Engel's theorem

Theorem in Lie representation theory

In representation theory, a branch of mathematics, Engel's theorem states that a finite-dimensional Lie algebra g {\displaystyle {\mathfrak {g}}} is a nilpotent Lie algebra if and only if for each X g {\displaystyle X\in {\mathfrak {g}}} , the adjoint map

ad ( X ) : g g , {\displaystyle \operatorname {ad} (X)\colon {\mathfrak {g}}\to {\mathfrak {g}},}

given by ad ( X ) ( Y ) = [ X , Y ] {\displaystyle \operatorname {ad} (X)(Y)=[X,Y]} , is a nilpotent endomorphism on g {\displaystyle {\mathfrak {g}}} ; i.e., ad ( X ) k = 0 {\displaystyle \operatorname {ad} (X)^{k}=0} for some k.[1] It is a consequence of the theorem, also called Engel's theorem, which says that if a Lie algebra of matrices consists of nilpotent matrices, then the matrices can all be simultaneously brought to a strictly upper triangular form. Note that if we merely have a Lie algebra of matrices which is nilpotent as a Lie algebra, then this conclusion does not follow (i.e. the naïve replacement in Lie's theorem of "solvable" with "nilpotent", and "upper triangular" with "strictly upper triangular", is false; this already fails for the one-dimensional Lie subalgebra of scalar matrices).

The theorem is named after the mathematician Friedrich Engel, who sketched a proof of it in a letter to Wilhelm Killing dated 20 July 1890 (Hawkins 2000, p. 176). Engel's student K.A. Umlauf gave a complete proof in his 1891 dissertation, reprinted as (Umlauf 2010).

Statements

Let g l ( V ) {\displaystyle {\mathfrak {gl}}(V)} be the Lie algebra of the endomorphisms of a finite-dimensional vector space V and g g l ( V ) {\displaystyle {\mathfrak {g}}\subset {\mathfrak {gl}}(V)} a subalgebra. Then Engel's theorem states the following are equivalent:

  1. Each X g {\displaystyle X\in {\mathfrak {g}}} is a nilpotent endomorphism on V.
  2. There exists a flag V = V 0 V 1 V n = 0 , codim V i = i {\displaystyle V=V_{0}\supset V_{1}\supset \cdots \supset V_{n}=0,\,\operatorname {codim} V_{i}=i} such that g V i V i + 1 {\displaystyle {\mathfrak {g}}\cdot V_{i}\subset V_{i+1}} ; i.e., the elements of g {\displaystyle {\mathfrak {g}}} are simultaneously strictly upper-triangulizable.

Note that no assumption on the underlying base field is required.

We note that Statement 2. for various g {\displaystyle {\mathfrak {g}}} and V is equivalent to the statement

  • For each nonzero finite-dimensional vector space V and a subalgebra g g l ( V ) {\displaystyle {\mathfrak {g}}\subset {\mathfrak {gl}}(V)} , there exists a nonzero vector v in V such that X ( v ) = 0 {\displaystyle X(v)=0} for every X g . {\displaystyle X\in {\mathfrak {g}}.}

This is the form of the theorem proven in #Proof. (This statement is trivially equivalent to Statement 2 since it allows one to inductively construct a flag with the required property.)

In general, a Lie algebra g {\displaystyle {\mathfrak {g}}} is said to be nilpotent if the lower central series of it vanishes in a finite step; i.e., for C 0 g = g , C i g = [ g , C i 1 g ] {\displaystyle C^{0}{\mathfrak {g}}={\mathfrak {g}},C^{i}{\mathfrak {g}}=[{\mathfrak {g}},C^{i-1}{\mathfrak {g}}]} = (i+1)-th power of g {\displaystyle {\mathfrak {g}}} , there is some k such that C k g = 0 {\displaystyle C^{k}{\mathfrak {g}}=0} . Then Engel's theorem implies the following theorem (also called Engel's theorem): when g {\displaystyle {\mathfrak {g}}} has finite dimension,

  • g {\displaystyle {\mathfrak {g}}} is nilpotent if and only if ad ( X ) {\displaystyle \operatorname {ad} (X)} is nilpotent for each X g {\displaystyle X\in {\mathfrak {g}}} .

Indeed, if ad ( g ) {\displaystyle \operatorname {ad} ({\mathfrak {g}})} consists of nilpotent operators, then by 1. {\displaystyle \Leftrightarrow } 2. applied to the algebra ad ( g ) g l ( g ) {\displaystyle \operatorname {ad} ({\mathfrak {g}})\subset {\mathfrak {gl}}({\mathfrak {g}})} , there exists a flag g = g 0 g 1 g n = 0 {\displaystyle {\mathfrak {g}}={\mathfrak {g}}_{0}\supset {\mathfrak {g}}_{1}\supset \cdots \supset {\mathfrak {g}}_{n}=0} such that [ g , g i ] g i + 1 {\displaystyle [{\mathfrak {g}},{\mathfrak {g}}_{i}]\subset {\mathfrak {g}}_{i+1}} . Since C i g g i {\displaystyle C^{i}{\mathfrak {g}}\subset {\mathfrak {g}}_{i}} , this implies g {\displaystyle {\mathfrak {g}}} is nilpotent. (The converse follows straightforwardly from the definition.)

Proof

We prove the following form of the theorem:[2] if g g l ( V ) {\displaystyle {\mathfrak {g}}\subset {\mathfrak {gl}}(V)} is a Lie subalgebra such that every X g {\displaystyle X\in {\mathfrak {g}}} is a nilpotent endomorphism and if V has positive dimension, then there exists a nonzero vector v in V such that X ( v ) = 0 {\displaystyle X(v)=0} for each X in g {\displaystyle {\mathfrak {g}}} .

The proof is by induction on the dimension of g {\displaystyle {\mathfrak {g}}} and consists of a few steps. (Note the structure of the proof is very similar to that for Lie's theorem, which concerns a solvable algebra.) The basic case is trivial and we assume the dimension of g {\displaystyle {\mathfrak {g}}} is positive.

Step 1: Find an ideal h {\displaystyle {\mathfrak {h}}} of codimension one in g {\displaystyle {\mathfrak {g}}} .

This is the most difficult step. Let h {\displaystyle {\mathfrak {h}}} be a maximal (proper) subalgebra of g {\displaystyle {\mathfrak {g}}} , which exists by finite-dimensionality. We claim it is an ideal of codimension one. For each X h {\displaystyle X\in {\mathfrak {h}}} , it is easy to check that (1) ad ( X ) {\displaystyle \operatorname {ad} (X)} induces a linear endomorphism g / h g / h {\displaystyle {\mathfrak {g}}/{\mathfrak {h}}\to {\mathfrak {g}}/{\mathfrak {h}}} and (2) this induced map is nilpotent (in fact, ad ( X ) {\displaystyle \operatorname {ad} (X)} is nilpotent as X {\displaystyle X} is nilpotent; see Jordan decomposition in Lie algebras). Thus, by inductive hypothesis applied to the Lie subalgebra of g l ( g / h ) {\displaystyle {\mathfrak {gl}}({\mathfrak {g}}/{\mathfrak {h}})} generated by ad ( h ) {\displaystyle \operatorname {ad} ({\mathfrak {h}})} , there exists a nonzero vector v in g / h {\displaystyle {\mathfrak {g}}/{\mathfrak {h}}} such that ad ( X ) ( v ) = 0 {\displaystyle \operatorname {ad} (X)(v)=0} for each X h {\displaystyle X\in {\mathfrak {h}}} . That is to say, if v = [ Y ] {\displaystyle v=[Y]} for some Y in g {\displaystyle {\mathfrak {g}}} but not in h {\displaystyle {\mathfrak {h}}} , then [ X , Y ] = ad ( X ) ( Y ) h {\displaystyle [X,Y]=\operatorname {ad} (X)(Y)\in {\mathfrak {h}}} for every X h {\displaystyle X\in {\mathfrak {h}}} . But then the subspace h g {\displaystyle {\mathfrak {h}}'\subset {\mathfrak {g}}} spanned by h {\displaystyle {\mathfrak {h}}} and Y is a Lie subalgebra in which h {\displaystyle {\mathfrak {h}}} is an ideal of codimension one. Hence, by maximality, h = g {\displaystyle {\mathfrak {h}}'={\mathfrak {g}}} . This proves the claim.

Step 2: Let W = { v V | X ( v ) = 0 , X h } {\displaystyle W=\{v\in V|X(v)=0,X\in {\mathfrak {h}}\}} . Then g {\displaystyle {\mathfrak {g}}} stabilizes W; i.e., X ( v ) W {\displaystyle X(v)\in W} for each X g , v W {\displaystyle X\in {\mathfrak {g}},v\in W} .

Indeed, for Y {\displaystyle Y} in g {\displaystyle {\mathfrak {g}}} and X {\displaystyle X} in h {\displaystyle {\mathfrak {h}}} , we have: X ( Y ( v ) ) = Y ( X ( v ) ) + [ X , Y ] ( v ) = 0 {\displaystyle X(Y(v))=Y(X(v))+[X,Y](v)=0} since h {\displaystyle {\mathfrak {h}}} is an ideal and so [ X , Y ] h {\displaystyle [X,Y]\in {\mathfrak {h}}} . Thus, Y ( v ) {\displaystyle Y(v)} is in W.

Step 3: Finish up the proof by finding a nonzero vector that gets killed by g {\displaystyle {\mathfrak {g}}} .

Write g = h + L {\displaystyle {\mathfrak {g}}={\mathfrak {h}}+L} where L is a one-dimensional vector subspace. Let Y be a nonzero vector in L and v a nonzero vector in W. Now, Y {\displaystyle Y} is a nilpotent endomorphism (by hypothesis) and so Y k ( v ) 0 , Y k + 1 ( v ) = 0 {\displaystyle Y^{k}(v)\neq 0,Y^{k+1}(v)=0} for some k. Then Y k ( v ) {\displaystyle Y^{k}(v)} is a required vector as the vector lies in W by Step 2. {\displaystyle \square }

See also

  • Lie's theorem
  • Heisenberg group

Notes

Citations

  1. ^ Fulton & Harris 1991, Exercise 9.10..
  2. ^ Fulton & Harris 1991, Theorem 9.9..

Works cited

  • Erdmann, Karin; Wildon, Mark (2006). Introduction to Lie Algebras (1st ed.). Springer. ISBN 1-84628-040-0.
  • Fulton, William; Harris, Joe (1991). Representation theory. A first course. Graduate Texts in Mathematics, Readings in Mathematics. Vol. 129. New York: Springer-Verlag. doi:10.1007/978-1-4612-0979-9. ISBN 978-0-387-97495-8. MR 1153249. OCLC 246650103.
  • Hawkins, Thomas (2000), Emergence of the theory of Lie groups, Sources and Studies in the History of Mathematics and Physical Sciences, Berlin, New York: Springer-Verlag, ISBN 978-0-387-98963-1, MR 1771134
  • Hochschild, G. (1965). The Structure of Lie Groups. Holden Day.
  • Humphreys, J. (1972). Introduction to Lie Algebras and Representation Theory. Springer.
  • Umlauf, Karl Arthur (2010) [First published 1891], Über Die Zusammensetzung Der Endlichen Continuierlichen Transformationsgruppen, Insbesondre Der Gruppen Vom Range Null, Inaugural-Dissertation, Leipzig (in German), Nabu Press, ISBN 978-1-141-58889-3