 ) + (α` , α` )
2
= .
4` − 4
This completes the proof.
Lemma 10.10.7. Let ` be an integer such that ` ≥ 3. Let F have characteristic
zero and be algebraically closed. The Dynkin diagram of so(2`, F ) is
D` : ...
Hence,
(βi , βj )2
hβi , βj ihβj , βi i = 4
(βi , βi )(βj , βj )
1 if i, j ∈ {1, . . . , ` − 1} and i and j are consecutive,
= 1 if {i, j} = {` − 2, `},
0 if none of the above conditions hold.
10.10. THE LIE ALGEBRA so(2`) 191
It follows that the Dynkin diagram of sp(2`, F ) is C` , and the Cartan matrix
of sp(2`, F ) is as stated. Since C` is connected, sp(2`, F ) is simple by Lemma
9.3.2 and Proposition 10.3.2.
192 CHAPTER 10. THE CLASSICAL LIE ALGEBRAS
Chapter 11
Representation theory
Lemma 11.1.1. Let F be algebraically closed and have characteristic zero. Let
L be a finite-dimensional, semi-simple Lie algebra over F . Let H be a Cartan
subalgebra of L, let
M
L=H⊕ Lα
α∈Φ
193
194 CHAPTER 11. REPRESENTATION THEORY
4. If V is finite-dimensional, then V 0 = V .
Proof. Proof of 1. Let λ : H → F be a linear functional, and let α ∈ Φ. Let
x ∈ Lα and v ∈ Vλ . We have
v 1 = v2 + · · · + vt
where vi ∈ Vλi for i ∈ {2, . . . , t}. The minimality of t implies that vi is non-zero
for i ∈ {2, . . . , t}. Let h ∈ H. Then
φ(h)v1 = φ(h)(v2 + · · · + vt )
λ1 (h)v1 = λ2 (h)v2 + · · · + λt (h)vt ,
Subtracting, we obtain:
[5]) that there exists a basis v1 , . . . , vn for V such that each vi for i ∈ {1, . . . , n}
is an eigenvector for every operator φ(h) for h ∈ H. Let i ∈ {1, . . . , n}. For
h ∈ H, let λ(h) ∈ F be such that φ(h)vi = λ(h)vi . Since the map H → gl(V )
given by h 7→ φ(h) is linear, and vi is non-zero, the function λ : H → F is also
linear. It follows that λ is a weight of H on V and that vi ∈ Vλ . We conclude
that V ⊂ V 0 .
be the root space decomposition of L with respect to L from Chapter 7, and let
B be a base for Φ. Let Φ+ be the positive roots in Φ with respect to B. Define
X
N= Lα
α∈Φ+
and X
P =H +N =H + Lα .
α∈Φ+
[P, P ] = N,
···
X
k+1
N = [N, N k ] ⊂ Lα1 +···+αk .
α1 ,...,αk ∈Φ+
Sk = {α1 + · · · + αk : α1 , . . . , αk ∈ Φ+ }.
Evidently, the sets Sk for k a positive integer do not contain the zero linear
functional. Recall the height function from page 93. Let m = max({ht(β) : β ∈
Φ+ }). Since ht(λ) ≥ k for all λ ∈ Sk , the set Sk for k ≥ m + 1 cannot contain
any elements of Φ+ . Also, it is clear that Sk does not contain any elements of
the set Φ− of negative roots (by the basic properties of the base B). Thus, if
k ≥ m + 1, then Lλ = 0 for all λ ∈ Sk . It follows that N m+2 = 0 so that N is
nilpotent.
Finally, P is solvable because [P, P ] = N and N is nilpotent.
Lemma 11.3.1. Let F be algebraically closed and have characteristic zero. Let
L be a finite-dimensional, semi-simple Lie algebra over F . Let H be a Cartan
subalgebra of L, let Φ be the roots of L with respect to H, and let B be a base for
Φ. Define N and the Borel subalgebra P as as in Lemma 11.2.1. Let (φ, V ) be
a representation of L. If V is finite-dimensional, then V has a maximal vector
of weight λ for some weight λ of H on V .
11.3. MAXIMAL VECTORS 197
Proof. Let P be the Borel subalgebra of L defined with respect to our chosen
base. By Lemma 11.2.1, P is solvable. Consider φ(P ) ⊂ gl(V ). Since φ is
a map of Lie algebras, φ(P ) is a Lie subalgebra of gl(V ). By Lemma 2.1.5,
φ(P ) is solvable. By Lemma 3.4.1, a version of Lie’s Theorem, there exists a
non-zero vector v of V such that v is a common eigenvector for the operators
φ(x) ∈ gl(V ), x ∈ P . For x ∈ P , let c(x) ∈ F be such that φ(x)v = c(x)v. It is
easy to see that the function c : P → F is F -linear. We claim that c(N ) = 0.
Let x, y ∈ P . Then
µ = λ − (c1 α1 + · · · + cn αn )
w = φ(yβ1 ) · · · φ(yβk )v
Since this equality holds for all h ∈ H, and the sum of Vµ1 , . . . , Vµt is direct, we
must have ν = µ1 = · · · = µt . Since µ1 , . . . , µt are mutually distinct, we obtain
t = 1 and ν = µ1 . Recalling the definition of the set M , and the fact that every
element of Φ− can be uniquely written as a linear combination of the elements
of B = {α1 , . . . , αn } with non-positive integral coefficients, we see that ν has
the form as stated in the theorem.
Finally, let µ be a weight of H on V . Let u ∈ Vµ be non-zero. By the
first paragraph, w can be written as linear combination of v and elements of
the form w = φ(yβ1 ) · · · φ(yβk )v. Hence, there exists a positive integer `, ele-
ments c0 , c1 , . . . , c` of F , and for each i ∈ {1, . . . , `} a positive integer ki and
βi,1 , . . . , βi,ki ∈ Φ− such that
`
X
u = c0 v + ci φ(yβi,1 ) · · · φ(yβi,ki )v.
i=1
Since φ(yβi,1 ) · · · φ(yβi,ki )v is contained in Vλ+βi,1 +···+βi,ki , and since the sum of
weight spaces is direct by 3 of Lemma 11.1.1, we see that for each i ∈ {1, . . . , `},
if
ci φ(yβi,1 ) · · · φ(yβi,ki )v
is non-zero, then
µ = λ + βi,1 + · · · + βi,ki ,
or equivalently,
µ − λ = βi,1 + · · · + βi,ki .
11.3. MAXIMAL VECTORS 199
w = wµ1 + · · · + wµk ,
i ∈ {1, . . . , k}, then we will say that w has property P. Suppose that there
exists a non-zero w ∈ W which has property P; we will obtain a contradiction.
We may assume that k is minimal. Since k is minimal, we must have k > 1:
otherwise, w = wµ1 ∈ W ∩ Vµ1 = Wµ1 , a contradiction. Also, we claim that
wµi ∈/ W for i ∈ {1, . . . , k}. To see this, let X = {i ∈ {1, . . . , k} : wµi ∈ W },
and assume that X is non-empty. Since w has property P , the set X is a proper
subset of {1, . . . , k}. We have
X X
w− wµi = wµj .
i∈X j∈{1,...,k}−X
Also, we have
Subtracting yields:
φ(h)w − µ2 (h)w
= (µ1 (h) − µ2 (h))wµ1 + (µ3 (h) − µ2 (h))wµ3 + · · · + (µk (h) − µ2 (h))wµk .
w1 = w1,µ1 + · · · + w1,µk ,
w2 = w2,ν1 + · · · + w2,ν`
T = T0 ⊕ T1 ⊕ T2 ⊕ ···
202 CHAPTER 11. REPRESENTATION THEORY
Tk = L ⊗ ··· ⊗ L.
| {z }
k
x ⊗ y − y ⊗ x − [x, y]
for x, y ∈ L. We define
U (L) = T /J,
and refer to U (L) as the universal enveloping algebra of L. We let
p
T −→ T /L = U (L)
Let
T+ = T 1 ⊕ T 2 ⊕ T 3 ⊕ · · · .
Then
T = T 0 ⊕ T+ = F · 1 ⊕ T+ .
Evidently, T+ is a two-sided ideal of T . Since x ⊗ y − y ⊗ x − [x, y] ∈ T+ for
x, y ∈ L, it follows that
J ⊂ T+ .
We claim that
p(T 0 ) ∩ p(T+ ) = 0.
To see this, let a ∈ F and z ∈ T+ be such that p(a·1) = p(z). Then p(a·1−z) = 0.
This means that a · 1 − z ∈ J. Since J ⊂ T+ , we get a · 1 − z ∈ T+ , and therefore
a · 1 ∈ T+ . As T 0 ∩ T+ = 0, this yields a = 0, as desired. Letting
U+ = p(T+ ),
U (L) = F · 1 ⊕ U+ .
where the first map is the inclusion map, and the second map is the projection
map p. We refer to σ as the canonical map of L into U (L). Let x, y ∈ L.
Then
That is,
σ(x)σ(y) − σ(y)σ(x) = σ([x, y])
for x, y ∈ L.
Lemma 11.4.1. Let F be a field, and let L be a Lie algebra over F . Let
σ : L → U (L) be the canonical map. Let A be an associative algebra with
identity, and assume that
τ
L −→ A
is a linear map such that
L
τ
σy &
τ0
U (L) −→ A
commutes.
Proof. To prove the existence of τ 0 , we note first that by the universal property
of T , there exists an algebra homomorphism
ϕ
T −→ A
= 0.
(τ 0 ◦ σ)(x) = τ 0 (σ(x))
= τ 0 (x + J)
= ϕ(x)
= τ (x).
This proves the existence of τ 0 . The uniqueness of τ 0 follows from the fact that
U (L) is generated by 1 and σ(L), the assumption that τ 0 is determined on these
elements.
We will consider sequences (i1 , . . . , ip ) where p is as positive integer, i1 , . . . , ip
are positive integers, and
i1 ≤ · · · ≤ ip .
We let X be the set consisting of all such sequences, along with the empty set ∅.
Let I ∈ X. If I 6= ∅, so that there exists a positive integer p, and i1 , . . . , ip ∈ Z>0
such that I = (i1 , . . . , ip ) with i1 ≤ · · · ≤ ip , then we define
d(I) = p.
If I = ∅, then we define
d(∅) = 0.
Let F be a field, and let L be a Lie algebra over F . Assume that L is
finite-dimensional and non-zero. We fix an ordered basis
x1 , x2 , x3 , . . . , xn
for L as a vector space over F . We define the images of these vectors in U (L)
as
y1 = σ(x1 ), y2 = σ(x2 ), y3 = σ(x3 ), . . . , yn = σ(xn ).
Let I ∈ X. If I 6= ∅, so that there exists a positive integer p, and i1 , . . . , ip ∈ Z>0
such that I = (i1 , . . . , ip ) with i1 ≤ · · · ≤ ip , then we define
If I = ∅, then we define
y∅ = 1 ∈ U (L).
Lemma 11.4.2. Let F be a field, and let L be a finite-dimensional Lie algebra
over F . Fix an ordered basis x1 , . . . , xn for L, and define yI for I ∈ X as above.
Then the elements yI for I ∈ X span U (L) as a vector space over F .
11.4. THE POINCARÉ-BIRKHOFF-WITT THEOREM 205
P = F [z1 , . . . , zn ].
If I = ∅, then we define
z∅ = 1 ∈ F [z1 , . . . , zn ].
Evidently, the elements zI for I ∈ X form a basis for F [z1 , . . . , zn ]. For conve-
nience, we define
P = F [z1 , . . . , zn ].
Also, if k is a non-negative integer, then we let Pk be the F -subspace of P of
polynomials of degree less than or equal to k. Evidently, if k is a non-negative
integer, then Pk has as basis the elements zI for I ∈ Xk .
Let I ∈ X. Let i ∈ {1, . . . , n}. We say that i ≤ I if and only if I = ∅, or,
if I 6= ∅, so that I = (i1 , . . . , ip ) for some positive integers p and i1 , . . . , ip with
i1 ≤ · · · ≤ ip , then i ≤ i1 .
fp (xi ⊗ zI ) = zi zI .
Proof. We will prove by induction that the following statement holds for all
non-negative integers p: (Sp ) there exists a unique linear map fp : L ⊗ Pp → P
satisfying (Ap ), (Bp ) and (Cp ) and such that the restriction of fp to L ⊗ Pp−1
is fp−1 when p is positive.
Suppose that p = 0. Clearly, there exists a unique linear map f0 : L⊗P0 → P
such that f0 (xi ⊗ 1) = zi for i ∈ {1, . . . , n}. It is clear that (A0 ), (B0 ) and (C0 )
hold; for this, note that, by definition, X0 = {I ∈ X : d(I) = 0} = {∅}, z∅ = 1,
and i ≤ ∅ for all i ∈ {1, . . . , n}. It follows that (S0 ) holds.
Suppose that p = 1. To define the linear map fp : L ⊗ P1 → P it suffices to
define f1 (xi ⊗ zI ) ∈ P for i ∈ {1, . . . , n} and I ∈ X0 t X1 . If i ∈ {1, . . . , n} and
I ∈ X0 , then I = ∅, and we define f1 (xi ⊗ zI ) = f0 (xi ⊗ zI ) = zi . Assume that
i ∈ {1, . . . , n} and I ∈ X1 . Write I = (j). There are two cases. Assume first
that i ≤ I, i.e., i ≤ j. In this case we define f1 (xi ⊗ zI ) = zi zj . Assume that
i 6≤ I, so that i > j. We define:
f1 (xi ⊗ zI ) = zi zI + f0 ([xi , xj ] ⊗ z∅ ),
i.e.,
f1 (xi ⊗ zj ) = zi zj + f0 ([xi , xj ] ⊗ 1).
It is straightfoward to verify that f1 satisfies (A1 ) and (B1 ). To see that f1
satisfies (C1 ), let i, j ∈ {1, . . . , n} and J ∈ X1−1 = X0 . Then J = ∅. We need
to prove that
which is
f1 (xi ⊗ zj ) = f1 (xj ⊗ zi ) + f0 ([xi , xj ] ⊗ 1).
Assume first that i ≤ j. In this case,
f1 (xi ⊗ zj ) = zi zj ,
and
This proves (C1 ) in the case i ≤ j. Now assume that i > j. Then
and
This proves (C1 ) in the remaining case i > j. It follows that (S1 ) holds.
11.4. THE POINCARÉ-BIRKHOFF-WITT THEOREM 207
Now suppose that p is a positive integer such that p ≥ 2 and that (Sk ) holds
for k = 0, . . . , p − 1. To define the linear map fp : L ⊗ Pp → P it suffices to
define fp (xi ⊗ zI ) ∈ P for i ∈ {1, . . . , n} and I ∈ Xq with q such that 0 ≤ q ≤ p.
Let i ∈ {1, . . . , n} and assume that I ∈ Xq with 0 ≤ q < p. In this case we
define fp (xi ⊗ zI ) = fp−1 (xi ⊗ zI ). Assume that i ∈ {1, . . . , n} and I ∈ Xp . If
i ≤ I, then we define
fp (xi ⊗ zI ) = zi zI .
Assume that i 6≤ I. To see how to define fp (xi ⊗ zI ) in this case, assume for
the moment that fk exists and satisfies (Ak ), (Bk ), and (Ck ) for non-negative
integers k and that fk−1 is the restriction of fk for k = 1, . . . , p; we will find a
formula for fp (xi ⊗ zI ) in terms of fp−1 . Let I = (j, i2 , . . . , ip ). By the definition
of X, j ≤ i2 ≤ · · · ≤ ip . Since i 6≤ I, we must have i > j. Define J = (i2 , . . . , ip );
note that the definition of J is meaningful since p ≥ 2. Clearly, J ∈ X with
d(J) = p − 1. We calculate, using (Ap−1 ) and then (Cp ):
fp (xi ⊗ zI ) = fp (xi ⊗ zj zJ )
= fp (xi ⊗ fp−1 (xj ⊗ zJ ))
= fp (xi ⊗ fp (xj ⊗ zJ ))
= fp (xj ⊗ fp (xi ⊗ zJ )) + fp ([xi , xj ] ⊗ zJ )
= fp (xj ⊗ fp−1 (xi ⊗ zJ )) + fp ([xi , xj ] ⊗ zJ ).
Define
w(i, J) = fp−1 (xi ⊗ zJ ) − zi zJ .
As just indicated, we have that w(i, J) ∈ Pp−1 . Continuing the calculation, we
get:
where I = (j, j1 , . . . , jp−1 ), J = (j1 , . . . , jp−1 ) and we have used the definition
of fp (xi ⊗ zI ) from the last paragraph. We note that j ≤ I. Also,
xj · zJ = xj · (zk zK )
= xj · (xk · zK )
= xk · (xj · zK ) + [xj , xk ] · zK
where the last equality follows from (Cp−1 ). Now xj · zK = zj zK + w for some
w ∈ Pp−2 . Therefore,
xj · zJ = xk · (zj zK ) + xk · w + [xj , xk ] · zK .
Applying xi , we get:
ρ0 (yI ) · 1 = zI .
ρ0 (yI ) · 1 = ρ0 (yi1 yJ ) · 1
= ρ0 (yi1 )(ρ0 (yJ ) · 1)
= ρ(xi1 )(zJ )
= zi1 zJ
= zI .
This proves the claim by induction. It follows now that the yI for I ∈ X are
linearly independent because the zI are linearly independent for I ∈ X.
Index
211
212 INDEX
root system, 83
Schur’s lemma, 24
semi-simple component, 49
semi-simple Lie algebra, 10
short roots, 120
simple ideal, 44
simple Lie algebra, 43
solvable, 8
subalgebra, 2
subrepresentation, 23
[1] Nicolas Bourbaki. Lie groups and lie algebras, chapter 1–3, 1989.
[2] Nicolas Bourbaki. Lie groups and lie algebras. chapters 4–6. translated from
the 1968 french original by andrew pressley. elements of mathematics, 2002.
[3] Nicolas Bourbaki. Lie groups and lie algebras. chapters 7–9. translated from
the 1975 french original by andrew pressley. elements of mathematics, 2008.
[4] Karin Erdmann and Mark J. Wildon. Introduction to Lie Algebras. Springer
Undergraduate Mathematics Series. Springer, 2006.
[5] Kenneth Hoffman and Ray Kunze. Linear Algebra. Prentice-Hall, second
edition, 1971.
[6] James E Humphreys. Introduction to Lie algebras and representation theory,
volume 9. Springer Science & Business Media, 2012.
213