Lie algebra cohomology

Similar documents
Lie Algebra Homology and Cohomology

Lie Algebra Cohomology

De Rham Cohomology. Smooth singular cochains. (Hatcher, 2.1)

Representations of algebraic groups and their Lie algebras Jens Carsten Jantzen Lecture III

9. The Lie group Lie algebra correspondence

Math 210B. Profinite group cohomology

On the singular elements of a semisimple Lie algebra and the generalized Amitsur-Levitski Theorem

MATH 101B: ALGEBRA II PART A: HOMOLOGICAL ALGEBRA 23

1 Categorical Background

TCC Homological Algebra: Assignment #3 (Solutions)

ON COSTELLO S CONSTRUCTION OF THE WITTEN GENUS: L SPACES AND DG-MANIFOLDS

Lecture 5 - Lie Algebra Cohomology II

A Leibniz Algebra Structure on the Second Tensor Power

Qualifying Exam Syllabus and Transcript

NOTES ON CHAIN COMPLEXES

Math 249B. Nilpotence of connected solvable groups

THE HARISH-CHANDRA ISOMORPHISM FOR sl(2, C)

Notes on p-divisible Groups

A Primer on Homological Algebra

Hom M (J P (V ), U) Hom M (U(δ), J P (V )),

LECTURE 4: REPRESENTATION THEORY OF SL 2 (F) AND sl 2 (F)

Derivations and differentials

LECTURE 3: REPRESENTATION THEORY OF SL 2 (C) AND sl 2 (C)

NOTES ON BASIC HOMOLOGICAL ALGEBRA 0 L M N 0

On the Universal Enveloping Algebra: Including the Poincaré-Birkhoff-Witt Theorem

The Structure of Compact Groups

Classification of discretely decomposable A q (λ) with respect to reductive symmetric pairs UNIVERSITY OF TOKYO

QUANTIZATION VIA DIFFERENTIAL OPERATORS ON STACKS

Thus we get. ρj. Nρj i = δ D(i),j.

Three Descriptions of the Cohomology of Bun G (X) (Lecture 4)

Math 210B. The bar resolution

A TALE OF TWO FUNCTORS. Marc Culler. 1. Hom and Tensor

Quillen cohomology and Hochschild cohomology

Notes on the definitions of group cohomology and homology.

Boolean Algebras, Boolean Rings and Stone s Representation Theorem

PBW for an inclusion of Lie algebras

Algebraic Geometry Spring 2009

Representations of semisimple Lie algebras

DEFORMATIONS OF ALGEBRAS IN NONCOMMUTATIVE ALGEBRAIC GEOMETRY EXERCISE SHEET 1

Algebraic Geometry Spring 2009

Draft: July 15, 2007 ORDINARY PARTS OF ADMISSIBLE REPRESENTATIONS OF p-adic REDUCTIVE GROUPS I. DEFINITION AND FIRST PROPERTIES

LECTURE 3: TENSORING WITH FINITE DIMENSIONAL MODULES IN CATEGORY O

LECTURE 4: SOERGEL S THEOREM AND SOERGEL BIMODULES

BLOCKS IN THE CATEGORY OF FINITE-DIMENSIONAL REPRESENTATIONS OF gl(m n)

Atiyah classes and homotopy algebras

Hochschild cohomology

REPRESENTATION THEORY WEEK 9

Lie algebra cohomology

Hungry, Hungry Homology

SEMISIMPLE LIE GROUPS

LECTURE 16: LIE GROUPS AND THEIR LIE ALGEBRAS. 1. Lie groups

3.1. Derivations. Let A be a commutative k-algebra. Let M be a left A-module. A derivation of A in M is a linear map D : A M such that

MATH 101B: ALGEBRA II PART A: HOMOLOGICAL ALGEBRA

Lecture 4 Super Lie groups

The Gauss-Manin Connection for the Cyclic Homology of Smooth Deformations, and Noncommutative Tori

Lemma 1.3. The element [X, X] is nonzero.

The Cartan Decomposition of a Complex Semisimple Lie Algebra

Cohomology and Base Change

LECTURE 3: RELATIVE SINGULAR HOMOLOGY

CHAPTER 1. AFFINE ALGEBRAIC VARIETIES

FILTERED RINGS AND MODULES. GRADINGS AND COMPLETIONS.

THE THEOREM OF THE HIGHEST WEIGHT

A GLIMPSE OF ALGEBRAIC K-THEORY: Eric M. Friedlander

Hodge Structures. October 8, A few examples of symmetric spaces

58 CHAPTER 2. COMPUTATIONAL METHODS

Variations on a Casselman-Osborne theme

Topics in Representation Theory: Roots and Weights

Conformal blocks for a chiral algebra as quasi-coherent sheaf on Bun G.

LOCAL VS GLOBAL DEFINITION OF THE FUSION TENSOR PRODUCT

The Gelfand-Tsetlin Basis (Too Many Direct Sums, and Also a Graph)

INTRO TO TENSOR PRODUCTS MATH 250B

SECTION 5: EILENBERG ZILBER EQUIVALENCES AND THE KÜNNETH THEOREMS

AHAHA: Preliminary results on p-adic groups and their representations.

The BRST complex for a group action

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

INTRODUCTION TO LIE ALGEBRAS. LECTURE 7.

Math Homotopy Theory Hurewicz theorem

Algebra Qualifying Exam Solutions January 18, 2008 Nick Gurski 0 A B C 0

ON DEGENERATION OF THE SPECTRAL SEQUENCE FOR THE COMPOSITION OF ZUCKERMAN FUNCTORS

Enveloping algebras of Hom-Lie algebras

Winter School on Galois Theory Luxembourg, February INTRODUCTION TO PROFINITE GROUPS Luis Ribes Carleton University, Ottawa, Canada

THE STEENROD ALGEBRA. The goal of these notes is to show how to use the Steenrod algebra and the Serre spectral sequence to calculate things.

Etale cohomology of fields by Johan M. Commelin, December 5, 2013

3. Categories and Functors We recall the definition of a category: Definition 3.1. A category C is the data of two collections. The first collection

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

Lecture 11 The Radical and Semisimple Lie Algebras

HOMOLOGY AND COHOMOLOGY. 1. Introduction

Formal power series rings, inverse limits, and I-adic completions of rings

GK-SEMINAR SS2015: SHEAF COHOMOLOGY

CATEGORY THEORY. Cats have been around for 70 years. Eilenberg + Mac Lane =. Cats are about building bridges between different parts of maths.

FILTRATIONS IN ABELIAN CATEGORIES WITH A TILTING OBJECT OF HOMOLOGICAL DIMENSION TWO

Tensor Product of modules. MA499 Project II

D-MODULES: AN INTRODUCTION

Categories and functors

10. Smooth Varieties. 82 Andreas Gathmann

Math 210C. A non-closed commutator subgroup

Category O and its basic properties

Manifolds and Poincaré duality

Exercises in Geometry II University of Bonn, Summer semester 2015 Professor: Prof. Christian Blohmann Assistant: Saskia Voss Sheet 1

HOPF ALGEBRAS AND LIE ALGEBRAS UCHICAGO PRO-SEMINAR - JANUARY 9, 2014

Transcription:

Lie algebra cohomology November 16, 2018 1 History Citing [1]: In mathematics, Lie algebra cohomology is a cohomology theory for Lie algebras. It was first introduced in 1929 by Élie Cartan to study the topology of Lie groups and homogeneous spaces by relating cohomological methods of Georges de Rham to properties of the Lie algebra. It was later extended by Claude Chevalley and Samuel Eilenberg (1948) to coefficients in an arbitrary Lie module. 2 Objects in Lie theory We begin and define Lie algebras, Lie groups and enveloping algebras. Their relation is given by: Lie algebras appear as... tangent spaces at 1 of Lie groups and by unenveloping associative algebras. We then continue to define representations (ie. modules) of such objects. 2.1 Lie algebras A Lie algebra over a field k 1 is a vector space g with a binary operation [, ] : g g g (called the Lie bracket) satisfying bilinearity, alternativity and the Jacobi identity. One should think of a Lie algebra as a (non-unitary) ring with [, ] being multiplication, although it need not be associative (or commutative). For example one defines a derivation on g as a linear map δ : g g satisfying Leibniz law: δ[x, y] = [δx, y] + [x, δy]. (1) 1 The definition for Lie algebra makes sense for any ring k. However, we will mainly discuss the case k {R, C}, namely when g is the Lie algebra for a Lie group G (real) or its complexification g C. 1

Also the notions of homomorphism, subalgebra, ideal, direct sum and quotient algebra extend to Lie algebras. For example a linear map f : g g between underlying vector spaces of g and g is a homomorphism of Lie algebras if it is compatible with multiplication: f[x, y] = [fx, fy]. If i is an ideal of g and g g/i splits (ie. has a section), then g = g/i i is called a semidirect product. Levi s theorem says that every finite dimensional Lie algebra is a semidirect product (by its radical and the complementary (Levi) subalgebra). The dimension of g is the cardinality of a minimal generating set of g 2. 2.2 Lie groups By a (real) Lie group G we mean a group object in the category of finite dimensional real smooth manifolds 3. 2.3 Enveloping algebras Let A be an associative algebra with multiplication. The underyling vector space of A can be equipped with a Lie bracket [x, y] = x y y x 4 resulting in a Lie algebra Lie(A). The algebra A is called an enveloping algebra of Lie(A). This construction is functorial and turns Lie into a (right adjoint) functor Lie : {k-associative algebras} {k-lie algebras} : U. (2) The left adjoint U assigns to g its universal enveloping algebra U(g). U(g) can be constructed as the quotient T (g)/i of T (g) = k g g 2 g 3... (the tensor algebra of g) by I the two-sided ideal in T (g) generated by elements of the form x y y x [x, y] g g 2. In particular, if g is abelian, then U(g) = Sym(g) is the symmetric algebra of the vector space underlying g. Example 2.1. Let M be a vector space. Then End(M) is an associative algebra and we denote gl(m) := Lie(End(M)). Also, if k = R, then gl(m) = T 1 GL(M). Theorem 2.2. (The PBW 5 -theorem) The natural map g LieU(g) is injective. 2 As usual the subalgebra h g generated by a set S g is the smallest subalgebra containing S. 3 Multiplication and inversion being smooth is equivalent to (x, y) x 1 y being smooth. 4 The Jacobi identity follows from associativity of. 5 Named after Henri Poincaré, Garret Birkhoff and Ernst Witt. 2

3 Representations Always g denotes a Lie algebra over a field (or ring) k. If we speak about G, it is assumed to be a real Lie group with Lie algebra g and in particular k = R and g possibly complexified. By f we mean a Lie subalgebra of g. By K we mean a subgroup of G, mostly assumed compact and maximal. We do simplify notation and always write π for the representation, whether from g or G or... If M is an k-vector space we can associate to it a Lie algebra gl(m) (over k), an associative algebra End(M) (over k) and if k {R, C} a Lie group GL(M). Then basically, morphisms to gl(m) (resp. End(M), resp. GL(M)) represent representations on M. 3.1 g-modules A g-module is a pair (M, π) where M is an k-vector space and π : g gl(m) in the category of Lie algebras over k. 3.2 U g-modules A Ug-module is an k-vector space equipped with an action π : Ug End(M) in the category of associative algebras. One has a one-to-one correspondence between g-modules and U g-modules: use gl(m) = Lie(End(M)) and (U, Lie) is an adjoint pair to get Hom(g, gl(m)) = Hom(U(g), End(M)). (3) Remark 3.1. Given any associative algebra A, an A-module M is a morphism π : A End(M) in the category of k-associative algebras. Applying Lie then makes M a Lie(A)-module: 3.3 Topological G-modules Lie(A) gl(m) (4) A topological G-module or simply G-module is a C vector space M (locally convex Hausdorff) equipped with a continuous action π : G Aut(M). Let C G be the category of topological G-modules and equivariant continuous linear maps. 3.4 Smooth G-modules Let M be a topological G-module. A vector v M is smooth if it is fixed by an open subgroup of G. The set of smooth vectors in M defines a G-submodule M M. A smooth G-module is a topological G-module M satisfying 1. M = M, 2. v (g gv) : M C (G, M) is a homeomorphism onto its image. 3

The full subcategory of smooth G-modules (in C G ) is denoted C G. The assignment M M is functorial and gives C G C G. If M is smooth we can differentiate and obtain (dπ) 1 : g gl(m). This makes M a g-module. Conversely, if G is simply connected, then there is an inverse functor associating to a g-module M a smooth G-module. Thus, there is a one-to-one correspondence between smooth G-modules and g-modules. 3.5 Adjoint representations As an example we discuss the adjoint representation. Every Lie algebra g is a module over itself via the adjoint representation ad, which is simply g acting on g by left-multiplication: ad : g gl(g) (5) x [x, ]. (6) Indeed this is a Lie algebra homomorphism (note that the Lie bracket for gl(g) is given by the commutator [x, y] = xy yx): [ad x, ad y ] = ad x ad y ad y ad x (7) = [x, [y, ]] [y, [x, ]] (8) = [[y, ], x] [[, x], y] (9) = [[x, y], ] (10) = ad [x,y] (11) If k {R, C} and G a Lie group with Lie algebra g = Lie(G), then this g- module structure on g lifts 6 to a G-module structure. This G-module is also called the adjoint representation, denoted by Ad : G GL(g) and defined by Ad(x) = (dc x ) 1, where c x : y xyx 1 is conjugation by x on G. 3.6 (g, K)-modules Let M be a G-module. Then M also has induced the structure of a g-module and in the case of simply connected G one can reconstruct the G-action on M from its induced g-action. We want to simplify the situation by restricting M to an even smaller subspace, but also remembering the action of a subgroup K < G. We say v M is K-finite if it is contained in a finite dimensional K-submodule. The space of K-finite vectors is denoted V (K). For K = G and V (K) = M, we call M locally finite. We are mostly interested in compact subgroups K. Then V (K) is a semi-simple K-module. Moreover, if K is maximal compact, then V 0 := M V (K) 7 also comes with an g-action and both actions (the one of K and the one of g) are compatible: for k K, X g and v V 0 we have k(xv) = (kxk 1 )kv = Ad(k)(X)kv. (12) 6 By this we mean g = T 1 (G) and ad = (dad) 1. 7 If M is smooth this extra step of filtering the smooth vectors is unnecessary. 4

This module V 0 is an example of what is called a (g, K)-module: A (g, K)-module is a real or complex vector space M that is both: a g-module and a locally finite and semi-simple K-module, satisfying the compatibility conditions: 1. kxv = (kxk 1 )kv = Ad(k)(X)kv for all k K, X Ug and v M, 2. if U M is a finite dimensional K-submodule, then the representation π of K on U is differentiable and its differential (dπ) 1 coincides with the action π of g restricted to f = T 1 K: (dπ) 1 = π f. (13) The category of (g, K)-modules and (g, K)-morphisms is denoted C g,k. Remark 3.2. One can think of V 0 equipped with the actions of K and g as an infinitesimal approximation to the G-module M. Indeed one defines smooth G-modules M, M to be infinitesimally equivalent if their associated (g, K)- modules are isomorphic. 4 Cohomology The following motivation is from [1]: If G is a compact simply connected Lie group, then it is determined by its Lie algebra, so it should be possible to calculate its cohomology from the Lie algebra. This can be done as follows. Its cohomology is the de Rham cohomology of the complex of differential forms on G. Using an averaging process, this complex can be replaced by the complex of left-invariant differential forms. The left-invariant forms, meanwhile, are determined by their values at the identity, so that the space of left-invariant differential forms can be identified with the exterior algebra of the Lie algebra, with a suitable differential. The construction of this differential on an exterior algebra makes sense for any Lie algebra, so is used to define Lie algebra cohomology for all Lie algebras. More generally one uses a similar construction to define Lie algebra cohomology with coefficients in a module. It should be noted that if G is a simply connected noncompact Lie group, the Lie algebra cohomology of the associated Lie algebra g does not necessarily reproduce the de Rham cohomology of G. The reason for this is that the passage from the complex of all differential forms to the complex of left-invariant differential forms uses an averaging process that only makes sense for compact groups. 4.1 (Co)homology of Lie algebras Let g be a Lie algebra over a field k and M be a left g-module. As the usual story goes we try to derive the following functors: 5

the invariant submodule functor M M g = {m M X g : Xm = 0}, The coinvariant quotient module functor M M g = M/gM. We then define homology and cohomology groups of g with coefficients in M by By definition: H n (g, M) = L n ( g)(m) (14) H n (g, M) = R n ( g)(m). (15) H 0 (g, M) = M g (16) H 0 (g, M) = M g. (17) Actually one has to make sure that there are enough injectives and enough projectives for those definitions to work. The case with projectives is clear: take free resolutions. Recall that Ug is the quotient of T g by the ideal generated by elements of the form i([x, y]) i(x)i(y) + i(y)i(x) with i : g T g. The k-algebra homomorphism ɛ : Ug k sending i(g) to zero is called augmentation and its kernel J augmentation ideal 0 J Ug k 0. (18) Proposition 4.1. Let k be equipped with the trivial g-module structure. We can compute Lie algebra cohomology as U g-cohomology: H n (g, M) = T orn Ug (k, M) (19) H n (g, M) = Ext n Ug(k, M). (20) Proof. Derived functors are isomorphic if their underlying functors are: M g = M/gM = M/JM = Ug/J Ug M = k Ug M (21) M g = Hom g (k, M) = Hom Ug (k, M). (22) Proposition 4.2. Let M, N be left g-modules. Then Hom k (M, N) can be made a g-module via Xf(m) = Xf(m) f(xm) for X g, m M. With this one has: Ext n Ug(M, N) = H n (g, Hom k (M, N)). (23) Proof. This follows from Hom g (M, N) = Hom k (M, N) g and the fact that for a field k, there are no nontrivial extensions of k-modules. 6

4.2 Homology of Lie algebras in degree one To understand the homology in low degrees we look at the long exact sequence for and T orn Ug (, M) (M a g-module): 0 J Ug k 0 (24) T or Ug 1 (J, M) T orug 1 (Ug, M) T orug 1 (k, M) (25) J Ug M Ug Ug M k Ug M 0. (26) Since Ug is free as Ug-module one has T orn Ug (Ug, M) = 0 for n 1. Hence, for n 2 and H n (g, M) = T orn Ug (k, M) = T or Ug n 1 (J, M) (27) 0 H 1 (g, M) J Ug M M M g 0. (28) Finally, tensoring 24 with J Ug and using that i : g Ug maps [g, g] to J 2, inducing g ab = g/[g, g] = J/J 2 yields J Ug k = g ab. (29) Theorem 4.3. If M is a trivial g-module, then H 1 (g, M) = g ab k M. Proof. This follows from 28 and M g = M: H 1 (g, M) = J Ug M = (J Ug k) k M = J/J 2 k M = g ab k M. (30) 4.3 Cohomology of Lie algebras in degree one Theorem 4.4. The first cohomology of a g-module M is the quotient of derivations Der(g, M) modulo inner derivation Der in (g, M): H 1 (g, M) = Der(g, M)/Der in (g, M). (31) A proof is easy after we introduce the Chevalley-Eilenberg complex, later. Recall, that a derivation D is a k-linear map D : g M, satisfying the Leibniz law (where g is considered a g-module via its adjoint representation: x y = [x, y] ): D([x, y]) = xdy ydx. Der(g, M) is the k-submodule of Hom k (g, M) of derivations. The subspace Der in (g, M) of inner derivations is defined as the image of M Der(g, M) (32) m D m : X Xm. (33) 7

4.4 The Chevalley-Eilenberg complex In this subsection we want to construct a complex C (g) satisfying: Theorem 4.5. The Chevalley-Eilenberg complex C (g) = Ug k p g is a projective resolution of the g-module k. In particular, if M is a g-module we can compute its (co)homology via: H n (g, M) = H n (M Ug C (g)), (34) H n (g, M) = H n (Hom g (C (g), M)). (35) Proof. Construction: First note that every C n (g) is actually a free Ug-module. We define the differentials as follows: For low degrees we use C 1 (g) = Ug k g d C 0 (g) = Ug k k = Ug ɛ k 0 (36) with d(u X) = ux and ɛ being augmentation. Note that im(d) = J, hence so far the sequence is exact. For higher degrees n 2 we define where θ 1 = d : C n (g) C n 1 (g) (37) u X 1 X n θ 1 + θ 2, (38) n ( 1) i+1 ux i X 1 ˆX i X n (39) i=1 θ 2 = i<j( 1) i+j u [X i, X j ] X 1 ˆX i ˆX j X n. (40) For example, if n = 2, then d(u X Y ) = ux Y uy X u [X, Y ]. Corollary 4.6. If g is n-dimensional over k, then (co)homology in degree > n must vanish. Proof. Then, n g = 0. Remark 4.7. We have M Ug C (g) = M Ug Ug k g = M k g (41) Hom g (C (g), M) = Hom g (Ug k g, M) = Hom k ( g, M) (42) and in the latter an n-cochain is simply an alternating k-multilinear function f(x 1,..., X n ) on g n with values in M. Its coboundary is df(x 1,..., X n+1 ) = ( 1) i X i f(x 1,..., ˆX i,..., X n+1 ) (43) + ( 1) i+j f([x i, X j ], X 1,..., ˆX i,..., ˆX j,..., X n+1 ). (44) 8

Example 4.8. Check that Z 1 (g, M) = Der(g, M) (45) B 1 (g, M) = Der in (g, M) (46) and hence H 1 (g, M) = Der(g, M)/Der in (g, M) as claimed. 4.5 Cohomology of Lie algebras in degrees two and three Theorem 4.9. Let M be a g-module. Then H 2 (g, M) is in bijection with the set Ext(g, M) of equivalence classes of extensions of g by M 8. Similar, H 3 (g, M) is in bijection with 2-extensions, ie. four term short exact sequences of Lie algebras 0 M h f g 0. (47) Proof. I think this is a standard argument from homological algebra. The keyword is: Yoneda extension. 5 Relative cohomology 5.1 Cohomology of (g, f)-modules As a motivation for relative cohomology there is a result of E. Cartan, stating that for G compact connected, one has H (g, f; M) = H (G/K, M), (48) where the left side denotes relative cohomology of g modulo f, where f = Lie(K). As usual it would be nice to have two characterizations: one abstract as a cohomology of a derived functor, the other concrete in terms of cohomology of a complex. Indeed, in terms of (Ug, Uf)-algebras one has Ext q Ug,Uf (k, M) = Hq (g, f; M) (49) and for the cochain complex version we define a subcomplex of the Chevalley- Eilenberg complex by C q (g, f; M) = Hom f ( q (g/f), M) (50) with the f-action on q (g/f) being induced from its adjoint action on f. In other words, it is the subspace of Hom k ( q (g/f), M) of elements f satisfying (for X f, X i g/f, i = 1,..., q): f(x 1,..., [X, X i ],..., X q ) = Xf(X 1,..., X q ). (51) i The cohomology groups H (g, f; M) of C (g, f; M) are called relative cohomology groups of g mod f with coefficients in M. 8 Actually, Ext(g, M) also admits an addition given by the Baer sum and the bijection stated then becomes an isomorphism. 9

5.2 Cohomology of (g, K)-modules Another relative cohomology considers the case with (g, K)-modules M. For this let K be a maximal compact subgroup of the real Lie group G 9 and let k = R. We only give a cochain computing cohomology: Let C q (g, K; M) = Hom K ( q (g/f), M), (52) where this time K acts on g/f via the adjoint representation. Note the relation to (g, f)-module cohomology (K 0 is the identity component of K) C q (g, K; M) C q (g, K 0 ; M) = C q (g, f; M). (53) Also K/K 0 naturally acts on C q (g, f; M) and we have C q (g, K; M) = C q (g, f; M) K/K0. (54) C (g, K; M) is a subcomplex of C (g, f; M) and it follows from 54 that: H q (g, K; M) = H q (g, f; M) K/K0 (55) 6 Application: Weyl s theorem Goal of this section is Theorem 6.1 (Weyl s theorem). Let g be a semisimple Lie algebra over a field k of characteristic 0. Then every finite dimensional g-module is semisimple, ie. a direct sum of simple g-modules. Proof. Sketch!: Step 1: first proof a vanishing theorem: If M is simple and k, then for all i H i (g, M) = H i (g, M) = 0. (56) The proof for this uses Schur s lemma and the so called Casimir operator. Step 2: Assume the opposite, ie. that M is not semisimple, ie. a direct sum of simple modules. Then, since M is finite dimensional, there is a submodule M 1 minimal with the property not being semisimple. Clearly, M 1 is not simple, so it contains a proper submodule M 0. By minimality, M 0 and M 2 = M 1 /M 0 are semisimple, while M 1 is not. This means, that 0 M 0 M 1 M 2 0 (57) is not split and hence induces a nontrivial element in Ext 1 Ug(M 2, M 0 ) = H 1 (g, Hom k (M 2, M 0 )). (58) 9 G is assumed to have a finite component group. 10

Step 3: We are left with showing that such an element does not exist. This follows from Whitehead s first lemma 10 : let g, k be as in our setting and M be a finite dimensional g-module. Then Use Hom k (M 2, M 0 ) for M to reach the contradiction. References H 1 (g, M) = 0. (59) [1] Wikipedia authors, https://en.wikipedia.org/wiki/lie_algebra_ cohomology [2] Shen-Ning Tung, Lie Algebra Homology and Cohomology, 2013 [3] A. Borel, N. Wallach, Continuous Cohomology, Discrete Subgroups, and Representations of Reductive Groups, Princeton, N.J.: Princeton University Press, 1980. 10 Here is a proof: If M is simple, then either M = k and H 1 (g, k) = g/[g, g] = 0 or else M k and H (g, M) = 0 by the vanishing theorem. Otherwise, M contains a proper submodule L and by induction H 1 (g, L) = H 1 (g, M/L) = 0. Since 0 L M M/L 0 induces an exact sequence of cohomology: H 1 (g, L) H 1 (g, M) H 1 (g, L/M)..., we have H 1 (g, M) = 0. 11