Lie's theorem

In mathematics, specifically the theory of Lie algebras, Lie's theorem states that,{{harvnb|Serre|2001|loc=Theorem 3}} over an algebraically closed field of characteristic zero, if \pi: \mathfrak{g} \to \mathfrak{gl}(V) is a finite-dimensional representation of a solvable Lie algebra, then there is a flag V = V_0 \supset V_1 \supset \cdots \supset V_n = 0 of invariant subspaces of \pi(\mathfrak{g}) with \operatorname{codim} V_i = i, meaning that \pi(X)(V_i) \subseteq V_i for each X \in \mathfrak{g} and i.

Put in another way, the theorem says there is a basis for V such that all linear transformations in \pi(\mathfrak{g}) are represented by upper triangular matrices.{{harvnb|Humphreys|1972|loc=Ch. II, § 4.1., Corollary A.}} This is a generalization of the result of Frobenius that commuting matrices are simultaneously upper triangularizable, as commuting matrices generate an abelian Lie algebra, which is a fortiori solvable.

A consequence of Lie's theorem is that any finite dimensional solvable Lie algebra over a field of characteristic 0 has a nilpotent derived algebra (see #Consequences). Also, to each flag in a finite-dimensional vector space V, there correspond a Borel subalgebra (that consist of linear transformations stabilizing the flag); thus, the theorem says that \pi(\mathfrak{g}) is contained in some Borel subalgebra of \mathfrak{gl}(V).

Counter-example

For algebraically closed fields of characteristic p>0 Lie's theorem holds provided the dimension of the representation is less than p (see the proof below), but can fail for representations of dimension p. An example is given by the 3-dimensional nilpotent Lie algebra spanned by 1, x, and d/dx acting on the p-dimensional vector space k[x]/(xp), which has no eigenvectors. Taking the semidirect product of this 3-dimensional Lie algebra by the p-dimensional representation (considered as an abelian Lie algebra) gives a solvable Lie algebra whose derived algebra is not nilpotent.

Proof

The proof is by induction on the dimension of \mathfrak{g} and consists of several steps. (Note: the structure of the proof is very similar to that for Engel's theorem.) The basic case is trivial and we assume the dimension of \mathfrak{g} is positive. We also assume V is not zero. For simplicity, we write X \cdot v = \pi(X)(v).

Step 1: Observe that the theorem is equivalent to the statement:{{harvnb|Serre|2001|loc=Theorem 3{{''}}}}

  • There exists a vector in V that is an eigenvector for each linear transformation in \pi(\mathfrak{g}).

Indeed, the theorem says in particular that a nonzero vector spanning V_{n-1} is a common eigenvector for all the linear transformations in \pi(\mathfrak{g}). Conversely, if v is a common eigenvector, take V_{n-1} to be its span and then \pi(\mathfrak{g}) admits a common eigenvector in the quotient V/V_{n-1}; repeat the argument.

Step 2: Find an ideal \mathfrak{h} of codimension one in \mathfrak{g}.

Let D\mathfrak{g} = [\mathfrak{g}, \mathfrak{g}] be the derived algebra. Since \mathfrak{g} is solvable and has positive dimension, D\mathfrak{g} \ne \mathfrak{g} and so the quotient \mathfrak{g}/D\mathfrak{g} is a nonzero abelian Lie algebra, which certainly contains an ideal of codimension one and by the ideal correspondence, it corresponds to an ideal of codimension one in \mathfrak{g}.

Step 3: There exists some linear functional \lambda in \mathfrak{h}^* such that

:V_{\lambda} = \{ v \in V | X \cdot v = \lambda(X) v, X \in \mathfrak{h} \}

is nonzero. This follows from the inductive hypothesis (it is easy to check that the eigenvalues determine a linear functional).

Step 4: V_{\lambda} is a \mathfrak{g}-invariant subspace. (Note this step proves a general fact and does not involve solvability.)

Let Y \in \mathfrak{g}, v \in V_{\lambda}, then we need to prove Y \cdot v \in V_{\lambda}. If v = 0 then it's obvious, so assume v \ne 0 and set recursively v_0 = v, \, v_{i+1} = Y \cdot v_i. Let U = \operatorname{span} \{ v_i | i \ge 0 \} and \ell \in \mathbb{N}_0 be the largest such that v_0,\ldots,v_\ell are linearly independent. Then we'll prove that they generate U and thus \alpha = (v_0,\ldots,v_\ell) is a basis of U. Indeed, assume by contradiction that it's not the case and let m \in \mathbb{N}_0 be the smallest such that v_m \notin \langle v_0,\ldots,v_\ell\rangle, then obviously m \ge \ell + 1. Since v_0,\ldots,v_{\ell+1} are linearly dependent, v_{\ell+1} is a linear combination of v_0,\ldots,v_\ell. Applying the map Y^{m-\ell-1} it follows that v_m is a linear combination of v_{m-\ell-1},\ldots,v_{m-1}. Since by the minimality of m each of these vectors is a linear combination of v_0,\ldots,v_\ell, so is v_m, and we get the desired contradiction. We'll prove by induction that for every n \in \mathbb{N}_0 and X \in \mathfrak{h} there exist elements a_{0,n,X},\ldots,a_{n,n,X} of the base field such that a_{n,n,X}=\lambda(X) and

:X \cdot v_n = \sum_{i=0}^{n} a_{i,n,X}v_i.

The n=0 case is straightforward since X \cdot v_0 = \lambda(X) v_0. Now assume that we have proved the claim for some n \in \mathbb{N}_0 and all elements of \mathfrak{h} and let X \in \mathfrak{h}. Since \mathfrak{h} is an ideal, it's [X,Y] \in \mathfrak{h}, and thus

:X \cdot v_{n+1} = Y \cdot (X \cdot v_n) + [X, Y] \cdot v_n = Y \cdot \sum_{i=0}^{n} a_{i,n,X}v_i + \sum_{i=0}^{n} a_{i,n,[X,Y]}v_i = a_{0,n,[X,Y]}v_0 + \sum_{i=1}^{n} (a_{i-1,n,X} + a_{i,n,[X,Y]})v_i + \lambda(X)v_{n+1},

and the induction step follows. This implies that for every X \in \mathfrak{h} the subspace U is an invariant subspace of X and the matrix of the restricted map \pi(X)|_U in the basis \alpha is upper triangular with diagonal elements equal to \lambda(X), hence \operatorname{tr}(\pi(X)|_U) = \dim(U) \lambda(X). Applying this with [X,Y] \in \mathfrak{h} instead of X gives \operatorname{tr}(\pi([X,Y])|_U) = \dim(U) \lambda([X,Y]). On the other hand, U is also obviously an invariant subspace of Y, and so

:\operatorname{tr}(\pi([X,Y])|_U) = \operatorname{tr}([\pi(X),\pi(Y)]|_U]) = \operatorname{tr}([\pi(X)|_U, \pi(Y)|_U]) = 0

since commutators have zero trace, and thus \dim(U) \lambda([X,Y]) = 0. Since \dim(U) > 0 is invertible (because of the assumption on the characteristic of the base field), \lambda([X, Y]) = 0 and

:X \cdot (Y \cdot v) = Y \cdot (X \cdot v) + [X, Y] \cdot v = Y \cdot (\lambda(X) v) + \lambda([X, Y]) v = \lambda(X) (Y \cdot v),

and so Y \cdot v \in V_{\lambda}.

Step 5: Finish up the proof by finding a common eigenvector.

Write \mathfrak{g} = \mathfrak{h} + L where L is a one-dimensional vector subspace. Since the base field is algebraically closed, there exists an eigenvector in V_{\lambda} for some (thus every) nonzero element of L. Since that vector is also eigenvector for each element of \mathfrak{h}, the proof is complete. \square

Consequences

The theorem applies in particular to the adjoint representation \operatorname{ad}: \mathfrak{g} \to \mathfrak{gl}(\mathfrak{g}) of a (finite-dimensional) solvable Lie algebra \mathfrak{g} over an algebraically closed field of characteristic zero; thus, one can choose a basis on \mathfrak{g} with respect to which \operatorname{ad}(\mathfrak{g}) consists of upper triangular matrices. It follows easily that for each x, y \in \mathfrak{g}, \operatorname{ad}([x, y]) = [\operatorname{ad}(x), \operatorname{ad}(y)] has diagonal consisting of zeros; i.e., \operatorname{ad}([x, y]) is a strictly upper triangular matrix. This implies that [\mathfrak g, \mathfrak g] is a nilpotent Lie algebra. Moreover, if the base field is not algebraically closed then solvability and nilpotency of a Lie algebra is unaffected by extending the base field to its algebraic closure. Hence, one concludes the statement (the other implication is obvious):{{harvnb|Humphreys|1972|loc=Ch. II, § 4.1., Corollary C.}}

:A finite-dimensional Lie algebra \mathfrak g over a field of characteristic zero is solvable if and only if the derived algebra D \mathfrak g = [\mathfrak g, \mathfrak g] is nilpotent.

Lie's theorem also establishes one direction in Cartan's criterion for solvability:

:If V is a finite-dimensional vector space over a field of characteristic zero and \mathfrak{g} \subseteq \mathfrak{gl}(V) a Lie subalgebra, then \mathfrak{g} is solvable if and only if \operatorname{tr}(XY) = 0 for every X \in \mathfrak{g} and Y \in [\mathfrak{g}, \mathfrak{g}].{{harvnb|Serre|2001|loc=Theorem 4}}

Indeed, as above, after extending the base field, the implication \Rightarrow is seen easily. (The converse is more difficult to prove.)

Lie's theorem (for various V) is equivalent to the statement:{{harvnb|Serre|2001|loc=Theorem 3'}}

:For a solvable Lie algebra \mathfrak g over an algebraically closed field of characteristic zero, each finite-dimensional simple \mathfrak{g}-module (i.e., irreducible as a representation) has dimension one.

Indeed, Lie's theorem clearly implies this statement. Conversely, assume the statement is true. Given a finite-dimensional \mathfrak g-module V, let V_1 be a maximal \mathfrak g-submodule (which exists by finiteness of the dimension). Then, by maximality, V/V_1 is simple; thus, is one-dimensional. The induction now finishes the proof.

The statement says in particular that a finite-dimensional simple module over an abelian Lie algebra is one-dimensional; this fact remains true over any base field since in this case every vector subspace is a Lie subalgebra.{{harvnb|Jacobson|1979|loc=Ch. II, § 6, Lemma 5.}}

Here is another quite useful application:{{harvnb|Fulton|Harris|1991|loc=Proposition 9.17.}}

:''Let \mathfrak{g} be a finite-dimensional Lie algebra over an algebraically closed field of characteristic zero with radical \operatorname{rad}(\mathfrak{g}). Then each finite-dimensional simple representation \pi: \mathfrak{g} \to \mathfrak{gl}(V) is the tensor product of a simple representation of \mathfrak{g}/\operatorname{rad}(\mathfrak{g}) with a one-dimensional representation of \mathfrak{g} (i.e., a linear functional vanishing on Lie brackets).

By Lie's theorem, we can find a linear functional \lambda of \operatorname{rad}(\mathfrak{g}) so that there is the weight space V_{\lambda} of \operatorname{rad}(\mathfrak{g}). By Step 4 of the proof of Lie's theorem, V_{\lambda} is also a \mathfrak{g}-module; so V = V_{\lambda}. In particular, for each X \in \operatorname{rad}(\mathfrak{g}), \operatorname{tr}(\pi(X)) = \dim(V) \lambda(X). Extend \lambda to a linear functional on \mathfrak{g} that vanishes on [\mathfrak g, \mathfrak g]; \lambda is then a one-dimensional representation of \mathfrak{g}. Now, (\pi, V) \simeq (\pi, V) \otimes (-\lambda) \otimes \lambda. Since \pi coincides with \lambda on \operatorname{rad}(\mathfrak{g}), we have that V \otimes (-\lambda) is trivial on \operatorname{rad}(\mathfrak{g}) and thus is the restriction of a (simple) representation of \mathfrak{g}/\operatorname{rad}(\mathfrak{g}). \square

See also

References

{{reflist}}

Sources

  • {{Fulton-Harris}}
  • {{Citation | last1=Humphreys | first1=James E. | title=Introduction to Lie Algebras and Representation Theory | publisher=Springer-Verlag | location=Berlin, New York | isbn=978-0-387-90053-7 | year=1972 | url-access=registration | url=https://archive.org/details/introductiontoli00jame }}.
  • {{Citation |last = Jacobson |first = Nathan| author-link = Nathan Jacobson| title = Lie algebras| edition = Republication of the 1962 original|publisher = Dover Publications, Inc.|location = New York|year = 1979|mr = 0559927|isbn = 0-486-63832-4}}
  • {{Citation |first = Jean-Pierre |last = Serre|title = Complex Semisimple Lie Algebras| series=Springer Monographs in Mathematics | publisher = Springer|location = Berlin|year = 2001|isbn = 3-5406-7827-1|doi = 10.1007/978-3-642-56884-8|mr = 1808366}}

Category:Theorems about algebras