In this chapter we shall get at the heart of the structure theory of finite-dimensional Lie algebras of characteristic zero. We shall obtain the structure of semi-simple algebras of this type, prove complete reducibility of finite-dimensional representations and prove the Levi radical splitting theorem. All these results are consequences of certain trace criteria for solvability and semi-simplicity. One of these is Cartan’s criterion that a finite-dimensional Lie algebra of characteristic 0 is solvable if and only if tr (ad a)2 = 0 for every a in ′ Clearly, this is a weakening of Engel’s condition— that ad a is nilpotent for every a ∈
′. The method which we employ to establish this result and others of the same type is classical and is based on the study of certain nilpotent subalgebras, called Cartan subalgebras. We shall pursue this method further in the next chapter to obtain the classification of simple Lie algebras over algebraically closed fields of characteristic 0.
If is a subalgebra of a Lie algebra
then the normalizer
of
is the set of x ∈
such that
, that is, [xb] ∈
for every b ∈
. It is immediate that
is a subalgebra containing
, and
is an ideal in
. In fact, as in group theory,
is the largest subalgebra in which
is contained as an ideal. We now give the following
DEFINITION 1. A subalgebra of a Lie algebra
is called a Cartan subalgebra if (1)
is nilpotent and (2)
is its own normalizer in
Let be a nilpotent subalgebra of a finite-dimensional Lie algebra
and let
=
0 ⊕
1 be the Fitting decomposition of
relative to ad
. We recall that
0 = {x | x(ad h)k = 0, h ∈
, for some integer k}. We can now establish the following criterion.
PROPOSITION 1. Let be a nilpotent subalgebra of a finite-dimenstona Lie algebra
. Then
is a Cart an subalgebra if and only if
coincides with the Fitting component
0 of
relative to ad
.
Proof: We note first that , the normalizer of
. Thus if x ∈
then [xh] ∈
for any h ∈
. Since
is nilpotent,
for some k. Hence x ∈ 0. Hence if
⊃
, then
0 ⊃
. Next assume
0 ⊃
. Now
0 is invariant under ad
and every restriction ad
, is nilpotent. Also
is an invariant subspace of
0 relative to ad
. Hence we obtain an induced Lie algebra
of linear transformations acting in the non-zero space
. Since these transformations are nilpotent, one of the versions of Engel’s theorem implies that there exists a non-zero vector x +
such that
. This means that we have [xh] ∈
for every h; hence x ∈
and
so that
. Thus
if and only if
, which is what we wished to prove
PROPOSITION 2. Let be a nilpotent subalgebra of the finite dimensional Lie algebra
and let
be the Fitting decomposition of
relative to ad
. Then
0 is a subalgebra and
1.
Proof: Let h ∈ and a ∈
0. Then
for some k. Hence
This relation and Lemma 1 of §2.4 implies that the Fitting spaces 0 ad h and
1 ad h of
relative to ad h are invariant under ad a. Since
and
, it follows that
0 ad
and
1 ad
. Since a is any element of
and
.
DEFINITION 2. An element h ∈ is called regular if the dimensionality of the Fitting null component of
relative to ad h is minimal. If this dimensionality is l, then n – l, where n = dim
, is called the rank of
.
We have seen that the dimensionality of the Fitting null component of a linear transformation A is the multiplicity of the root 0 of the characteristic polynomial of A. Hence h is regular if and only if the multiplicity of the characteristic root 0 of ad h is minimal. Since [hh] = 0 for every h it is clear that ad h is singular for every h. Hence the number l of the above definition is > 0. We remark also that
is of rank 0 if and only if every ad h is nilpotent. By Engel’s theorem this holds if and only if
is a nilpotent Lie algebra. Regular elements can be used to construct Cartan subalgebras, for we have the following
THEOREM 1. If is a finite-dimensional Lie algebra over an infinite field Φ and a is a regular element of
then the Fitting null component
of
relative to ad a is a Cartan subalgebra.
Proof. Let =
⊕
be the Fitting decomposition of
relative to ad a. Then, by Proposition 2,
is a subalgebra and
. We assert that every ad
b, b ∈
, is nilpotent. Otherwise let b be an element of
such that ad
b is not nilpotent. We choose a basis for
which consists of a basis for
and a basis for
. Then the matrix of any ad h, h ∈
, relative to this basis has the form
where (ρ1) is a matrix of adh and (ρ2) is a matrix of ad
h. Let
respectively be the matrices for ad a and ad b. Then we know that (α2) is non-singular; hence det(α2) ≠ 0. Also, by assumption, (β1) is not nilpotent. Hence if n – l is the rank then dim = l and the characteristic polynomial of (β1) is not divisible by λl. Now let λ, μ v be (algebraically independent) indeterminates and let F(λ, μ, v) be the characteristic polynomial F(λ, μ, v) = det(λ1 – μA – vB). We have F(λ, μ, v) = F1(λ, μ, v)F2(λ, μ, v) where
We have seen that F2(λ, 1, 0) = det(λ1 — (α2)) is not divisible by λ and F1(λ, 0, 1) = det(λ1 – (β1)) is not divisible by λl. Hence the highest power of λ dividing F(λ, μ, v) is λl′, l′ < l. Since Φ is infinite we can choose μ0, v0 in Φ such that F(λ, μ0, v0) is not divisible by λl′+1. Set c = μ0a + v0b. Then the characteristic polynomial det (λ1 – ad c) = det(λ1 – μ0A – v0B) = F(λ, μ0, v0) is not divisible by λl′+1. Hence the multiplicity of the characteristic root 0 of ad c is l′ < l. This contradicts the regularity of a. We have therefore proved that for every b ∈ , ad
b is nilpotent. Consequently, by Engel’s theorem,
is a nilpotent Lie algebra. Let
0 be the Fitting null component of
relative to ad
. Then
since the latter is the Fitting null component of ad a and a ∈
. On the other hand, we always have that
for a nilpotent subalgebra. Hence
0 =
, and
is a Cartan subalgebra by Proposition 1.
Another useful remark about regular elements and Cartan subalgebras is that if a Cartan subalgebra contains a regular element a, then
is uniquely determined by a as the Fitting null component of ad a. Thus if
is this component then it is clear that
since
is nilpotent. On the other hand, we have just seen that
is nilpotent so that if
, then
contains an element
such that
(cf. Exercise 1). This contradicts the assumption that
is a Cartan subalgebra. An immediate consequence of our result is that if two Cartan subalgebras have a regular element in common then they coincide. We shall see later (Chapter IX) that if Φ is algebraically closed of characteristic zero, then every Cartan subalgebra contains a regular element. We now indicate a fairly concrete way of determining the regular elements assuming again that Φ is infinite. For this purpose we need to introduce the notion of a generic element and the characteristic polynomial of a Lie algebra.
Let be a Lie algebra with basis (e1, e2, …, en) over the field Φ. Let
be indeterminates and let
, the field of rational expressions in the ξi. We form the extension
. The element
of
P is called a generic element of
and the characteristic polynomial fx(λ) of ad x (in
P) is called the characteristic polynomial of the Lie algebra
. If we use the basis (e1, e2, …, en) for
P, then we can write
where the pij are homogeneous expressions of degree one in the ξk. It follows that
where τi is a homogeneous polynomial of degree i in the ξ’s and but τn–1+k = 0, if k > 0. Since x ad x = 0 and x ≠ 0, det(ρ) = 0 and l > 0. The characteristic polynomial of any a = Σαiei ∈
is obtained by specializing ξi = αi, i = 1, 2, …, n, in (4). Hence it is clear that the multiplicity of the root 0 for the characteristic polynomial of ad a is at least l. On the other hand, if Φ is an infinite field then, since the polynomial
in the polynomial algebra
, we can choose
so that
. Then ad a for a = Σαiei has exactly l characteristic roots 0, and so a is regular. Thus we see that for an infinite field, a is regular if and only if
In this sense “almost all” the elements of are regular. (In the sense of algebraic geometry the regular elements form an open set.) It is also clear that n – l is the rank of
.
All of this depends on the choice of the basis (e). However, it is easy to see what happens if we change to another basis (f1, f2 …, fn) where fi = Σμijej. Thus if η1, η2, …, ηn are indeterminates, then . Hence the characteristic polynomial fy(λ) is obtained from fx(λ) by the substitutions
in its coefficients (of the powers of λ).
If Ω is any extension field of Φ, then (e) is a basis for Ω over Ω. Hence
can be considered also as a generic element of
Ω and the characteristic polynomial fx(λ) is unchanged on extending the base field Φ to Ω. It is clear from this also that if Φ is infinite, a ∈
is regular in
if and only if a is regular as an element in
Ω. (In either case, (5) is the condition for regularity.) We have seen that the Fitting null component
of ad a, a regular is a Cartan subalgebra. The dimensionality of
is l, which is the multiplicity of the characteristic root 0. It follows that the Cartan subalgebra determined by a in
Ω is
Ω.
It is convenient to carry over the notion of weights and weight spaces for a Lie algebra of linear transformations to an abstract Lie algebra and a representation R of
. Let
be the module for
. A mapping a → α(a) of
into Φ is called a weight of
if there exists a non-zero vector x in
such that
for a suitable k. The set of vectors satisfying this condition together with 0 is a subspace α called the weight space corresponding to the weight α. If
is nilpotent, then Lemma 2.1 shows that
α is a submodule. If
=
α, then we shall say that
is a weight module for
corresponding to the weight α.
Let be a Lie algebra and let
be a finite-dimensional weight module for
relative to the weight α. Then for any x ∈
,
if k is sufficiently large. Moreover, if dim
= n, then (λ – α(a))n is the characteristic polynomial of aR. Hence we have x(aR – α(a)1)n = 0 for all x ∈
. We consider the contragredient module
* which carries the representation R* satisfying
x ∈ , y* ∈
*. We have
which we can add to (7) to obtain
Iteration of this gives
If k = n, x(aR – α(a)1)n = 0 for all x and consequently, by (10), x, y*(aR* + α(a)1)n
= 0. Hence y*(aR* + α(a)1)n = 0 for all y* ∈
*. This shows that
* is a weight module With the weight – α.
PROPOSITION 3. If is a finite-dimensional weight module for
with the weight α, then the contragredient module
* is a weight module with the weight –α.
We consider next what happens if we take the tensor product of two weight spaces. Thus let ,
be weight modules of
relative to the weights α and β. Let R and S denote the representations in
and
, respectively. Then any x ∈
satisfies
for some positive integer k, and every y ∈ satisfies
for some positive integer k′. Let =
⊗
and denote the representation of
in
by T. Then we have
or aT = aR ⊗ 1 + 1 ⊗ aS. Hence
Since the two transformations in the parentheses commute we can apply the binomial theorem to obtain
If we apply this to x ⊗ y, we obtain
If we take m = k + k′ – 1, then for every i, either x(aR – α(a)1)i = 0 or y(as – β(a)1)m–i = 0. Hence (x ⊗ y)(aT – (α(a) + β(a))1)m = 0 and we have the following
PROPOSITION 4. If and
are weight modules for
for the weights α and β, respectively, then
=
⊗
is a weight module with the weight α + β.
We suppose now that is a finite-dimensional Lie algebra,
a nilpotent subalgebra, and
is a finite-dimensional module for
, hence for
. If R denotes the representation in
and ad
the adjoint representation in
, then we assume that
R and ad
are split Lie algebras of linear transformations, that is, the characteristic roots of hR and ad
h, h ∈
, are in the base field Φ. This will be automatically satisfied if Φ is algebraically closed. If hR → α(hR) is a weight on
R, then h → α(hR) is a weight for
in the module
. The result on weight spaces for a split nilpotent Lie algebra of linear transformations (Theorem 2.7) implies that
is a direct sum of weight modules
ρ. Similarly, we have a decomposition of
into weight modules
α. Thus we have
and
where ρ, σ, …, are mappings of into Φ such that if xP ∈
ρ then xρ(hR – ρ(h)1)m = 0 for some m, etc. The weights α, β, …, associated with ad
will be called the roots of
in
. Since h → hR and h → ad
h are linear, it is clear that in the characteristic 0 case the weights ρ, …, τ and the roots α, …, δ are linear functions on
which vanish on
′.
The following important result relates the decompositions (17) and (18) relative to .
PROPOSITION 5. if ρ + α is a weight of
relative to
; otherwise
ρ
α = 0.
Proof. The elements of ρ
α are of the form
,
. The characteristic property of the tensor product of two spaces shows that we have a linear mapping π of
ρ ⊗
α onto
ρ
α such that
We shall now show that π is in fact a homomorphism for the -modules. Thus let h ∈
. Then we have
On the other hand, the image of xρ ⊗ αα under π is xρaα. Following this with the module product by h we again obtain (xρaα)h. We have therefore proved that ρ
α is a homomorphic image of
P ⊗
α. Moreover, the latter is a weight module for the weight ρ + α. Now it is clear from the definition that the homomorphic image of a weight module with weight β is either 0 or it is a weight module with weight β. The result stated follows from this.
If we apply the last result to the case in which =
and the representation is the adjoint representation of
, we obtain the
COROLLARY. is a root, and [
α
β] = 0 otherwise.
Before plunging into the structure theory it will be well to look at an example. Let be an n-dimensional vector space over an algebraically closed field Φ,
the associative algebra of linear transformations in
,
=
L, the corresponding Lie algebra. We wish to determine the regular elements of
and the corresponding Cartan subalgebras.
Let A ∈ and let
be the decomposition of
into the weight spaces relative to Φ A. The αi are distinct elements of Φ and these are just the characteristic roots of the linear transformation A and the
are the corresponding characteristic spaces. It is well known that a decomposition of a space as a direct sum leads to a decomposition of the dual space into a direct sum of subspaces which can be considered as the conjugate spaces of the components. Hence we have
where can be identified with the conjugate space of
, is invariant under –A* and Proposition 3 shows that this is a weight module for the Lie algebra ΦA with the weight – αi. Accordingly, we write
for
and we have
We now consider the module ⊗
* relative to ΦA. As is well known, we have
. By Proposition 4,
is a weight module for ΦA for the weight αi – αj. If Aij denotes the linear transformation in
corresponding to A, then (Aij – (αi – αj)l)kij = 0 for suitable kij. Hence Aij is non-singular if
and Aij is nilpotent if αi = αj. This implies that the Fitting null component of
⊗
* relative to the linear transformation
corresponding to A (A →
is the representation of
in
⊗
*) is
. If dim
, then the dimensionality of this Fitting space is
. Since
,
is minimal if and only if every ni = 1, which is equivalent to saying that A has n distinct characteristic roots. Now, we recall that the module
⊗
* is isomorphic to the module
=
L relative to the adjoint mapping X → [XA]. Hence we see that the dimensionality of the Fitting null component of
relative to ad A is minimal if and only if A has n distinct characteristic roots. Thus A is regular in
=
L if and only if A has n distinct characteristic roots. The corresponding Cartan subalgebra
is the Fitting null component of
relative to ad A and dim
= n.
Since A has n distinct characteristic roots we can choose a basis for such that the matrix of A relative to this basis is diagonal. Let
1 be the set of linear transformations whose matrices relative to this basis are diagonal. Then dim
1 = n and [HA] = 0 for every H ∈
1. Hence
. On the other hand, dim
= n. Hence
and the Cartan subalgebra determined by A is just the centralizer of the linear transformation A.
Let (e1, e2, …, en) be a basis such that eiA = αiei. Then we have seen that the Cartan subalgebra is the set of linear transformations which have diagonal matrices relative to the basis (ei). Hence if H ∈
then we have
, and we may assume the notation is such that
. The spaces
are weight spaces and Φei corresponds to the weight
. We shall therefore write
for Φei. As for the single linear transformation A we can now write
and
is a weight space for the Cartan subalgebra
corresponding to the weight λi – λj. We can summarize our results in the following
THEOREM 2. Let =
L, the Lie algebra of linear transformations in an n-dimensional vector space
over an algebraically closed field Φ. Then A ∈
is a regular element if and only if the characteristic polynomial det (λ1 – A) has n distinct roots. The Cartan subalgebra
determined by A is the set of linear transformations H such that [HA] = 0. If the weights of H acting in
are
then the roots (weights of the adjoint representation in
) are
.
If Φ is not algebraically closed but is infinite then the extension field argument of § 1 shows that again A is regular in L if and only if A has n distinct characteristic roots (in the algebraic closure Ω of Φ). It is well known that the centralizer of such an A is the algebra Φ[A] of polynomials in A and dim Φ[A] = n. Since the dimensionality of the Cartan subalgebra determined by A is n, it follows that this subalgebra is Φ[A].
If is the orthogonal or symplectic Lie algebra in a 2l-dimensional space over an infinite field one can show that the rank is N – l, N = dim
(cf. Exercises 4 and 5). It is not difficult to determine Cartan subalgebras for the other important examples of Lie algebras which we have encountered (cf. Exercise 3).
We now consider a finite-dimensional Lie algebra over an algebraically closed field Φ, a nilpotent subalgebra of
and
a module for
, which is finite-dimensional over Φ. Let
be the decompositions of and
into weight modules relative to
. We have seen that
Now suppose is a Cartan subalgebra. Then
=
0, the root module corresponding to the root 0. Also we have
= [
] = Σ[
α
β] where the sum is taken over all the roots α, β. The formula for [
α
β] shows that
where the summation is taken over all the α such that — α is also a root (e.g., α = 0).
We now prove the following
LEMMA 1. Let Φ be algebraically closed of characteristic 0, a Cartan subalgebra of
over Φ,
a module for
. Suppose a is a root such that — α is also a root. Let
. Then ρ(hα) is a rational multiple of α(hα) for every weight ρ of
in
.
Proof: Consider the functions of the form ρ(h) + iα(h), i = 0, ± 1, ± 2, …, which are weights, and form the subspace =
summed over the corresponding weight spaces. This space is invariant relative to
and, by (20), it is also invariant relative to the linear transformations eαR, e-αR where R is the representation of
determined by
. Hence, if tr
denotes the trace of an induced mapping in
, then
. On the other hand, the restriction of hαR to
σ has the single characteristic root σ(hα). Hence
where, in general, nσ = dim σ. Thus we have
Since Σinρ + iα is a positive integer this shows that ρ(hα) is a rational multile of α(hα).
We can now prove
Cartan’s criterion for solvable Lie algebras. Let be a finitedimensional Lie algebra over a field of characteristic 0. Suppose
has a finite-dimensional module
such that (1) the kernel
of the associated representation R is solvable and (2) tr (aR)2 = 0 for every a ∈
. Then
is solvable.
Proof: Assume first that the base field Φ is algebraically closed. It suffices to prove that for, conditions (1) and (2) carry over to
and
as
-module. Hence we shall have that
We therefore suppose that =
. Let
be a Cartan subalgebra and let the decomposition of
and
relative to
be as in (19). Then (21) implies that
= Σ[
α
β] summed on the α such that – α is also a root. Choose such an α, let eα ∈
α, e-α ∈
-α, and consider the element hα = [eαe-α]. The formula
=
implies that every element of
is a sum of terms of the form [eαe-α]. The restriction of hαR to
ρ has the single characteristic root ρ(hα). Hence the restriction of (hRα)2 has the single characteristic root ρ(hα)2 and if nρ is the dimensionality of
ρ, then we have
By the lemma, ρ(hα) = rρα(hα), rρ rational. Hence α(hα)2(Σnρr2ρ) = 0. Since the nρ are positive integers, this implies that α(hα) = 0 and hence ρ(hα) = 0. Since the ρ are linear functions and every h ∈ is a sum of elements of the form hα, hβ, …, etc., we see that ρ(h) = 0. Thus 0 is the only weight for
; that is, we have
=
0. If α is a root then (20) now implies that
α = 0 for every α ≠ 0. This means the kernel
of R contains all the
α, a ≠ 0. Hence
is a homomorphic image of
. Thus
R is nilpotent and
is solvable contrary to
' =
.
If the base field Φ is not algebraically closed, then let Ω be its algebraic closure. Then Ω is a module for
Ω and
Ω is the kernel of the corresponding representation. Since
is solvable,
Ω is solvable. Next we note that the condition tr (aR)2 = 0 and tr aRbR = tr bRaR imply that
Hence if the ai ∈ and ωi∈ Ω, then tr
Hence the condition (2) holds also in
Ω. The first part of the proof therefore implies that
Ω is solvable. Hence
is solvable and the proof is complete.
COROLLARY. If Φ is of characteristic 0 then is solvable if and only if tr(ad a)2 = 0 for every a ∈
.
Proof: The sufficiency of the condition is a consequence of Cartan’s criterion since the kernel of the adjoint representation is the center. Conversely, assume solvable. Then, by Corollary 2 to Theorem 2.8, applied to ad
, the elements ad
a, a ∈
, are in the radical of (ad
)*. Hence ad
a is nilpotent and tr(ad
a)2 = 0.
Let R be a representation of a Lie algebra in a finite-dimensional space . Then the function
is evidently a symmetric bilinear form on with values in Φ. Such a form will be called a trace form for
. In particular, if
is finite-dimensional, then we have the trace form tr(ad a)(ad b), which we shall call the Killing form of
. If f is the trace form determined by the representation R then
A bilinear form f(a, b) on which satisfies this condition
is called an invariant form on . Hence we have verified that trace forms are invariant. We note next that if f(a, b) is any symmetric invariant form on
, then the radical
⊥ of the form; that is, the set of elements z such that f(a, z) = 0 for all a ∈
, is an ideal. This is clear since f(a, [zb]) = — f([ab], z) = 0.
We can now derive
Cartan’s criterion for semi-simplicity. If is a finite-dimensional semi-simple Lie algebra over a field of characteristic 0, then the trace form of any 1:1 representation of
is non-degenerate. If the Killing form is non-degenerate, then
is semi-simple.
Proof: Let R be a 1:1 representation of in a finite-dimensional space
and let f(a, b) be the associated trace form. Then
⊥ is an ideal of
and f(a, a) = tr(aR)2 = 0 for every a ∈
⊥. Hence
⊥ is solvable by the first Cartan criterion. Since
is semi-simple,
⊥ = 0 and f(a, b) is non-degenerate. Next suppose that
is not semi-simple. Then
has an abelian ideal
≠ 0. If we choose a basis for
such that the first vectors form a basis for
, then the matrices of ad a, a ∈
, and ad b, b ∈
, are, respectively, of the forms
This implies that tr(ad b)(ad a) = 0. Hence and the Killing form is degenerate.
If f(a, b) is a symmetric bilinear form in a finite-dimensional space and (e1e2, …, en) is a basis for the space, then it is well known that f is non-degenerate if and only if det (f(ei, ej)) ≠ 0. If is a finite-dimensional Lie algebra of characteristic zero with basis (e1e2, …, en) and we set βij = tr ad ei ad ej, then
is semisimple if and only if det (βij) ≠ 0. This is the determinant form of Cartan’s criterion which we have just proved. If Ω is an extension of the base field of
then (e1e2, …, en) is a basis for
ω over Ω. Hence it is clear that we have the following consequence of our criterion.
COROLLARY. A finite-dimensional Lie algebra over a field Φ of characteristic zero is semi-simple if and only if
Ω is semi-simple for every extension field Ω of Φ.
We are now in a position to obtain the main structure theorem on semi-simple Lie algebras. The proof of this result which we shall give is a simplification, due to Dieudonné, of Cartan’s original proof. The argument is actually applicable to arbitrary non-associative algebras and we shall give it in this general form.
Let be a non-associative algebra over a field Φ. A bilinear form f(a, b) on
(to Φ) is called associative if
If f is an invariant form in a Lie algebra then f([ac], b) + f(a, [bc]) = 0. Hence f([ac], b) — f(a, [cb]) = 0 and f is associative. Now let f(a, b) be a symmetric associative bilinear form on and let
be an ideal in
. Let a ∈
⊥ so that f(a, b) = 0 for all b ∈
. Then for any c in
, f(ac, b) = f(a, cb) = 0 since cb ∈
. Also f(ca, b) = f(b, ca) = f(bc, a) = 0 since bc ∈
. Hence
⊥ is an ideal.
The importance of associative forms is indicated in the following result.
THEOREM 3. Let be a finite-dimensional non-associative algebra over a field Φ such that (1)
has a non-degenerate symmetric associative form f and (2)
has no ideals
with
2 = 0. Then
is a direct sum of ideals which are simple algebras.
(We recall that simple means that
has no ideals ≠ 0,
, and
2 ≠ 0.)
Proof: Let be a minimal ideal (≠ 0) in
. Then
⊥; is an ideal contained in
. Hence either
⊥ =
or
⊥ = 0. Suppose the first case holds and let b1, b2 ∈
, a ∈
. Then f(b1b2, a) = f(b1, b2a) = 0. Since f is non-degenerate b1b2= 0 and
2 = 0 contrary to hypothesis. Hence
⊥ = 0. It is well known that this implies that
=
⊕
⊥ and
⊥ is an ideal. This decomposition implies that
⊥ = 0 =
⊥
hence every
-ideal is an ideal. Consequently,
is simple. Moreover,
⊥ satisfies the same conditions as
since the restriction of f to
⊥ is nondegenerate and any
⊥-ideal is an ideal. Hence, induction on dim
implies that
⊥ =
2 ⊕ … ⊕
r where the
i are ideals and are simple algebras. Then for
1 =
we have
=
1⊕
2⊕ …
r,
i, simple and ideals.
This result and the non-degeneracy of the Killing form for a semi-simple Lie algebra of characteristic zero imply the difficult half of the fundamental
Structure theorem. A finite-dimensional Lie algebra over a field of characteristic 0 is semi-simple if and only if where the
i are ideals which are simple algebras.
Proof: If is semi-simple, then
has the structure indicated. Conversely, suppose
ideals and simple. We consider the set of linear transformations ad
= {ad a | a ∈
} acting in
. The invariant subspaces relative to this set are the ideals of
. Since
where the
i are irreducible, we see that the set ad
is completely reducible. Hence if
is any ideal ≠ 0 in
, then
=
⊕
where
is an ideal (Theorem 2.9). Moreover, the proof of the theorem referred to shows that we can take
to have the form
for a subset {
iu} of the
i. Then
where the
iv are the remaining
i. Since
i is simple,
2i =
i. Hence
and consequently
2 =
. Thus
is not solvable. We have therefore proved that
has no non-zero solvable ideals; so
is semi-simple.
The argument just given that has the following consequence.
COROLLARY 1. Any ideal in a semi-simple Lie algebra of characteristic 0 is semi-simple.
If i is simple then the derived algebra
hence the structure theorem implies the following
COROLLARY 2. If is semi-simple of characteristic 0, then
=
.
Remark. We have proved in Chapter II that if is a completely reducible Lie algebra of linear transformations in a finite-dimensional vector space over a field of characteristic 0, then
=
⊕
1 where
is the center and
1 is a semi-simple ideal. Then
' =
' =
1. Hence
=
⊕
,
semi-simple.
We prove next the following general uniqueness theorem.
THEOREM 4. If is a non-associative algebra and
where the i and
j are ideals and are simple, then r = s and the
i’s coincide with the
i’s (except for order).
Proof: Consider . This is an ideal contained in
1 and
j. Hence if
since
1 and
j are simple. It follows that
for at most one
j. On the other hand, if
for all j then
for all j. Since
, this implies that
1
= 0 contrary to the assumption that
21 ≠ 0. Hence there is a j such that
1 =
j. Similarly, we have that every
i coincides with one of the
j and every
j coincides with one of the
i.
It is easy to see also that if is as in Theorem 4, then
has just 2r ideals, namely, the ideals
, a subset of {1, 2, …, r). We omit the proof.
The main structure theorem fails if the characteristic is p ≠ 0. To obtain a counter example we consider the Lie algebra L of linear transformations in a vector space
whose dimensionality n is divisible by p. It is easy to prove (Exercise 1.20) that the only ideals in
L are
'L and Φ1, the set of multiples of 1. Since
′L is the set of linear transformations of trace 0 and tr 1 = n = 0,
. Hence
=
has only one ideal, namely,
, and the latter is simple. This implies that
is semi-simple, but since
is the only ideal in
is not a direct sum of simple ideals. This and Theorem 3 imply that
L/Φ1 possesses no nondegenerate symmetric associative bilinear form.
We conclude this section with the following characterization of the radical in the characteristic 0 case.
THEOREM 5. If is a finite-dimensional Lie algebra over a field of characteristic 0, then the radical
of
is the orthogonal complement
⊥ of
relative to the Killing form f(a, b).
Proof: =
⊥ is an ideal and if b ∈
', then tr(ad
b)2 = f(b, b) = 0. The kernel of the representation
is abelian. Hence
is solvable, by Cartan’s criterion, and
. Next let s ∈
, a, b ∈
. Then f(s, [ab]) = f([sa], b). We have seen (Corollary 2 to Theorem 2.8) that ad [sa] is contained in the radical of the enveloping associative algebra (ad
)*. Consequently, ad [sa] ad b is nilpotent for every b and hence f([sa], b) = 0. Thus f(s, [ab]) = 0 and s ∈
⊥. Thus
.
We recall that ad a is a derivation called inner and the set ad of these derivations is an ideal in the derivation algebra
(
). In fact, we have the formula [ad a, D] = ad aD for D a derivation. Hence
which implies that
Thus, for the Killing form f(a, b) = tr a ad b we have
that is, every derivation is a skew-symmetric transformation relative to the Killing form.
We prove next the following theorem which is due to Zassenhaus.
THEOREM 6. If is a finite-dimensional Lie algebra which has a non-degenerate Killing form, then every derivation D of
is inner.
Proof: The mapping is a linear mapping of
into Φ; that is, it is an element of the conjugate space
* of
Since f(a, b) is non-degenerate it follows that there exists an element d ∈
such that f(d, x) — (ad x)D for all x ∈
.
Let E be the derivation D — ad d. Then
Thus
Now consider
by (26). Since f is non-degenerate, this implies that E = 0. Hence D = ad d is inner.
This result implies that the derivations of any finite dimensional semi-simple Lie algebra over a field of characteristic zero are all inner. We recall also that if is solvable, finite-dimensional of characteristic 0, then
is mapped into the nil radical
by every derivation of
(Corollary 2 to Theorem 2.13). We can now prove
THEOREM 7. (1) Let be a finite-dimensional Lie algebra over a field of characteristic 0,
the radical,
the nil radical. Then any derivation D of
maps
into
. (2) Let
be an ideal in a finitedimensional algebra
1,
1,
1 the radical and nil radical of
1.
Proof: We first prove (2) for the radical. Thus it is clear that is a solvable ideal in
, hence
. Then
is a solvable ideal in
. On the other hand,
, which is an ideal in
. Hence
and
are semi-simple. Hence
. Now let
1 be the holomorph of
the radical and nil radical of
1. Then we know that
(Theorem 2.13). Since
by the first part of the argument,
. This implies that every derivation of
maps
into
, which proves (1). Now let
1 be any finite-dimensional Lie algebra containing
as an ideal. If a1 ∈
1 then ad a1 induces a derivation in
. Hence
. This means that
is an ideal in
1 so that
the nil radical of
1. Since the reverse inequality is clear,
.
This result fails for characteristic p ≠ 0. To construct a counter example we consider first the commutative associative algebra with the basis
with zp = 0. The radical
of
has the basis (z, z2, …, zp-l) and
. It is easy to prove that if w is any element of
then there exists a derivation of
mapping z into w. In particular, there is a derivation D such that zD = 1. Now let
be any simple Lie algebra and let
be the Lie algebra
. The elements of this algebra have the form
and
. Then
is a Lie algebra (Exercise 1.23) and
is a nilpotent ideal in
. Moreover,
is simple. Hence
is the radical and the nil radical of
. If D is any derivation in the associative algebra
, then the mapping
is a derivation in
. If we take D so that zD = 1 and let b ≠ 0 in
then
. Hence we have a derivation which does not leave the radical invariant.
In this section we shall prove the main structure theorem for modules of a semi-simple Lie algebra of characteristic zero and we shall obtain its most important consequences. The main theorem is due to Weyl and was proved by him by transcendental methods based on the connection between Lie algebras and compact groups.
The first algebraic proof of the result was given by Casimir and van der Waerden. The proof we shall give is in essence due to Whitehead. It should be mentioned that Whitehead’s proof was one of the stepping stones to the cohomology theory of Lie algebras which we shall consider in § 10. We note also that in the characteristic p case there appears to be little connection between the structure of a Lie algebra and the structure of its modules since, as will be shown later, every finite-dimensional Lie algebra of characteristic p ≠ 0 has faithful representations which are not completely reducible and also faithful representations which are completely reducible.
We obtain first a criterion that a set Σ of linear transformations in a finite-dimensional vector space be completely reducible. We have seen (Theorem 2.9) that Σ is completely reducible if and only if every invariant subspace
of
has a complement
which is invariant relative to Σ. Now let
’ be any complementary subspace to
:
=
⊕
’. Such a decomposition is associated with a projection E of
onto
. Thus if x ∈
, then we can write x in one and only one way as x = y + y', y ∈
, y’ ∈
’, and E is the linear mapping x → y. Conversely, if E is any idempotent linear mapping such that
=
E then
’ =
(l — E) is a complement of
in
. Now let A ∈ Σ and consider the linear transformation [AE] = AE — EA. If x ∈
, xAE ∈
and xEA ∈
hence [AE] maps
into
. If y ∈
then yAE = yA. Hence [AE] maps
into 0. Then if
denotes the set of linear tranformations of
which map
into
and
into 0, [AE] ∈
. It is clear that
is a subspace of the space
of linear transformations in
. We now prove the following
LEMMA 2. has a complement
which is invariant if and only if there exists a D ∈
such that [AE] = [AD] for all A ∈ Σ. Here E is any projection onto
.
Proof: Let be a complement of
which is invariant relative to Σ and let F be the projection of
onto
determined by the decomposition
=
⊕
Since
is invariant, F commutes with every A ∈ Σ, that is, [AF] = 0. Hence [AD] = [AE] for D = E — F. Also, since E and F are projections on
, E — F maps
into
and
into 0. Hence D ∈
as required. Conversely, suppose there exists a D ∈
such that [AE] = [AD]. Then F = E — D commutes with every A ∈ Σ. If x ∈
then xF = x(E — D) ∈
and if y ∈
then yF = yE = y. Hence F2 = F and
=
F. Then
=
(1 — F) is a complement of
and
is invariant under Σ since F commutes with every A ∈ Σ.
Suppose now that is a Lie algebra and
is an
-module,
a submodule. We can apply our considerations to the set
R of representing transformations aR determined by
. Let E be a projection of
onto
. If a ∈
, set f(a) = [aR, E]. Then a → f(a) is a linear mapping of
into the space
of linear transformations of
which map
into
and
into 0. If X ∈
and a ∈
then [XaR] ∈
. Thus if x ∈
then x[XaR] ∈
and if y ∈
then yXaR = 0 and yaRX = 0. We denote the mapping X → [XaR by aR. It is immediate that a → αR is a representation of
whose associated module is the space
. We have
We are now led to consider the following situation: We have a module for
and a linear mapping a → f(a) of
into
such that
A “trivial” example of such a mapping is obtained by taking f(a) = da, where d is an element of . For, we have
The key result for the proof of complete reducibility of the modules for a semi-simple Lie algebra of characteristic zero is the following
LEMMA 3. (Whitehead). Let be finite-dimensional semi-simple of characteristic zero and let
be a finite-dimensional module for
and a → f(a) a linear mapping of
into
satisfying (27). Then there exists a d ∈
such that f(a) = da.
Proof. The proof will be based on the important notion of a Casimir operator. First, suppose that is a Lie algebra and
1 and
2 are ideals in
such that the representations of
in
1 and
2 are contragredient. Thus we are assuming that the spaces
1 and
2 are connected by a bilinear form (b1, b2), bi ∈
i, (b1 b2) ∈ Φ, which is non-degenerate and that for any a ∈
, we have
If (u1, …, um) is a basis for 1 then we can choose a complementary or dual basis (u1, u2, …, um) for
2 satisfying (ui uj) = δij. Let [uia] =
Then
Hence (28) implies that = αik = — βki that is, the matrices (α) and (β) determined by dual bases satisfy (β) = — (α)′ ((α)′ the transpose of (α)). Now let R be a representation of
. Then the element
is called a Casimir operator of R. We have
Hence we have the important property that Γ commutes with all the representing transformations aR.
Now let satisfy the hypotheses of the lemma. Let
be the kernel of the representation R determined by
. Then we can write
where
1 is an ideal. Then the restriction of R to
is semi-simple. Hence the trace form (b1 b2) ≡
is non-degenerate on
1. Also we know that the trace form of a representation is invariant. Hence the equation (28) holds for bi ∈
1 and α ∈
. Thus the representation of
in
1 coincides with its contragredient and if (u1, …, um), (u1, …, um)are bases for
1 satisfying
is a Casimir operator which commutes with every aR. We note also that tr
We now decompose into its Fitting components
0 and
1 relative to Γ so that Γ induces a nilpotent linear transformation in
0 and a non-singular one in
1. Since
so that the
j are submodules. We can write f(a) = f0(a) + (a) where
and it is immediate that a → fj(a) is a linear mapping of
into
îj satisfying (27). Now if both spaces
are ≠ 0, then dim
< dim
for j = 0, 1. Hence we can use induction on dim
to conclude that there is a dj ∈
j such that fj(α) = dja. Then d — d0 + d1 satisfies f(α) = da as required. Thus it remains to consider the following two cases:
=
0 and
=
1.
=
0: In this case Γ is nilpotent. Hence m — tr Γ = 0. This means that the kernel of R is the whole of
, that is, ar = 0 for all α. Then the condition (27) is that f([ab]) = 0, a, b ∈ S. Thus f(a’) = 0 for all a’ ∈
'. Since
' =
, this implies that f(a) = 0 so that d = 0 satisfies the condition.
=
1: Set
where the (ui) and (ui) are dual bases for
i as before. Then
Since Γ is non-singular, d = Γ1 satisfies the required condition f(α) — da. This completes the proof of Whitehead’s lemma.
We can now prove the following fundamental theorem:
THEOREM 8. If is finite-dimensional semi-simple of characteristic0, then every finite-dimensional module for
is completely reducible.
Proof: Let be a finite-dimensional
-module,
a submodule. Let
be the space of linear transformations of
which map
into
,
into 0, and consider
as
-module relative to the composition Xa ≡ [X, aR], R the representation of
. Let E be any projection of
onto
and set f(α) — [αR, E]. Then f(a) satisfies the conditions of Whitehead’s lemma. Hence there exists a D ∈
such that f(a) = Da — [D, aR]. As we saw before, this implies that
has a complementary subspace which is invariant under
. Since this applies to every submodule
,
is completely reducible.
If is a subalgebra of a Lie algebra
, then a derivation D of
into
is a linear mapping of
into
such that
for every l1, l2 ∈ . It is immediate that the set
(
,
) of derivations of
into
is a subspace of the space of linear transformations of
into
. Whitehead’s lemma to the theorem on complete reducibility has the following important consequence on derivations:
THEOREM 9 Let be a finite-dimensional Lie algebra of characteristic 0 and let
be a semi-simple subalgebra of
. Then every derivation of
into
can be extended to an inner derivation of
.
Proof: Consider as
-module relative to the multiplication [bl], l ∈
, b ∈
. Then a derivation D of
into
defines f(l) = lD satisfying the condition
of Whitehead’s lemma. Hence there exists a d ∈ such that lD = f(l) = [d, l]. Then D can be extended to the inner derivation determined by the element — d.
We recall that we have shown in Chapter II (Theorem 2.11) that if is a completely reducible Lie algebra of linear transformations in a finite-dimensional vector space
over a field of characteristic 0, then
=
1 ⊕
where
1 is a semi-simple ideal and
is the center. Moreover, the elements C ∈
are semi-simple in the sense that their minimum polynomials are products of distinct irreducible polynomials. We are now in a position to establish the converse of this result. Our proof will be based on a field extension argument of the following type: Suppose we have a set Σ of linear transformations in
over Φ. If Ω is an extension field of Φ, every A ∈ Σ has a unique extension to a linear transformation, denoted again by A, in
Ω. In this way we get a set Σ = {A} of linear transformations in
Ω over Ω. We shall now prove the following
LEMMA 4. Let Σ be a set of linear transformations in a finitedimensional vector space over Φ and let Σ be the set of extensions of these transformations to
Ω over Ω, Ω an extension field of Φ. Suppose the set Σ in
Ω is completely reducible. Then the original Σ is completely reducible in
.
Proof: Let be a subspace of
which is invariant under Σ and let E be a projection on
. Then our criterion for complementation (Lemma 2) shows that
will have a complement which is invariant relative to Σ if and only if there exists a linear transformation D of
mapping
into
,
into 0 such that [AE] = [AD] for all A ∈ Σ. If A1, A2, …, Ak is a maximal set of linearly independent elements in Σ, and we set Bi = [AiE], then it suffices to find a D such that [AiD] = Bi, i = 1, 2, …, k. This is a system of k linear equations for D in the finite-dimensional space
of linear transformations of
mapping
into
,
into 0. Thus if we have the basis (U1, U2, …, Ur) for
we can write Bi =
then our equations are equivalent to the ordinary system:
s = 1, 2, …, k for the δj in Φ. Hence
has a Σ-invariant complement if and only if this system has a solution. We now pass to
Ω and the invariant subspace
Ω relative to the set Σ of extensions of the A ∈ Σ. Then our hypothesis is that
Ω has a Σ-invariant complement in
Ω. Now the extension E of E is a projection of
Ω onto
Ω. Hence we have a linear mapping D of
Ω mapping
Ω into
Ω,
Ω into 0, such that
The extensions Uu U2, …, Ur form a basis for the space of linear transformations of
Ω mapping
Ω into
Ω,
Ω into 0. Hence if
satisfy the system
. Since the γihs and βis belong to Φ, it follows that this system has a solution (δ1, …, δr), δ’s in Φ. Hence there exists a D ∈
such that [AD] = [AE], A ∈ Σ, and so
has a Σ-invariant complement in
.
We can now prove the following
THEOREM 10. Let be a Lie algebra of linear transformations in a finite-dimensional vector space
over a field of characteristic zero. Then
is completely reducible in
if and only if the following conditions hold: (1)
=
1 ⊕
,
1 a semi-simple ideal and
the center and (2) the elements of
are semi-simple.
Proof: The necessity has been proved before. Now assume (1) and (2) and let Ω be the algebraic closure of the base field. Then the lemma shows that it suffices to prove that the set of extensions of the elements of
is completely reducible in
Ω. The set of Ω-linear combinations of the elements of
can be identified with
Ω and similar statements hold for
1 and
Now let C ∈
Since the minimum polynomial of C in
has distinct irreducible factors and since the field is of characteristic 0, the minimum polynomial of C in
Ω has distinct linear factors in Ω. Consequently, we can decompose
Ω as
α1 ⊕ … ⊕
αk where
and α1, α2, …, αr are the different characteristic roots of C. Since AC = CA for A ∈ ,
αi A ⊆
αi. We can apply the same procedure to the
αi relative to any other D ∈
. This leads to a decomposition of
Ω =
1 ⊕
2 ⊕ … ⊕
r into
-invariant subspaces such that the transformation induced in the
i by every C ∈
is a scalar multiplication. To prove
completely reducible in
Ω it suffices to show that
1 the sets of induced transformations in the
i are completely reducible and since the elements of
are scalars in
i it suffices to show that
1 is completely reducible in every
i. The invariant subspaces of
i relative to
1 are invariant relative to Ω
1, the set of Æ-linear combinations of the elements of
1. Now Ω
1 is a homomorphic (actually isomorphic) image of the extension algebra
1 Ω, which is semi-simple. Hence Ω
1 is semi-
1mple and consequently this Lie algebra of linear transformations is completely reducible by Theorem 8. Thus we have proved that
is completely reducible in
Ω and hence in
.
We now shift our point of view and consider a finite dimensional Lie algebra of characteristic 0 and two finite-dimensional completely reducible modules
and
for
. We shall show that
⊗
is completely reducible. Now the space
=
⊕
is a module relative to the product (x + y)l = xl + yl, x ∈
, y ∈
. Evidently
is completely reducible and
⊗
is a submodule of
⊗
Hence it suffices to prove that
⊗
is completely reducible. If we replace
by
/
where
is the kernel of the representation in
, then we may assume that the associated representation R in
is 1: 1. Then we know that
=
1 ⊕
where
1 is a semi-simple ideal and
is the center. Moreover, the elements CR C ∈
, are semi-simple. Now, in general, if R is a faithful representation of a Lie algebra
, then the representation R ⊗ R in
⊗
is also faithful. Thus, if a ∈
and aR is not a scalar multiplication, then, since the algebra of linear transformations in
⊗
is the tensor product of the algebras of linear transformations in
aR ⊗ aR, aR ⊗ 1, 1 ⊗ aR and 1 ⊗ 1 are linearly independent, so aR ⊗ 1 + 1 ⊗ aR Φ 0. Hence if aR ⊗ R = 0, aR must be a scalar, say aR = α. Then aR⊗R = 2α (in
⊗
) and α = 0. since R is 1:1 this implies that a = 0. We can now conclude that
where
is semi-simple and
R ⊗ R is the center. Our result will therefore follow from the criterion of Th. 10 provided that we can prove that every
C ∈ , is semi-simple.
Let Ω be the algebraic closure of the base field and let α1 α2, …, αk be the different characteristic roots of CR. Then the proof of Theorem 10 shows that where xσiCR =
for
Hence
and
for every
It follows that the minimum polynomial of CR⊗R has distinct roots in (
⊗
)Ω since this is also the minimum polynomial of CR⊗R in
⊗
, it follows that this polynomial is a product of distinct irreducible factors. Thus CR⊕R is semi-simple and we have proved
THEOREM 11. Let be a finite-dimensional Lie algebra over a field of characteristic zero and let
and
be finite-dimensional completely reducible modules for
. Then
⊗
is completely reducible.
In § 1.4 we called a three-dimensional simple Lie algebra split if
contains an element h such that ad h has a non-zero characteristic root ρ belonging to the base field. We showed that any such algebra has a basis (e, f, h) with the multiplication table
The representation theory of this algebra is the key for unlocking the deeper parts of the structure and representation theory of semisimple Lie algebras (Chapters IV, VII, and VIII). We consider this now for the case of a field Ф of characteristic 0. We suppose first that Ф is algebraically closed and that is a finite-dimensional module for
. The representation in
is determined by the images E, F, H of the base elements e, f, h and we have
Conversely, any three linear transformations E, F, H satisfying these relations determine a representation of and hence
module. Let α be a characteristic root of H and x a corresponding characteristic vector: x ≠ 0, xH = αx. Then
If xE ≠ 0 then (32) shows that α + 2 is a characteristic root for H and xE a corresponding characteristic vector. We can replace x by xE and repeat the process. This leads to a sequence of nonzero vectors x, xE, xE2, …, belonging to the characteristic roots α, α + 2, α + 4, …, respectively, for H. Now H has only a finite number of distinct characteristic roots; hence, our sequence breaks off and this means that we obtain a k such that xEk ≠ 0 and xEk + 1 = 0.
If we replace x by xEk we may suppose at the start that x ≠ 0 and
Now set x0 = x and let xi = xi – 1F. Then, analogous to (32), we obtain
and the argument used for the vectors xEi shows that there exists a non-negative integer m such that x0, x1, …, xm are ≠ 0 but xm + 1 = 0. Thus xFm + 1 = 0, xFm ≠ 0.
Then xi 0 i
m, is a characteristic vector of H belonging to the characteristic root α ‒ 2i. since α, α – 2, α – 4, …, α – 2m are all different it follows that the xi are linearly independent. Let
so that
is an (m + 1)-dimensional subspace of
. We shall now show that
is invariant and irreducible relative to We first establish the formula
Thus we have x0E = 0 as given in (35). Assume (35) for i – 1. Then
as required. It is now clear from (34), (35), and xiF = xi + 1 that is a
-subspace of
. since H = [EF] we must have tr
H = 0. This, using (34), gives (m + 1)α – m(m + 1) = 0. Hence we obtain the result that α – m. Our formulas now read
and we note that in the last equation
– mi + i(i – 1) ≠ 0.
Now let 1 be a non-zero invariant subspace of
and let
βi ≠ 0, be in 1. Then
. Hence by the last equation of (36) every xi ∈
1 and
1 =
. Hence if
is
-irreducible to begin with, then
=
. In general, the theorem on complete reducibility shows that
is a direct sum of irreducible invariant subspaces which are like the space
.
We can now drop the hypothesis that Ф is algebraically closed, assuming only that Ф is of characteristic 0. We note first the following
LEMMA 5. Let be the split three-dimensional simple Lie algebra over a field Ω of characteristic zero and let e → E, f → F, h → H define a finite-dimensional representation of
. Then the characteristic roots of H are integers.
Proof. If is the module of the representation and Ω is the algebraic closure of Ф then
Ω is a module for
Ω which satisfies the same conditions over Ω as
over Φ. Then
Ω is a direct sum of irreducible subspaces
with bases (x0, x1, …, xm) satisfying (36). Hence if we choose a suitable basis for
Ω then the matrix of H relative to this is a diagonal matrix with integral entries. Hence the characteristic roots of H in
Ω are integers. These are also the characteristic roots of H in
.
We can now prove the following
THEOREM 12. Let be the split three-dimensional simple Lie algebra over a field of characteristic 0. Then for each integer m = 0, 1, 2, … there exists one and, in the sense of isomorphism, only one irreducible
-module
of dimension m + 1.
has a basis (x0, x1, …, xm) such that the representing transformations (E, F, H) corresponding to the canonical basis (e, f, h) are given by (36).
Proof. Let be a finite-dimensional irreducible module for
. Then the characteristic roots of H are integers. Hence we can find an integer α and a vector x ≠ in
such that xH = αx. As before we may suppose xE = 0. Then we obtain that α = m and that
has a basis (x0, x1, xm) such that (36) hold. These formulas are completely determined by the dimensionality m + 1 of
. Hence any two (m + 1)-dimensional irreducible modules for
are isomorphic. It remains to show that there is an irreducible (m + 1)-dimensional module for
for every m = 0, 1, …. To see this we let
be a space with the basis (x0, x1, …, xm) and we define the linear transformations E, F, H by (36). Then we have
Hence E, F, and H satisfy the required commutation relations and so they define a representation of As before, is
-irreducible.
The theorem of complete reducibility applies here also and together with the foregoing result gives the structure of any finitedimensional -module.
The “radical splitting” theorem of Levi asserts that if is a finite-dimensional Lie algebra of characteristic 0 with solvable radical
, then
contains a semi-simple subalgebra
=
+
such that
=
+
. It will follow that
= 0 so that
=
⊕
and
/
. Thus the subalgebra
is isomorphic to the difference algebra of
modulo its radical. Conversely, if
contains a subalgebra
isomorphic to
/
, then
is semi-simple. Hence
= 0 and since dim
= dim
+ dim
/
= dim
+ dim
,
=
+
.
We note next that it suffices to prove the theorem for the case 2 = 0, that is
is abelian. Thus suppose
2 ≠ 0. Then if
=
/
2, dim
< dim
. Hence if we use induction on the dimensionality we may assume the result for
. Now
=
/
2 is the radical of
and
/
/
. Hence
contains a subalgebra
. As subalgebra of
,
has the form
1/
2 where
1 is a subalgebra of
containing
2. Now
2 is the radical of
1 and
so that dim
1 < dim
. The induction hypothesis can therefore be used to conclude that
1 contains a subalgebra
≅
/
, and this completes the proof for
.
We now assume that 2 = 0 and for the moment we drop the assumption that
=
/
is semi-simple. Now
is a submodule of
for
(adjoint representation). since
2 = 0,
is in the kernel of the representation of
determined by the module
. Hence we have an induced representation for
=
/
. For the corresponding module we have s
= [s, b], s ∈
, b ∈
.
We can find a 1:1 linear mapping of
into
such that
Such a mapping is obtained by writing
=
⊕
where
is a subspace. Then we have a projection of
onto
defined by this decomposition. since
is the kernel, we have an induced linear isomorphism σ of
onto
hence into
. If b = s + g, s ∈
, then by definition
and
so that
as required. Conversely, let
be any 1: 1 linear mapping of
into
such that
Then
is a complement of
in
. If s ∈
and b ∈
, then
holds for the module multiplications in
.
Let and consider the element
If we apply the algebra homomorphism of
onto
and make use of the property
we obtain
and
Hence we see that
One verifies immediately that is a bilinear mapping of
into
.
Now suppose is a subalgebra of
. Then
∈
hence
so that we must have
for all
The converse is also clear since
implies that
Hence
is a subalgebra if and only if the bilinear mapping g is 0.
If σ is not a subalgebra, then we seek to modify σ to obtain a second mapping τ of
so that
τ is a subalgebra. Suppose this is possible. Then we have a 1: 1 linear mapping τ of
into
such that
and
for all
Now let ρ = σ – τ. Then ρ is a linear mapping of
into
such that
Hence , and we can consider ρ as a linear mapping of
into
. Also we have
If s is defined as before, we have
Thus, if we can somehow choose a complement of
which is a subalgebra then the bilinear mapping
of
×
into
can be expressed in terms of the linear mapping ρ of
into
by the formula
Conversely, suppose we have a linear mapping p of into
satisfying this condition. Then τ = σ – ρ is another 1: 1 linear mapping of
into
such that
and one can re-trace the steps to show that
so that
t is subalgebra.
Our results can be stated in the following way:
Criterion. Let be a Lie algebra,
an ideal in
such that
2 = 0 and set
Then
is a
-module relative to the composition s
= [sb]. Also there exist 1: 1 linear mappings σ of
into
such that
If σ is such a mapping then
Moreover, has a complementary space which is a subalgebra if and only if there exists a linear mapping ρ of
into
such that
We observe next that the bilinear mapping g, which we shall call a factor set in
, satisfies certain conditions which are consequences of the special properties of the multiplication in a Lie algebra. Thus it is clear that
which implies . We next write
and calculate
If we permute 1,
2,
3, cyclically, add, and make use of the Jacobi identities in
and
, we obtain
Our proof of Levi’s theorem will be completed by proving the following lemma, which is due to Whitehead.
LEMMA 6. Let be a finite-dimensional semi-simple Lie algebra of characteristic 0,
a finite-dimensional
-module and (l1, l2) → g(l1, l2) a bilinear mapping of
×
into
such that
Then there exists a linear mapping l → lρ of into
such that
Proof: Let ,
1, ui, ui, Γ be as in the proof of Whitehead’s first lemma:
is the kernel of the representation,
1 is an ideal such that
=
⊕
1, (ui) and (ui), i = 1, …, m, are dual bases of
1 relative to the trace form of the given representation, and Γ is the Casimir operator determined by the ui and ui We recall that γ is the mapping
in
. Set l3 = ui in (ii) and take the module product with ui. Add for i. This gives
If we make use of and recall that βij = – αji (cf. (28)) we can verify that
These and the skew symmetry of g permit the cancellation of four terms in the foregoing equations. Hence we obtain
If Γ is non-singular we define
Then (43) gives the required relation (iii). If Γ is nilpotent, then, as in the proof of Whitehead’s first lemma, m = 0, =
, so that the representation is a zero representation. Then (ii) reduces to
Now let denote the vector space of linear mappings of
into
. We make this into an
-module by defining for A ∈
, x, l ∈
, x(Al) ≡ – [xl]A that is, Al ≡ – (ad l)A. It is easy to see that this satisfies the module conditions (cf. § 1.6). For each l ∈
we define an element Al ∈
as the mapping x → g(x, l) ∈
. Then l → Al is a linear mapping of
into
and
Hence the skew symmetry of g and (ii′) imply that
Thus he hypothesis of Whitehead’s first lemma holds. The conclusion states that there exists a ρ ∈ such that Al = ρl. This means that we have a linear mapping ρ of
into
such that
By definition of as module, this gives (iii). This proves the result for the case Γ nilpotent. If Γ is neither non-singular nor nilpotent, then we have the decomposition of
as
0 ⊕
1 where the
i are the Fitting components of
relative to Γ and these are
0. These spaces are submodules and we can write g(l1, l2) = g0(l1, l2) + g1(l1, l2), gi ∈
i. Then the gi satisfy the conditions imposed on g, so we can represent these in the form (iii), by virtue of an induction hypothesis on the dimensionality of
. This gives the result for
by adding the linear transformations for the
i.
As we have noted before, the lemma completes our proof of
Levi’s theorem. If is a finite-dimensional Lie algebra of characteristic zero with radical
then there exists a semi-simple subalgebra
of
such that
=
⊕
.
A subalgebra satisfying these conditions is called a Levi factor of
. A first consequence of Levi’s theorem is the following result:
COROLLARY 1. Let ,
and
be as in the theorem. Then
.
Proof: We have =
⊕
so that
. since
have
.
We have seen that the nil radical of
(Theorem 2.13), so we can now state that
. We know also that the radical of an ideal is the intersection of the ideal with the radical of the containing algebra. Hence
′
is the radical of
′. We therefore have the following
COROLLARY 2. The radical of the derived algebra of a finitedimensional Lie algebra of characteristic 0 is nilpotent.
We take up next the question of uniqueness of the Levi factors. It will turn out that these are not usually unique; however, they are conjugate in a rather strong sense which we shall now define. We recall that if z ∈ , the nil radical of
, then ad z is nilpotent. since ad z is a derivation we know also that A = exp (ad z) is an automorphism. Let
denote the group of automorphisms generated by the elements exp (ad z), z ∈
. Then we have the following conjugacy
Theorem of Malcev-Harish-Chandra. Let =
⊕
where
is a solvable ideal and
is a semi-simple subalgebra and let
1 be a semi-simple subalgebra of
. Assume
finite-dimensional and of characteristic 0. Then there exists an automorphism A ∈
such that
.
Proof: Any l1 ∈ 1 can be written in one and only one way as
where
and
so that we have the linear mappings λ and σ of
1 into
and
, respectively. since
1 is semi-simple,
1
= 0; hence λ is 1:1. If l2 ∈
1 then
Hence
The second of these equations shows that the nil radical of
. since
this implies that
for every l1 ∈
1 and so
We shall prove by induction that there exists an automorphism
such that
where
(i) is the ith derived algebra of
. since
is solvable this will prove the result. since we have proved that
it suffices to prove the inductive step and we may simplify the notation and assume that
. Then we shall show that there exists A ∈
such that
. If we use the notation introduced before,
1
+
(k) implies that
The first equation in (48) implies that if we set
then this makes
(k) into an
1-module. Now
(k + 1) a submodule so that
(k)/
(k + 1) is an
1-module relative to
where z ∈
(k) and
. We now take the cosets relative to ^α+1)0f the terms in the second equation of (48). Since
we have
Now set is a linear mapping of
1 into the
1-module
(k)/
(k + 1) and the foregoing equation can be re-written as
Hence by Whitehead’s first lemma there exists a such that
which means that
Let A = exp (ad z). Then
Now and (51) shows that
We can therefore prove the result by induction on k.
COROLLARY 1. Any semi-simple subalgebra of a finite-dimensional Lie algebra of characteristic zero can be imbedded in a Levi factor.
Proof: If A is as in the theorem, then 1 is contained in the Levi factor
A – 1.
COROLLARY 2. If where
1 and
2 are semisimple subalgebras then there exists an automorphism A ∈
such that
.
This is an immediate consequence of the theorem.
The two lemmas of Whitehead can be formulated as theorems in the cohomology theory of Lie algebras. Historically, these constituted one of the clues which led to the discovery of this theory. Another impetus to the theory came from the study of the topology of Lie groups which was initiated by Cartan. In this section we give the definition of the cohomology groups which is concrete and we indicate an extension of the “Γ non-singular” case of Whitehead’s lemmas to a general cohomology theorem. Later (Chapter V) we shall give the definition of the cohomology groups which follows the general pattern of derived functors of Cartan-Eilenberg.
Let be a Lie algebra,
an
-module. If i
1, an i-dimensional
-cochain for
is a skew symmetric i-linear mapping of
×
× … ×
(i times) into
. Such a mapping f sends an i-tuple (l1, l2, … li), lq ∈
into f(l1, …, li) ∈
in such a way that for fixed values of l1, …, lq – 1, lq + 1, …, li the mapping lq → f(l1, …, fi) is a linear mapping of
into
. The skew symmetry means that f is changed to —f if any two of the li are interchanged (the remaining ones unchanged). If i = 0 one defines a 0-dimensional
-cochain for
as a “constant” function from
to
, that is, a mapping l →u, u a fixed element of
. If f is an i-dimensional cochain (or simply “an i-cochain”), i
0, f determines an (i + 1)-dimensional cochain fδ, called the coboundary of f, defined by the formula
Here the ^ over an argument means that this argument is omitted (e.g., . For i = 0 this is to be interpreted as (fδ)(l) = ul if f is the mapping x → u ∈
.
The set Ci(,
) of i-cochains for
is a vector space relative to the usual definitions of addition and scalar multiplication of functions. Moreover f → fδ is a linear mapping, the coboundary operator, of Ci(
,
) into Ci + 1(
,
), i
0. Besides the case
we have
An i-cochain f is called a cocycle if fδ = 0 and a coboundary if f = gδ for some (i – 1)-cochain g. The set Zi(,
) of i-cocycles is the kernel of the homomorphism δ of Ci into Ci + 1, so Zi is a subspace of Ci similarly, the set
of i-coboundaries is a subspace of Ci since it is the image under δ of Ci – 1. It can be proved fairly directly that
, that is, coboundaries are cocycles.
This amounts to the fundamental property: δ2 = 0 of the coboundary operator. We shall not give the verification in the general case at this point since it will follow from the abstract point of view later on. At this point we shall be content to verify fδ2 = 0 for f a 0-or a 1-cochain. Thus if f = u, that is, f is the mapping x → u then fδ(l) = ul and fδ2(l1 l2) = – ul2l1 + ul1l2 – u[l1l2 = 0 by the definition of a module. If f is a 1-cochain, fd(l1l2) is given by (54). Hence, by (55),
One checks that this sum is 0; hence fδ2 = 0 for any 1-cochain f.
Once the verification δ2 = 0 has been made, one can define the i-dimensional cohomology group (space) of relative to the module
as the factor space
. If i = 0 we agree to take Bi = 0 since there are no (i – 1)-cochains. Hence in this case it is understood that H0(
,
) = Z0(
,
). This can be identified with the subspace I(
) of elements u ∈
such that ul = 0 for all l. Such elements are called invariants of the module
. Hi(
,
) = 0 means that
that is, every i-cocycle is a coboundary. For i = 1 this states that if l → f(l) is a linear mapping of
into
such that
, then there exists a u in
such that f(f) = ul. This is just the type of statement which appears in Whitehead’s first lemma. similarly, Whitehead’s second lemma is a statement about the second cohomology groups. In fact, these two results can now be stated in the following way.
THEOREM 13. If is finite-dimensional semi-simple of characteristic 0, then H1(
,
) = 0 and H2(
,
) = 0 for every finite-dimensional module
of
.
It is easy to see that if =
⊕
2 where the
i are sub-modules of
, then
This and the theorem of complete reducibility permits the reduction of the Hi(
,
), for finite-dimensional
to the case
irreducible and here one distinguishes two cases: (1)
≠ 0 and (2)
= 0. In the second case irreducibility implies dim
= l, so
can be identified with the field Φ. Then an i-cochain is a skew symmetric i-linear function of (l1 · ··, li) with values in Φ, and since the representation is a zero representation, the coboundary formula reduces to
It turns out that for semi-simple Lie algebras the cohomology groups with values in = Φ are the really interesting ones, since these correspond to cohomology groups of Lie groups. On the other hand, the case
≠ 0 is not very interesting (for semi-simple 8, finite-dimensional irreducible
) except for its applications to the theorem of complete reducibility and the Levi theorem, since one has the following general result.
THEOREM 14 (Whitehead). Let be a finite-dimensional semisimple Lie algebra over a field of characteristic 0 and let
be a finite-dimensional irreducible module such that
≠ 0. Then
for all i
0.
If i = 0 the irreducibility and ≠ 0 imply that u
= 0 holds only for u = 0. This means that H0(
) = 0. The proof for i > 0 is similar to the proof of the case: Γ non-singular, in the two Whitehead lemmas. We leave the details to the reader.
For our further study of this question we require a notion of a type of closure for Lie algebras of linear transformations and an imbedding theorem for nilpotent elements in three-dimensional split simple algebras. The first of these is based on a special case of a property of associative algebras (the so-called Wedderburn princeal theorem), which is the analogue of Levi’s theorem on Lie algebras. The result is the following
THEOREM 15. Let be a finite-dimensional algebra (associative with identity 1) generated by a single element x over Φ of characteristic zero and let
be the radical of
. Then
contains a semisimple ubalgebra
1 such that
Proof: Let f(λ) be the minimum polynomial of x and let
be the factorization of f(λ) into irreducible polynomials with the leading coefficients one such that if i = j and deg πi(λ) > 0, We note first that if all the ei = 1, has no nonzero nilpotent elements (cf. p. 47), so
is semi-simple and there is nothing to prove. In any case, set
and z = f1(x). Then if e = max(ei), so that z is nilpotent. since
is commutative, the ideal (z) generated by z is nilpotent; hence (z) ⊆
. On the other hand, f1(x) = 0 (mod (z)). Hence the minimum polynomial of the coset
in
/(z) is a product of distinct prime factors. since x generates
/(z), this means that
/(z) is semi-simple. Hence (z) =
. It follows also easily that the minimum polynomial of
. Hence it suffices to prove that
contains an element y whose minimum polynomial is We shall obtain such an element by a method of ‘ ‘successive approximations” beginning with x1 = x To begin with we have f1(x1) ≡ 0 (mod
) and x ≡ xk (mod
). Suppose we have already determined xk such that f1(xk) ≡ 0 (mod
k) and x ≡ xk(mod
). Set xk + 1 = ×k + w where w is to be determined in
k so that f1(xk + i) ≡ 0 (mod
k +1). We have, by Taylor’s theorem for polynomials,
since the base field is of characteristic 0, f1(λ) has distinct roots in the algebraic closure of Φ. Hence fι(Λ) is prime to the derivative It follows that
has an inverse v in
/
. Set w = — fi(xk)v. Then w ≡ 0 (mod
k) so that
Thus we have determined xk + 1 such that f1(xk + i) = 0 (mod ) and y = x (mod
). since
is nilpotent this process leads to a y such that f1(y) = 0 and y = x (mod
). Hence
1 = λ [y] satisfies
=
⊕
. since the minimum polynomial of x is it follows that the minimum polynomial of y is f1(λ) also.
We can now prove
THEOREM 16. Let X be a linear transformation in a finite-dimensional vector space over a field of characteristic zero. Then we can write X = Y + Z where Y and Z are polynomials in X such that Y is semi-simple and Z is nilpotent. Moreover, if X = Yx + Zi where Y1 is semi-simple and Z1 is nilpotent and Yx and Zi commute with Z, then Y1 = Y, Z1 = Z.
Proof: The existence of the decomposition X = Y + Z is obtained by applying Theorem 15 to the algebra Φ[X]. Now suppose X = Y1 + Z1 where Y1 and Z1 have the properties stated in the theorem. since Y and Z are polynomials in X they commute with Y, and Z1. We have Y - Yt = Z1 - Z. since Z and Z1 are nilpotent and commute, Z - Z1 is nilpotent. since Y and Y1 are semi-simple and the base field is of characteristic zero, the proof of Theorem 11 shows that Y - Y1 is semi-simple. since the only transformation which is both semi-simple and nilpotent is 0,
Hence Y = Y1 Z = Z1.
We call the uniquely determined linear transformations Y and Z of Theorem 16 the semi-simple and the nilpotent components of X.
DEFINITION 3. A Lie algebra of linear transformations of a finite-dimensional vector space over a field of characteristic 0 is called almost algebraic* if it contains the nilpotent and semi-simple components of every X ∈
.
To prove our imbedding theorem we require the following two lemmas.
LEMMA 7. (Morozov). Let be a finite-dimensional Lie algebra of characteristic 0 and suppose
contains elements f, h such that [fh] = –2f and h ∈ [
f]. Then there exists an element e ∈
such that
* This concept is due to Malcev, who used the term splittable. We have changed the term to “almost algebraic” since this is somewhat weaker than Chevalley’s notion of an algebraic Lie algebra of linear transformations. Moreover, we have preferred to use the term “split Lie algebra” in a connection which is totally unrelated to Malcev’s notion.
Proof: There exists a z ∈ such that h = [zf]. Set F = adf, H = ad h, Z = ad z so that we have
The first of these relations implies that F is nilpotent (Lemma 2.4). Also
Hence [zh] = 2z + x1 where x1 ∈ the subalgebra of elements x such that [xf] = 0. since [FH] = –2F, if b ∈
then
Hence bH ∈ and so
H ⊆
. Also we have
and since HFk = FkH + 2kFk, we have
Let Then b = aFi-1 and bF= aFi = 0. Hence
Hence b(H +(i – 1)) ∈
Fi. It follows from this relation and the nilpotency of F that if b is any element of
then
for some positive integer m. Thus the characteristic roots of the restriction of H to are non-positive integers. Hence H – 2 induces a non-singular linear transformation in
and consequently there exists a y, ∈
such that y1(H – 2) where Xi is the element such that [zh] = 2z + x1. Then [y1h] = 2y1 + x1 Hence if we set e = z – y1 we have [eh] = 2e. Also [ef] = [zf] = h. Hence (59) holds.
LEMMA 8. Let be a Lie algebra of linear transformations in a finite-dimensional vector space over a field of characteristic 0. Suppose every nilpotent element F ≡ 0 of
can be imbedded in a subalgebra with basis (E, F, H) such that [EH] = 2E, [FH] — — 2F, [EF] = H. Let ⊕ be any subalgebra of
which has a complementary space
in
invariant under multiplication by
Then ⊕ has the property stated for
.
Proof Let F be a non-zero nilpotent of . Then we can choose E and H in
so that the indicated relations hold. Write H = H1 + H2 H1, ∈
, H2 ∈
, E = E1 + E2, E1 ∈
, E2 ∈
. Then we have –2F = [FH] = [FH] + [FH2] and [FH1] ∈ [FH2] ∈
. Hence 2F = [FH1]. Also H = [EF] = [E, F] + [E2F]. This implies that Hi = [E1 F1] ∈ [
F]. Thus H1 satisfies for F,
the conditions on H in Lemma 7. Hence there exists E′, H′ in
such that [FH1] = 2F, [E′H′] = 2E′, [E′F] = H′. The subalgebra generated by F, E′, H′ is a homomorphic image of the split three-dimensional simple algebra. since F ≠ 0 we have an isomorphism, so that F, E′, H′ are linearly independent and satisfy the required conditions
We can now establish our second criterion for complete reducibility.
THEOREM 17. Let be a Lie algebra of linear transformations in a finite-dimensional vector space
over a field of characteristic 0. (1) Assume
completely reducible. Then every non-zero nilpotent element of
can be imbedded in a three-dimensional split simple subalgebra of
and
is almost algebraic. (2) Assume that every non-zero nilpotent element of
can be imbedded in a three-dimensional simple subalgebra of
and that the center
of
is almost algebraic. Then
is completely reducible.
Proof: (1) Assume is completely reducible and let
denote the complete algebra of linear transformations in
. Let F be a nilpont linear transformation and let
be a decomposition of
into cyclic invariant subspaces relative to F. Thus in
i we have a basis (x0, x1, …, xmi) such that xjF = xj + 1 xmi F = 0. We define H and E to be the linear transformations leaving every
i invariant and satisfying xjH = (mi = 2j)xj, x0E = 0, xjE = mij + j(j – 1))xj– 1, j > 0 (cf. (36)). Then as in §8, [EH] = 2E, [FH] = – F, [EF] = H. This shows that F can be imbedded in a subalgebra ΦE + ΦF + ΦH of the type indicated. We shall show next that we can write
L =
⊕
where
is a subspace such that [
] ⊆
. It will then follow from Lemma 8 that every nilpotent element ≠ 0 of
can be imbedded in a split three-dimensional simple subalgebra of
. We recall that
L as module relative to
(adjoint representation) is equivalent to
⊕
*,
* the contragredient module. It is also easy to see that
* is completely reducible. Hence, by Theorem 11,
⊗
*, and consequently
L, is completely reducible relative to
. since
is a submodule of
L relative to
, there exists a complement
such that
L =
. This completes the proof of the first assertion in (1). Now let X be any element of
and let Y and Z be the semi-simple and nilpotent components of X. Then
and ad
Z is nilpotent (Zm = 0 implies (ad
Z)2m-1 = 0). Also the identification of
with
⊗
* and the proof of Theorem 11 show that ad
Y is semi-simple. Hence ad
Y and ad
Z are the semi-simple and nilpotent components of ad
X and so these are polynomials in ad
X. since
ad
X
and ad
Y, ad
Z are polynomials in ad
X,
ad
Y
,
ad
Z
. Thus L → [LY], L → [LZ] are derivations in
. We can write
=
′ ⊕
where
′ is semisimple and
is the center. since the derivations of
′ are all inner it follows that any derivation of
which maps
into 0 is an inner derivation determined by an element of
′. Since Z is a polynomial in X, [XC] = 0 implies [ZC] = 0. This implies that the derivation L → [LZ] maps
into 0. Hence there exists a Z1 ∈
′ such that [LZ] = [LZ1, L ∈
. since Z is nilpotent, ad
′Z1 = ad
′Z is nilpotent. since
′ is semi-simple, the result just proved (applied to ad
′) implies that there exists an element U ∈
′ such that [ad
′Z1, ad
′U] = 2ad
′Z1. Then [Z1U] = 2Z1, which implies that Z1 is nilpotent. since [XZ] = 01 [XZ1] = 0 and since Z is a polynomial in X, [ZZ1] = 0. It now follows that Z — Z1 is nilpotent. since [L, Z – Z1] = 0, L ∈
, and Z – Z1 is in the enveloping associative algebra
*, Z – Z1 is in the center of
*. since
* is completely reducible and Z – Z1 is nilpotent, this implies Z – Z1 = 0 so Z = Z1 ∈
. Hence also Y = X – Z ∈
. This completes the proof that
is almost algebraic. (2) Assume
has an almost algebraic center and has the property stated for nilpotent elements. Let
be the radical of
and let F ∈ [
]. Then we know that F is nilpotent (Corollary 2 to Theorem 2.8). If F is not zero it can be imbedded in a three-dimensional simple subalgebra
. since
∩
≠ 0 and
is simple,
which is impossible because of the solvability of
. Hence F = 0 and [
] = 0. This implies that
=
the center. By Levi’s theorem
=
⊕
1 where
1 is a semi-simple subalgebra. since
is the center this implies that
1 is an ideal. We can now invoke Theorem 10 to prove that
is completely reducible, provided that we can show that every C ∈
is semi-simple. Now we are assuming that C = D + E where D is semi-simple, E nilpotent, and D and E are in
. If E ≠ 0 we can imbed this in a three-dimensional simple subalgebra. Clearly this is impossible since E is in the center. This completes the proof of (2).
It is immediate that if is almost algebraic, then the center
of
is almost algebraic. Hence we can replace the assumption in (2) that
is almost algebraic by the assumption that
is almost algebraic. We recall that the centralizer
of a subset S is the set of elements y ∈
such that [sy] = 0 for all s ∈ S. This is a subalgebra of
. We shall now use the foregoing criterion to prove
THEOREM 18. Let be a completely reducible Lie algebra of linear transformations in a finite-dimensional vector space
of characteristic zero and let
1 be a completely reducible subalgebra of
. Then the centralizer
is completely reducible.
Proof. Let X ∈ 2. Then since
is almost algebraic, the semisimple and nilpotent parts Y, Z of X are in
. since these are polynomials in X, [CX] = 0 for C ∈
1 implies [CY] = 0 = [CZ]. Hence Y, Z ∈
2 and
2 is almost algebraic. We shall show next that
where
3 is a subspace of
such that
. It will then follow from Theorem 17 and Lemma 8 that every nil-potent element of
2 can be imbedded in a three-dimensional split simple algebra. Then
2 will be completely reducible by Theorem 17. Now we know that ad
1 is completely reducible (proof of Theorem 17). since
is a submodule of
relative to
1,
is completely reducible relative to ad
1. Thus we may write
where [i
1]
i, i = 1, …, k, and
i is irreducible relative to ad
1. We assume the
i are ordered so that [
i
1] = 0, i = 1, …, h, and [
i
1] ≠ 0 if j > h. since the subset
i of elements zi such that [zi
1] = 0 is a submodule of
i, it is immediate that
Set 3 =
h+1 + … +
k. Then
=
2 ⊕
3. If i > h, then [
i
1] ≠ 0 and [
i
1] + [[
i
1]
1] + … is an
1-submodule ≠ 0 of
i. Hence
i = [
i
1] + [[
i
1]
1] + …. This implies that
On the other hand, =
2 ⊕
3; hence, [
1] = [
3
1]
3. Hence
3 = [
1] and
This shows that 3 is a complement of
2 in
such that [
3
2]
3, which is what we needed to prove.
In all these exercises the characteristic of the base field will be zero, and unless the number is indicated with an asterisk the dimensionalities of the spaces will be finite.
1. Show that if is a Cartan subalgebra of
then
is a maximal nil-potent subalgebra of
. Show that the converse is false for ФnL (n
2).
2. Let be a nilpotent Lie algebra of linear transformations in
and let
=
0 ⊕
1 be the Fitting decomposition relative to
. Show that if Ф is infinite, then there exists an A ∈
such that
0 =
0A,
1 =
1A,
iA, the Fitting components relative to A.
3. Show that the diagonal matrices of trace 0 form a Cartan subalgebra in the Lie algebra of triangular matrices of trace 0. Show that
is complete
4. Let be the subalgebra of Ф2lL of matrices A satisfying S–lA′S = – A where
(This is isomorphic to an orthogonal Lie algebra.) Show that the diagonal matrices
form a Cartan subalgebra of .
5. Same as Exercise 4 but with S replaced by
6. Generalize Exercise 2.9 to the following: Let be a Lie algebra,
a nilpotent subalgebra of the derivation algebra of
. Suppose the only element l ∈
such that lD = 0 for all D ∈
is l = 0. Then prove that
is nilpotent.
7. Show that if si is a semi-simple ideal in 1 then
=
1 ⊕
2 where
2 is a second ideal.
8. Let be an ideal in
such that
/
is semi-simple. Show that there exists a subalgebra
1 of
such that
=
⊕
1.
9. Let be simple over an algebraically closed field Ф and let f(a, b) be an invariant symmetric bilinear form on
. Show that f is a multiple of the Killing form. Generalize this to semi-simple
.
10. Let n be the n-dimensional irreducible module for a split three-dimensional simple algebra
. Obtain a decomposition of
n ⊗
r into irreducible submodules.
11*. Let e, h be elements of an associative algebra such that [[eh]h] = 0. Show that if h is algebraic, that is, there exists a non-zero polynomial φ(λ) such that φ(h) = 0, then [eh] is nilpotent.
12*. Let be an associative algebra with an identity element 1 and suppose
contains elements e, f, h such that [eh] = 2e, [fh] = —2f, [ef] = h. Show that if φ(h) ∈
is a polynomial in h then
Also prove that if r and n are positive integers, r
n then
where
13*. , h, e, f as in Exercise 12. Show that if em = 0, then
14. Prove that if e is an element of a semi-simple Lie algebra of characteristic zero such that ad e is nilpotent, then eR is nilpotent for every representation R of
.
15. Prove that if is semi-simple over an algebraically closed field then
contains an e ≠ 0 with ad e nilpotent.
16. Prove that every finite-dimensional Lie algebra ≠ 0 over an algebraically closed field has indecomposable modules of arbitrarily high finite dimensionalities. (Hint: Show that there exists an e ∈ and a representation R such that eR is nilpotent ≠ 0. If
is the corresponding module, then the dimensionalities of the indecomposable components of
,
⊗
,
⊗
⊗
, … are not bounded.)
17. Prove that any semi-simple algebra has irreducible modules of arbitrarily high dimensionalities.
18. Show that the derivation algebra of any Lie algebra is algebraic. (Hint: Use Exercise 2.8.)
19. A Lie algebra is called reductive if ad
is completely reducible. Show that
is reductive if and only if
has a 1: 1 completely reducible representation.
20. A subalgebra of
is called reductive in
if ad
is completely reducible. Prove that if
is a completely reducible Lie algebra of linear transformations and
is reductive in
then
is completely reducible.
21. Show that any reductive commutative subalgebra of a semi-simple Lie algebra can be imbedded in a Cartan subalgebra.
22. Show that any semi-simple Lie algebra contains commutative Cartan subalgebras.
23. Let A be an automorphism of a semi-simple Lie algebra . Show that the subalgebra of elements y such that y (A — l)m = 0 for some m is a reductive subalgebra. (Hint: Use Exercise 2.5.)
24. (Mostow-Taft). Let G be a finite group of automorphisms in a Lie algebra. Show that has a Levi factor which is invariant under G.
25. Let fa(λ) = det (λ1 — ad a), the characteristic polynomial of ad a in a Lie algebra , and let D be a derivation of
;. Show that if t is an indeterminate, then
(Hint. Use the fact if A is an automorphism and the fact that exp tD = 1 + tD + (t2D2/2!) + … is a well-defined automorphism in
P, P the field of power series in t with coefficients in Ф)
26. Write and let τi(a1, …, ai) be the linearized form of τi defined by
Show that τi(a1, …, ai) is a symmetric i-linear function and that