Weyl's theorem on complete reducibility

In algebra, Weyl's theorem on complete reducibility is a fundamental result in the theory of Lie algebra representations (specifically in the representation theory of semisimple Lie algebras). Let g {\displaystyle {\mathfrak {g}}} be a semisimple Lie algebra over a field of characteristic zero. The theorem states that every finite-dimensional module over g {\displaystyle {\mathfrak {g}}} is semisimple as a module (i.e., a direct sum of simple modules.)[1]

The enveloping algebra is semisimple

Weyl's theorem implies (in fact is equivalent to) that the enveloping algebra of a finite-dimensional representation is a semisimple ring in the following way.

Given a finite-dimensional Lie algebra representation π : g g l ( V ) {\displaystyle \pi :{\mathfrak {g}}\to {\mathfrak {gl}}(V)} , let A End ( V ) {\displaystyle A\subset \operatorname {End} (V)} be the associative subalgebra of the endomorphism algebra of V generated by π ( g ) {\displaystyle \pi ({\mathfrak {g}})} . The ring A is called the enveloping algebra of π {\displaystyle \pi } . If π {\displaystyle \pi } is semisimple, then A is semisimple.[2] (Proof: Since A is a finite-dimensional algebra, it is an Artinian ring; in particular, the Jacobson radical J is nilpotent. If V is simple, then J V V {\displaystyle JV\subset V} implies that J V = 0 {\displaystyle JV=0} . In general, J kills each simple submodule of V; in particular, J kills V and so J is zero.) Conversely, if A is semisimple, then V is a semisimple A-module; i.e., semisimple as a g {\displaystyle {\mathfrak {g}}} -module. (Note that a module over a semisimple ring is semisimple since a module is a quotient of a free module and "semisimple" is preserved under the free and quotient constructions.)

Application: preservation of Jordan decomposition

Here is a typical application.[3]

Proposition — Let g {\displaystyle {\mathfrak {g}}} be a semisimple finite-dimensional Lie algebra over a field of characteristic zero.[a]

  1. There exists a unique pair of elements x s , x n {\displaystyle x_{s},x_{n}} in g {\displaystyle {\mathfrak {g}}} such that x = x s + x n {\displaystyle x=x_{s}+x_{n}} , ad ( x s ) {\displaystyle \operatorname {ad} (x_{s})} is semisimple, ad ( x n ) {\displaystyle \operatorname {ad} (x_{n})} is nilpotent and [ x s , x n ] = 0 {\displaystyle [x_{s},x_{n}]=0} .
  2. If π : g g l ( V ) {\displaystyle \pi :{\mathfrak {g}}\to {\mathfrak {gl}}(V)} is a finite-dimensional representation, then π ( x ) s = π ( x s ) {\displaystyle \pi (x)_{s}=\pi (x_{s})} and π ( x ) n = π ( x n ) {\displaystyle \pi (x)_{n}=\pi (x_{n})} , where π ( x ) s , π ( x ) n {\displaystyle \pi (x)_{s},\pi (x)_{n}} denote the Jordan decomposition of the semisimple and nilpotent parts of the endomorphism π ( x ) {\displaystyle \pi (x)} .

In short, the semisimple and nilpotent parts of an element of g {\displaystyle {\mathfrak {g}}} are well-defined and are determined independent of a faithful finite-dimensional representation.

Proof: First we prove the special case of (i) and (ii) when π {\displaystyle \pi } is the inclusion; i.e., g {\displaystyle {\mathfrak {g}}} is a subalgebra of g l n = g l ( V ) {\displaystyle {\mathfrak {gl}}_{n}={\mathfrak {gl}}(V)} . Let x = S + N {\displaystyle x=S+N} be the Jordan decomposition of the endomorphism x {\displaystyle x} , where S , N {\displaystyle S,N} are semisimple and nilpotent endomorphisms in g l n {\displaystyle {\mathfrak {gl}}_{n}} . Now, ad g l n ( x ) {\displaystyle \operatorname {ad} _{{\mathfrak {gl}}_{n}}(x)} also has the Jordan decomposition, which can be shown (see Jordan–Chevalley decomposition) to respect the above Jordan decomposition; i.e., ad g l n ( S ) , ad g l n ( N ) {\displaystyle \operatorname {ad} _{{\mathfrak {gl}}_{n}}(S),\operatorname {ad} _{{\mathfrak {gl}}_{n}}(N)} are the semisimple and nilpotent parts of ad g l n ( x ) {\displaystyle \operatorname {ad} _{{\mathfrak {gl}}_{n}}(x)} . Since ad g l n ( S ) , ad g l n ( N ) {\displaystyle \operatorname {ad} _{{\mathfrak {gl}}_{n}}(S),\operatorname {ad} _{{\mathfrak {gl}}_{n}}(N)} are polynomials in ad g l n ( x ) {\displaystyle \operatorname {ad} _{{\mathfrak {gl}}_{n}}(x)} then, we see ad g l n ( S ) , ad g l n ( N ) : g g {\displaystyle \operatorname {ad} _{{\mathfrak {gl}}_{n}}(S),\operatorname {ad} _{{\mathfrak {gl}}_{n}}(N):{\mathfrak {g}}\to {\mathfrak {g}}} . Thus, they are derivations of g {\displaystyle {\mathfrak {g}}} . Since g {\displaystyle {\mathfrak {g}}} is semisimple, we can find elements s , n {\displaystyle s,n} in g {\displaystyle {\mathfrak {g}}} such that [ y , S ] = [ y , s ] , y g {\displaystyle [y,S]=[y,s],y\in {\mathfrak {g}}} and similarly for n {\displaystyle n} . Now, let A be the enveloping algebra of g {\displaystyle {\mathfrak {g}}} ; i.e., the subalgebra of the endomorphism algebra of V generated by g {\displaystyle {\mathfrak {g}}} . As noted above, A has zero Jacobson radical. Since [ y , N n ] = 0 {\displaystyle [y,N-n]=0} , we see that N n {\displaystyle N-n} is a nilpotent element in the center of A. But, in general, a central nilpotent belongs to the Jacobson radical; hence, N = n {\displaystyle N=n} and thus also S = s {\displaystyle S=s} . This proves the special case.

In general, π ( x ) {\displaystyle \pi (x)} is semisimple (resp. nilpotent) when ad ( x ) {\displaystyle \operatorname {ad} (x)} is semisimple (resp. nilpotent).[clarification needed] This immediately gives (i) and (ii). {\displaystyle \square }

Proofs

Analytic proof

Weyl's original proof (for complex semisimple Lie algebras) was analytic in nature: it famously used the unitarian trick. Specifically, one can show that every complex semisimple Lie algebra g {\displaystyle {\mathfrak {g}}} is the complexification of the Lie algebra of a simply connected compact Lie group K {\displaystyle K} .[4] (If, for example, g = s l ( n ; C ) {\displaystyle {\mathfrak {g}}=\mathrm {sl} (n;\mathbb {C} )} , then K = S U ( n ) {\displaystyle K=\mathrm {SU} (n)} .) Given a representation π {\displaystyle \pi } of g {\displaystyle {\mathfrak {g}}} on a vector space V , {\displaystyle V,} one can first restrict π {\displaystyle \pi } to the Lie algebra k {\displaystyle {\mathfrak {k}}} of K {\displaystyle K} . Then, since K {\displaystyle K} is simply connected,[5] there is an associated representation Π {\displaystyle \Pi } of K {\displaystyle K} . Integration over K {\displaystyle K} produces an inner product on V {\displaystyle V} for which Π {\displaystyle \Pi } is unitary.[6] Complete reducibility of Π {\displaystyle \Pi } is then immediate and elementary arguments show that the original representation π {\displaystyle \pi } of g {\displaystyle {\mathfrak {g}}} is also completely reducible.

Algebraic proof 1

Let ( π , V ) {\displaystyle (\pi ,V)} be a finite-dimensional representation of a Lie algebra g {\displaystyle {\mathfrak {g}}} over a field of characteristic zero. The theorem is an easy consequence of Whitehead's lemma, which says V Der ( g , V ) , v v {\displaystyle V\to \operatorname {Der} ({\mathfrak {g}},V),v\mapsto \cdot v} is surjective, where a linear map f : g V {\displaystyle f:{\mathfrak {g}}\to V} is a derivation if f ( [ x , y ] ) = x f ( y ) y f ( x ) {\displaystyle f([x,y])=x\cdot f(y)-y\cdot f(x)} . The proof is essentially due to Whitehead.[7]

Let W V {\displaystyle W\subset V} be a subrepresentation. Consider the vector subspace L W End ( V ) {\displaystyle L_{W}\subset \operatorname {End} (V)} that consists of all linear maps t : V V {\displaystyle t:V\to V} such that t ( V ) W {\displaystyle t(V)\subset W} and t ( W ) = 0 {\displaystyle t(W)=0} . It has a structure of a g {\displaystyle {\mathfrak {g}}} -module given by: for x g , t L W {\displaystyle x\in {\mathfrak {g}},t\in L_{W}} ,

x t = [ π ( x ) , t ] {\displaystyle x\cdot t=[\pi (x),t]} .

Now, pick some projection p : V V {\displaystyle p:V\to V} onto W and consider f : g L W {\displaystyle f:{\mathfrak {g}}\to L_{W}} given by f ( x ) = [ p , π ( x ) ] {\displaystyle f(x)=[p,\pi (x)]} . Since f {\displaystyle f} is a derivation, by Whitehead's lemma, we can write f ( x ) = x t {\displaystyle f(x)=x\cdot t} for some t L W {\displaystyle t\in L_{W}} . We then have [ π ( x ) , p + t ] = 0 , x g {\displaystyle [\pi (x),p+t]=0,x\in {\mathfrak {g}}} ; that is to say p + t {\displaystyle p+t} is g {\displaystyle {\mathfrak {g}}} -linear. Also, as t kills W {\displaystyle W} , p + t {\displaystyle p+t} is an idempotent such that ( p + t ) ( V ) = W {\displaystyle (p+t)(V)=W} . The kernel of p + t {\displaystyle p+t} is then a complementary representation to W {\displaystyle W} . {\displaystyle \square }

See also Weibel's homological algebra book.

Algebraic proof 2

Whitehead's lemma is typically proved by means of the quadratic Casimir element of the universal enveloping algebra,[8] and there is also a proof of the theorem that uses the Casimir element directly instead of Whitehead's lemma.

Since the quadratic Casimir element C {\displaystyle C} is in the center of the universal enveloping algebra, Schur's lemma tells us that C {\displaystyle C} acts as multiple c λ {\displaystyle c_{\lambda }} of the identity in the irreducible representation of g {\displaystyle {\mathfrak {g}}} with highest weight λ {\displaystyle \lambda } . A key point is to establish that c λ {\displaystyle c_{\lambda }} is nonzero whenever the representation is nontrivial. This can be done by a general argument [9] or by the explicit formula for c λ {\displaystyle c_{\lambda }} .

Consider a very special case of the theorem on complete reducibility: the case where a representation V {\displaystyle V} contains a nontrivial, irreducible, invariant subspace W {\displaystyle W} of codimension one. Let C V {\displaystyle C_{V}} denote the action of C {\displaystyle C} on V {\displaystyle V} . Since V {\displaystyle V} is not irreducible, C V {\displaystyle C_{V}} is not necessarily a multiple of the identity, but it is a self-intertwining operator for V {\displaystyle V} . Then the restriction of C V {\displaystyle C_{V}} to W {\displaystyle W} is a nonzero multiple of the identity. But since the quotient V / W {\displaystyle V/W} is a one dimensional—and therefore trivial—representation of g {\displaystyle {\mathfrak {g}}} , the action of C {\displaystyle C} on the quotient is trivial. It then easily follows that C V {\displaystyle C_{V}} must have a nonzero kernel—and the kernel is an invariant subspace, since C V {\displaystyle C_{V}} is a self-intertwiner. The kernel is then a one-dimensional invariant subspace, whose intersection with W {\displaystyle W} is zero. Thus, k e r ( V C ) {\displaystyle \mathrm {ker} (V_{C})} is an invariant complement to W {\displaystyle W} , so that V {\displaystyle V} decomposes as a direct sum of irreducible subspaces:

V = W k e r ( C V ) {\displaystyle V=W\oplus \mathrm {ker} (C_{V})} .

Although this establishes only a very special case of the desired result, this step is actually the critical one in the general argument.

Algebraic proof 3

The theorem can be deduced from the theory of Verma modules, which characterizes a simple module as a quotient of a Verma module by a maximal submodule.[10] This approach has an advantage that it can be used to weaken the finite-dimensionality assumptions (on algebra and representation).

Let V {\displaystyle V} be a finite-dimensional representation of a finite-dimensional semisimple Lie algebra g {\displaystyle {\mathfrak {g}}} over an algebraically closed field of characteristic zero. Let b = h n + g {\displaystyle {\mathfrak {b}}={\mathfrak {h}}\oplus {\mathfrak {n}}_{+}\subset {\mathfrak {g}}} be the Borel subalgebra determined by a choice of a Cartan subalgebra and positive roots. Let V 0 = { v V | n + ( v ) = 0 } {\displaystyle V^{0}=\{v\in V|{\mathfrak {n}}_{+}(v)=0\}} . Then V 0 {\displaystyle V^{0}} is an h {\displaystyle {\mathfrak {h}}} -module and thus has the h {\displaystyle {\mathfrak {h}}} -weight space decomposition:

V 0 = λ L V λ 0 {\displaystyle V^{0}=\bigoplus _{\lambda \in L}V_{\lambda }^{0}}

where L h {\displaystyle L\subset {\mathfrak {h}}^{*}} . For each λ L {\displaystyle \lambda \in L} , pick 0 v λ V λ {\displaystyle 0\neq v_{\lambda }\in V_{\lambda }} and V λ V {\displaystyle V^{\lambda }\subset V} the g {\displaystyle {\mathfrak {g}}} -submodule generated by v λ {\displaystyle v_{\lambda }} and V V {\displaystyle V'\subset V} the g {\displaystyle {\mathfrak {g}}} -submodule generated by V 0 {\displaystyle V^{0}} . We claim: V = V {\displaystyle V=V'} . Suppose V V {\displaystyle V\neq V'} . By Lie's theorem, there exists a b {\displaystyle {\mathfrak {b}}} -weight vector in V / V {\displaystyle V/V'} ; thus, we can find an h {\displaystyle {\mathfrak {h}}} -weight vector v {\displaystyle v} such that 0 e i ( v ) V {\displaystyle 0\neq e_{i}(v)\in V'} for some e i {\displaystyle e_{i}} among the Chevalley generators. Now, e i ( v ) {\displaystyle e_{i}(v)} has weight μ + α i {\displaystyle \mu +\alpha _{i}} . Since L {\displaystyle L} is partially ordered, there is a λ L {\displaystyle \lambda \in L} such that λ μ + α i {\displaystyle \lambda \geq \mu +\alpha _{i}} ; i.e., λ > μ {\displaystyle \lambda >\mu } . But this is a contradiction since λ , μ {\displaystyle \lambda ,\mu } are both primitive weights (it is known that the primitive weights are incomparable.[clarification needed]). Similarly, each V λ {\displaystyle V^{\lambda }} is simple as a g {\displaystyle {\mathfrak {g}}} -module. Indeed, if it is not simple, then, for some μ < λ {\displaystyle \mu <\lambda } , V μ 0 {\displaystyle V_{\mu }^{0}} contains some nonzero vector that is not a highest-weight vector; again a contradiction.[clarification needed] {\displaystyle \square }

External links

  • A blog post by Akhil Mathew

References

  1. ^ Editorial note: this fact is usually stated for a field of characteristic zero, but the proof needs only that the base field be perfect.
  1. ^ Hall 2015 Theorem 10.9
  2. ^ Jacobson 1979, Ch. II, § 5, Theorem 10.
  3. ^ Jacobson 1979, Ch. III, § 11, Theorem 17.
  4. ^ Knapp 2002 Theorem 6.11
  5. ^ Hall 2015 Theorem 5.10
  6. ^ Hall 2015 Theorem 4.28
  7. ^ Jacobson 1979, Ch. III, § 7.
  8. ^ Hall 2015 Section 10.3
  9. ^ Humphreys 1973 Section 6.2
  10. ^ Kac 1990, Lemma 9.5.
  • Hall, Brian C. (2015). Lie Groups, Lie Algebras, and Representations: An Elementary Introduction. Graduate Texts in Mathematics. Vol. 222 (2nd ed.). Springer. ISBN 978-3319134666.
  • Humphreys, James E. (1973). Introduction to Lie Algebras and Representation Theory. Graduate Texts in Mathematics. Vol. 9 (Second printing, revised ed.). New York: Springer-Verlag. ISBN 0-387-90053-5.
  • Jacobson, Nathan (1979). Lie algebras. New York: Dover Publications, Inc. ISBN 0-486-63832-4. Republication of the 1962 original.
  • Kac, Victor (1990). Infinite dimensional Lie algebras (3rd ed.). Cambridge University Press. ISBN 0-521-46693-8.
  • Knapp, Anthony W. (2002), Lie Groups Beyond an Introduction, Progress in Mathematics, vol. 140 (2nd ed.), Boston: Birkhäuser, ISBN 0-8176-4259-5
  • Weibel, Charles A. (1995). An Introduction to Homological Algebra. Cambridge University Press.