lie-algebras-and-their-representations
Source code for my notes on representations of semisimple Lie algebras and Olivier Mathieu's classification of simple weight modules
complete-reducibility.tex (47064B)
1 \chapter{Semisimplicity \& Complete Reducibility} 2 3 \label{start-47} 4 5 Having hopefully established in the previous chapter that Lie algebras and 6 their representations are indeed useful, we are now faced with the Herculean 7 task of trying to understand them. We have seen that representations can be 8 used to derive geometric information about groups, but the question remains: 9 how do we go about classifying the representations of a given Lie algebra? This 10 question has sparked an entire field of research, and we cannot hope to provide 11 a comprehensive answer in the \pagedifference{start-47}{end-47} pages we have 12 left. Nevertheless, we can work on particular cases. 13 14 For instance, one can readily check that a \(K^n\)-module \(M\) -- here \(K^n\) 15 denotes the \(n\)-dimensional Abelian Lie algebra -- is nothing more than a 16 choice of \(n\) commuting operators \(M \to M\) -- corresponding to the action 17 of the canonical basis elements \(e_1, \ldots, e_n \in K^n\). In particular, a 18 \(1\)-dimensional \(K^n\)-module is just a choice of \(n\) scalars \(\lambda_1, 19 \ldots, \lambda_n\). Different choices of scalars yield non-isomorphic modules, 20 so that the \(1\)-dimensional \(K^n\)-modules are parameterized by points in 21 \(K^n\). 22 23 This goes to show that classifying the representations of Abelian algebras is 24 not that interesting of a problem. Instead, we focus on a less trivial, yet 25 reasonably well behaved case: the finite-dimensional modules of a 26 finite-dimensional semisimple Lie algebra \(\mathfrak{g}\) over an 27 algebraically closed field \(K\) of characteristic \(0\). But why are the 28 modules of a semisimple Lie algebras simpler -- or perhaps \emph{semisimpler} 29 -- to understand than those of any old Lie algebra? We will get back to this 30 question in a moment, but for now we simply note that, when solving a 31 classification problem, it is often profitable to break down our structure is 32 smaller pieces. This leads us to the following definitions. 33 34 \begin{definition}\index{\(\mathfrak{g}\)-module!indecomposable module} 35 A \(\mathfrak{g}\)-module is called \emph{indecomposable} if it is 36 not isomorphic to the direct sum of two nonzero \(\mathfrak{g}\)-modules. 37 \end{definition} 38 39 \begin{definition}\index{\(\mathfrak{g}\)-module!simple module}\index{simple!\(\mathfrak{g}\)-module} 40 A \(\mathfrak{g}\)-module is called \emph{simple} if it has no nonzero proper 41 \(\mathfrak{g}\)-modules. 42 \end{definition} 43 44 \begin{example} 45 The trivial \(\mathfrak{g}\)-module \(K\) is an example of a simple 46 \(\mathfrak{g}\)-module. In fact, every \(1\)-dimensional 47 \(\mathfrak{g}\)-module \(M\) is simple: \(M\) has no nonzero proper 48 \(K\)-subspaces, let alone \(\mathfrak{g}\)-submodules. 49 \end{example} 50 51 \begin{example}\label{ex:all-simple-reps-are-tensor-prod} 52 Given a finite-dimensional simple \(\mathfrak{g}_1\)-module \(M_1\) and a 53 finite-dimensional simple \(\mathfrak{g}_2\)-module \(M_2\), the tensor 54 product \(M_1 \otimes M_2\) is a simple \(\mathfrak{g}_1 \oplus 55 \mathfrak{g}_2\)-module. All finite-dimensional simple \(\mathfrak{g}_1 56 \oplus \mathfrak{g}_2\)-modules have the form \(M_1 \otimes M_2\) for unique 57 (up to isomorphism) \(M_1\) and \(M_2\). In light of 58 Example~\ref{ex:univ-enveloping-of-sum-is-tensor}, this is a particular case 59 of the fact that, given \(K\)-algebras \(A\) and \(B\), all 60 finite-dimensional simple \(A \otimes_K B\)-modules are given tensor products 61 of simple \(A\)-modules with simple \(B\)-modules -- see 62 \cite[ch.~3]{etingof}. 63 \end{example} 64 65 The general strategy for classifying finite-dimensional modules over an algebra 66 is to classify the indecomposable modules. This is because\dots 67 68 \begin{theorem}[Krull-Schmidt]\label{thm:krull-schmidt} 69 Let \(\mathfrak{g}\) be a Lie algebra. 70 Then every finite-dimensional \(\mathfrak{g}\)-module can be uniquely -- 71 up to isomorphisms and reordering of the summands -- decomposed into a direct 72 sum of indecomposable \(\mathfrak{g}\)-modules. 73 \end{theorem} 74 75 Hence finding the indecomposable \(\mathfrak{g}\)-modules suffices to find 76 \emph{all} finite-dimensional \(\mathfrak{g}\)-modules: they are the direct sum 77 of indecomposable \(\mathfrak{g}\)-modules. The existence of the decomposition 78 should be clear from the definitions. Indeed, if \(M\) is a finite-dimensional 79 \(\mathfrak{g}\)-modules a simple argument via induction in \(\dim M\) suffices 80 to prove the existence: if \(M\) is indecomposable then there is nothing to 81 prove, and if \(M\) is not indecomposable then \(M = N \oplus L\) for some 82 nonzero submodules \(N, L \subsetneq M\), so that their dimensions are both 83 strictly smaller than \(\dim M\) and the existence follows from the induction 84 hypothesis. For a proof of uniqueness please refer to \cite{etingof}. 85 86 Finding the indecomposable modules of an arbitrary Lie algebra, however, turns 87 out to be a bit of a circular problem: the indecomposable 88 \(\mathfrak{g}\)-modules are the ones that cannot be decomposed, which is to 89 say, those that are \emph{not} decomposable. Ideally, we would like to find 90 some other condition, equivalent to indecomposability, but which is easier to 91 work with. It is clear from the definitions that every simple 92 \(\mathfrak{g}\)-module is indecomposable, but there is no reason to believe 93 the converse is true. Indeed, this is not always the case. For instance\dots 94 95 \begin{example}\label{ex:indecomposable-not-irr} 96 The space \(M = K^2\) endowed with the action 97 \begin{align*} 98 x \cdot e_1 & = e_1 & x \cdot e_2 = e_1 + e_2 99 \end{align*} 100 of the Lie algebra \(K[x]\) is a \(K[x]\)-module. Notice \(M\) has a single 101 nonzero proper submodule, which is spanned by the vector \(e_1\). This is 102 because if \((a + b) e_1 + b e_2 = x \cdot (a e_1 + b e_2) = \lambda \cdot (a 103 e_1 + b e_2)\) for some \(\lambda \in K\) then \(\lambda = 1\) and \(b = 0\). 104 Hence \(M\) is indecomposable -- it cannot be broken into a direct sum of 105 \(1\)-dimensional submodules -- but it is evidently not simple. 106 \end{example} 107 108 This counterexample poses an interesting question: are there conditions one can 109 impose on an algebra \(\mathfrak{g}\) under which every indecomposable 110 \(\mathfrak{g}\)-module is simple? This is what is known in representation 111 theory as \emph{complete reducibility}. 112 113 \begin{definition}\index{\(\mathfrak{g}\)-module!completely reducible module} 114 A \(\mathfrak{g}\)-module \(M\) is called \emph{completely reducible} if 115 every \(\mathfrak{g}\)-submodule of \(M\) has a \(\mathfrak{g}\)-invariant 116 complement -- i.e. given \(N \subset M\), there is a submodule \(L \subset 117 M\) such that \(M = N \oplus L\). 118 \end{definition} 119 120 \begin{definition}\index{\(\mathfrak{g}\)-module!semisimple module}\index{semisimple!\(\mathfrak{g}\)-module} 121 A \(\mathfrak{g}\)-module \(M\) is called \emph{semisimple} if it is the 122 direct sum of simple \(\mathfrak{g}\)-modules. 123 \end{definition} 124 125 In case the relationship between complete reducibility, semisimplicity of 126 \(\mathfrak{g}\)-modules and the simplicity of indecomposable modules is 127 unclear, the following results should clear things up. 128 129 \begin{proposition} 130 The following conditions are equivalent. 131 \begin{enumerate} 132 \item Every submodule of a finite-dimensional \(\mathfrak{g}\)-module is 133 completely reducible. 134 135 \item Every exact sequence of finite-dimensional \(\mathfrak{g}\)-modules 136 splits. 137 138 \item Every indecomposable finite-dimensional \(\mathfrak{g}\)-module is 139 simple. 140 141 \item Every finite-dimensional \(\mathfrak{g}\)-module is semisimple. 142 \end{enumerate} 143 \end{proposition} 144 145 \begin{proof} 146 We begin by \(\textbf{(i)} \implies \textbf{(ii)}\). Let 147 \begin{center} 148 \begin{tikzcd} 149 0 \rar & 150 N \rar{f} & 151 M \rar{g} & 152 L \rar & 153 0 154 \end{tikzcd} 155 \end{center} 156 be an exact sequence of \(\mathfrak{g}\)-modules. We can suppose without loss 157 of generality that \(N \subset M\) is a submodule and \(f\) is its inclusion 158 in \(M\), for if this is not the case there is an isomorphism of sequences 159 \begin{center} 160 \begin{tikzcd} 161 0 \rar & 162 N \rar{f} \dar[swap]{f} & 163 M \rar{g} \dar[Rightarrow, no head] & 164 L \rar \dar[Rightarrow, no head] & 165 0 \\ 166 0 \rar & 167 f(N) \rar & 168 M \rar[swap]{g} & 169 L \rar & 170 0 171 \end{tikzcd} 172 \end{center} 173 174 It then follows from \textbf{(i)} that there exists a 175 \(\mathfrak{g}\)-submodule \(L' \subset M\) such that \(M = N \oplus L'\). 176 Finally, the projection \(s : M \to N\) is \(\mathfrak{g}\)-homomorphism 177 satisfying 178 \begin{center} 179 \begin{tikzcd} 180 0 \rar & 181 N \rar{f} & 182 M \rar{g} \lar[bend left=30]{s} & 183 L \rar & 184 0 185 \end{tikzcd} 186 \end{center} 187 188 Next is \(\textbf{(ii)} \implies \textbf{(iii)}\). If \(M\) is an 189 indecomposable \(\mathfrak{g}\)-module and \(N \subset M\) is a submodule, we 190 have an exact sequence 191 \begin{center} 192 \begin{tikzcd} 193 0 \rar & 194 N \rar & 195 M \rar & 196 \mfrac{M}{N} \rar & 197 0 198 \end{tikzcd} 199 \end{center} 200 of \(\mathfrak{g}\)-modules. 201 202 Since our sequence splits, we must have \(M \cong N \oplus \mfrac{M}{N}\). 203 But \(M\) is indecomposable, so that either \(M = N\) or \(M \cong 204 \mfrac{M}{N}\), in which case \(N = 0\). Since this holds for all \(N \subset 205 M\), \(M\) is simple. For \(\textbf{(iii)} \implies \textbf{(iv)}\) it 206 suffices to apply Theorem~\ref{thm:krull-schmidt}. 207 208 Finally, for \(\textbf{(iv)} \implies \textbf{(i)}\), if we assume 209 \(\textbf{(iv)}\) and let \(M\) be a \(\mathfrak{g}\)-module with 210 decomposition into simple submodules 211 \[ 212 M = \bigoplus_i M_i 213 \] 214 and \(N \subset M\) is a submodule. Take some maximal set of indexes \(\{i_1, 215 \ldots, i_r\}\) so that \(\left( \bigoplus_k M_{i_k} \right) \cap M = 0\) and 216 let \(L = \bigoplus_k M_{i_k}\). We want to establish \(M = N \oplus L\). 217 218 Suppose without any loss in generality that \(i_k = k\) for all \(k\) and let 219 \(j > r\). By the maximality of our set of indexes, there is some nonzero \(n 220 \in (M_j \oplus L) \cap N\). Say \(n = m_j + m_1 + \cdots + m_r\) with each 221 \(m_i \in M_i\). Then \(m_j = n - m_1 - \cdots - m_r \in M_j \cap (N \oplus 222 L)\) is nonzero. Indeed, if this is not the case we find \(0 \ne n = m_1 + 223 \cdots + m_r \in \left( \bigoplus_{i = 1}^r M_i \right) \cap N\), a 224 contradiction. This implies \(M_j \cap (N \oplus L)\) is a nonzero submodule 225 of \(M_j\). Since \(M_j\) is simple, \(M_j = M_j \cap (N \oplus L)\) and 226 therefore \(M_j \subset N \oplus L\). Given the arbitrary choice of \(j\), it 227 then follows \(M = N \oplus L\). 228 \end{proof} 229 230 While we are primarily interested in indecomposable \(\mathfrak{g}\)-modules -- 231 which is usually a strictly larger class of representations than that of simple 232 \(\mathfrak{g}\)-modules -- it is important to note that simple 233 \(\mathfrak{g}\)-modules are generally much easier to find. The relationship 234 between simple \(\mathfrak{g}\)-modules is also well understood. This is 235 because of the following result, known as \emph{Schur's Lemma}. 236 237 \begin{lemma}[Schur] 238 Let \(M\) and \(N\) be simple \(\mathfrak{g}\)-modules and \(f : M \to N\) be 239 a \(\mathfrak{g}\)-homomorphism. Then \(f\) is either \(0\) or an 240 isomorphism. Furthermore, if \(M = N\) is finite-dimensional then \(f\) is a 241 scalar operator. 242 \end{lemma} 243 244 \begin{proof} 245 For the first statement, it suffices to notice that \(\ker f\) and 246 \(\operatorname{im} f\) are both submodules. In particular, either \(\ker f = 247 0\) and \(\operatorname{im} f = N\) or \(\ker f = M\) and \(\operatorname{im} 248 f = 0\). Now suppose \(M = N\) is finite-dimensional. Let \(\lambda \in K\) 249 be an eigenvalue of \(f\) -- which exists because \(K\) is algebraically 250 closed -- and \(M_\lambda\) be its corresponding eigenspace. Given \(m \in 251 M_\lambda\), \(f(X \cdot m) = X \cdot f(m) = \lambda X \cdot m\). In other 252 words, \(M_\lambda\) is a \(\mathfrak{g}\)-submodule. It then follows 253 \(M_\lambda = M\), given that \(M_\lambda \ne 0\). 254 \end{proof} 255 256 We are now ready to answer our first question: the special thing about 257 semisimple algebras is that the relationship between their indecomposable 258 modules and their simple modules is much clearer. Namely\dots 259 260 \begin{proposition} 261 Given a finite-dimensional Lie algebra \(\mathfrak{g}\) over \(K\), 262 \(\mathfrak{g}\) is semisimple if, and only if every finite-dimensional 263 \(\mathfrak{g}\)-module is completely reducible. 264 \end{proposition} 265 266 The proof of the fact that a finite-dimensional Lie algebra \(\mathfrak{g}\) 267 whose finite-dimensional modules are completely reducible is semisimple is 268 actually pretty simple. Namely, it suffices to note that the adjoint 269 \(\mathfrak{g}\)-module is the direct sum of simple submodules, which are all 270 simple ideals of \(\mathfrak{g}\) -- so \(\mathfrak{g}\) is the direct sum of 271 simple Lie algebras. The proof of the converse is more nuanced, and this will 272 be our next milestone. 273 274 Before proceeding to the proof of complete reducibility, however, we would like 275 to introduce some basic tools which will come in handy later on, known as\dots 276 277 \section{Invariant Bilinear Forms} 278 279 \begin{definition}\index{invariant bilinear form} 280 A symmetric bilinear form \(B : \mathfrak{g} \times \mathfrak{g} \to K\) is 281 called \emph{\(\mathfrak{g}\)-invariant} if the operator 282 \(\operatorname{ad}(X) : \mathfrak{g} \to \mathfrak{g}\) is antisymmetric 283 with respect to \(B\) for all \(X \in \mathfrak{g}\). 284 \[ 285 B(\operatorname{ad}(X) Y, Z) + B(Y, \operatorname{ad}(X) Z) = 0 286 \] 287 \end{definition} 288 289 \begin{note} 290 The etymology of the term \emph{invariant form} comes from group 291 representation theory. Namely, given a linear action of a group \(G\) on a 292 vector space \(V\) equipped with a bilinear form \(B\), \(B\) is called 293 \(G\)-invariant if all \(g \in G\) act via \(B\)-orthogonal operators. The 294 condition of \(\mathfrak{g}\)-invariance can thus be though-of as an 295 \emph{infinitesimal approximation} of the notion of a \(G\)-invariant form. 296 Indeed \(\operatorname{Lie}(\operatorname{O}(B))\) is precisely the Lie 297 subalgebra of \(\mathfrak{gl}(V)\) consisting of antisymmetric operators \(V 298 \to V\). 299 \end{note} 300 301 An interesting example of an invariant bilinear form is the so called 302 \emph{Killing form}. 303 304 \begin{definition}\index{invariant bilinear form!Killing form}\index{Killing form} 305 Given a finite-dimensional Lie algebra \(\mathfrak{g}\), the symmetric 306 bilinear form 307 \begin{align*} 308 \kappa : \mathfrak{g} \times \mathfrak{g} & \to K \\ 309 (X, Y) & 310 \mapsto \operatorname{Tr}(\operatorname{ad}(X) \operatorname{ad}(Y)) 311 \end{align*} 312 is called \emph{the Killing form of \(\mathfrak{g}\)}. 313 \end{definition} 314 315 The fact that the Killing form is an invariant form follows directly from the 316 identity \(\operatorname{Tr}([X, Y] Z) = \operatorname{Tr}(X [Y, Z])\), \(X, Y, 317 Z \in \mathfrak{gl}_n(K)\). In fact this same identity show\dots 318 319 \begin{lemma}\index{invariant bilinear form!bilinear form of a \(\mathfrak{g}\)-module} 320 Given a finite-dimensional \(\mathfrak{g}\)-module \(M\), the symmetric 321 bilinear form 322 \begin{align*} 323 \kappa_M : \mathfrak{g} \times \mathfrak{g} & \to K \\ 324 (X, Y) & \mapsto \operatorname{Tr}(X\!\restriction_M \, Y\!\restriction_M) 325 \end{align*} 326 is \(\mathfrak{g}\)-invariant. 327 \end{lemma} 328 329 The reason why we are discussing invariant bilinear forms is the following 330 characterization of finite-dimensional semisimple Lie algebras, known as 331 \emph{Cartan's criterion for semisimplicity}. 332 333 \begin{proposition} 334 Let \(\mathfrak{g}\) be a Lie algebra. The following conditions are 335 equivalent. 336 \begin{enumerate} 337 \item \(\mathfrak{g}\) is semisimple. 338 \item For each non-trivial finite-dimensional \(\mathfrak{g}\)-module 339 \(M\), the \(\mathfrak{g}\)-invariant bilinear form 340 \begin{align*} 341 \kappa_M : \mathfrak{g} \times \mathfrak{g} & 342 \to K \\ 343 (X, Y) & 344 \mapsto \operatorname{Tr}(X\!\restriction_M \, Y\!\restriction_M) 345 \end{align*} 346 is non-degenerate\footnote{A symmetric bilinear form $B : \mathfrak{g} 347 \times \mathfrak{g} \to K$ is called non-degenerate if $B(X, Y) = 0$ for 348 all $Y \in \mathfrak{g}$ implies $X = 0$.}. 349 \item The Killing form \(\kappa\) is non-degenerate. 350 \end{enumerate} 351 \end{proposition} 352 353 This proof is somewhat technical, but the idea behind it is simple. First, for 354 \strong{(i)} \(\implies\) \strong{(ii)} we show that \(\mathfrak{a} = \{ X \in 355 \mathfrak{g} : \kappa_M(X, Y) = 0 \, \forall Y \in \mathfrak{g}\}\) is a 356 solvable ideal of \(\mathfrak{g}\). Hence \(\mathfrak{a} = 0\). For 357 \strong{(ii)} \(\implies\) \strong{(iii)} it suffices to take \(M = 358 \mathfrak{g}\) the adjoint \(\mathfrak{g}\)-module. Finally, for \strong{(iii)} 359 \(\implies\) \strong{(i)} we note that the orthogonal complement of any 360 \(\mathfrak{a} \normal \mathfrak{g}\) with respect to the Killing form 361 \(\kappa\) is an ideal \(\mathfrak{b}\) of \(\mathfrak{g}\) with \(\mathfrak{g} 362 = \mathfrak{a} \oplus \mathfrak{b}\). Furthermore, the Killing form of 363 \(\mathfrak{a}\) is the restriction \(\kappa\!\restriction_{\mathfrak{a}}\) of 364 the Killing form of \(\mathfrak{g}\) to \(\mathfrak{a} \times \mathfrak{a}\), 365 which is non-degenerate. It then follows from induction in \(\dim 366 \mathfrak{a}\) that \(\mathfrak{g}\) is the sum of simple ideals. 367 368 We refer the reader to \cite[ch. 5]{humphreys} for a complete proof. Without 369 further ado, we may proceed to our\dots 370 371 \section{Proof of Complete Reducibility} 372 373 Let \(\mathfrak{g}\) be a finite-dimensional Lie algebra over \(K\). We want to 374 establish that if \(\mathfrak{g}\) is semisimple then all finite-dimensional 375 \(\mathfrak{g}\)-modules are semisimple. Historically, this was first proved by 376 Herman Weyl for \(K = \mathbb{C}\), using his knowledge of smooth 377 representations of compact Lie groups. Namely, Weyl showed that any 378 finite-dimensional semisimple complex Lie algebra is (isomorphic to) the 379 complexification of the Lie algebra of a unique simply connected compact Lie 380 group, known as its \emph{compact form}. Hence the category of the 381 finite-dimensional modules of a given complex semisimple algebra is equivalent 382 to that of the finite-dimensional smooth representations of its compact form, 383 whose representations are known to be completely reducible because of Maschke's 384 Theorem -- see \cite[ch. 3]{serganova} for instance. 385 386 This proof, however, is heavily reliant on the geometric structure of 387 \(\mathbb{C}\). In other words, there is no hope for generalizing this for some 388 arbitrary \(K\). Fortunately for us, there is a much simpler, completely 389 algebraic proof of complete reducibility, which works for algebras over any 390 algebraically closed field of characteristic zero. The algebraic proof included 391 in here is mainly based on that of \cite[ch. 6]{kirillov}, and uses some basic 392 homological algebra. Admittedly, much of the homological algebra used in here 393 could be concealed from the reader, which would make the exposition more 394 accessible -- see \cite{humphreys} for instance. 395 396 However, this does not change the fact the arguments used in this proof are 397 essentially homological in nature. Hence we consider it more productive to use 398 the full force of the language of homological algebra, instead of burring the 399 reader in a pile of unmotivated, yet entirely elementary arguments. 400 Furthermore, the homological algebra used in here is actually \emph{very 401 basic}. In fact, all we need to know is\dots 402 403 \begin{theorem}\label{thm:ext-exacts-seqs}\index{\(\operatorname{Ext}\) functors} 404 There is a sequence of bifunctors \(\operatorname{Ext}^i : 405 \mathfrak{g}\text{-}\mathbf{Mod}^{\operatorname{op}} \times 406 \mathfrak{g}\text{-}\mathbf{Mod} \to K\text{-}\mathbf{Vect}\), \(i \ge 0\) 407 such that, given a \(\mathfrak{g}\)-module \(L'\), every exact sequence of 408 \(\mathfrak{g}\)-modules 409 \begin{center} 410 \begin{tikzcd} 411 0 \rar & N \rar{f} & M \rar{g} & L \rar & 0 412 \end{tikzcd} 413 \end{center} 414 induces long exact sequences 415 \begin{center} 416 \begin{tikzcd} 417 0 \rar & 418 \operatorname{Hom}_{\mathfrak{g}}(L', N) 419 \rar{f \circ -}\ar[draw=none]{d}[name=X, anchor=center]{} & 420 \operatorname{Hom}_{\mathfrak{g}}(L', M) \rar{g \circ -} & 421 \operatorname{Hom}_{\mathfrak{g}}(L', L) 422 \ar[rounded corners, 423 to path={ -- ([xshift=2ex]\tikztostart.east) 424 |- (X.center) \tikztonodes 425 -| ([xshift=-2ex]\tikztotarget.west) 426 -- (\tikztotarget)}]{dll}[at end]{} \\ & 427 \operatorname{Ext}^1(L', N) 428 \rar\ar[draw=none]{d}[name=Y, anchor=center]{} & 429 \operatorname{Ext}^1(L', M) \rar & 430 \operatorname{Ext}^1(L', L) 431 \ar[rounded corners, 432 to path={ -- ([xshift=2ex]\tikztostart.east) 433 |- (Y.center) \tikztonodes 434 -| ([xshift=-2ex]\tikztotarget.west) 435 -- (\tikztotarget)}]{dll}[at end]{} \\ & 436 \operatorname{Ext}^2(L', N) \rar & 437 \operatorname{Ext}^2(L', M) \rar & 438 \operatorname{Ext}^2(L', L) \rar[dashed] & 439 \cdots 440 \end{tikzcd} 441 \end{center} 442 and 443 \begin{center} 444 \begin{tikzcd} 445 0 \rar & 446 \operatorname{Hom}_{\mathfrak{g}}(L, L') 447 \rar{- \circ g}\ar[draw=none]{d}[name=X, anchor=center]{} & 448 \operatorname{Hom}_{\mathfrak{g}}(M, L') \rar{- \circ f} & 449 \operatorname{Hom}_{\mathfrak{g}}(N, L') 450 \ar[rounded corners, 451 to path={ -- ([xshift=2ex]\tikztostart.east) 452 |- (X.center) \tikztonodes 453 -| ([xshift=-2ex]\tikztotarget.west) 454 -- (\tikztotarget)}]{dll}[at end]{} \\ & 455 \operatorname{Ext}^1(L, L') 456 \rar\ar[draw=none]{d}[name=Y, anchor=center]{} & 457 \operatorname{Ext}^1(M, L') \rar & 458 \operatorname{Ext}^1(N, L') 459 \ar[rounded corners, 460 to path={ -- ([xshift=2ex]\tikztostart.east) 461 |- (Y.center) \tikztonodes 462 -| ([xshift=-2ex]\tikztotarget.west) 463 -- (\tikztotarget)}]{dll}[at end]{} \\ & 464 \operatorname{Ext}^2(L, L') \rar & 465 \operatorname{Ext}^2(M, L') \rar & 466 \operatorname{Ext}^2(N, L') \rar[dashed] & 467 \cdots 468 \end{tikzcd} 469 \end{center} 470 \end{theorem} 471 472 \begin{theorem}\label{thm:ext-1-classify-short-seqs} 473 Given \(\mathfrak{g}\)-modules \(N\) and \(L\), there is a one-to-one 474 correspondence between elements of \(\operatorname{Ext}^1(L, N)\) and 475 isomorphism classes of short exact sequences 476 \begin{center} 477 \begin{tikzcd} 478 0 \rar & N \rar & M \rar & L \rar & 0 479 \end{tikzcd} 480 \end{center} 481 482 In particular, \(\operatorname{Ext}^1(L, N) = 0\) if, and only if every short 483 exact sequence of \(\mathfrak{g}\)-modules with \(N\) and \(L\) in the 484 extremes splits. 485 \end{theorem} 486 487 We should point out that, although we have not provided an explicit definition 488 of the bifunctors \(\operatorname{Ext}^i\), they are uniquely determined by 489 the conditions of Theorem~\ref{thm:ext-exacts-seqs} and some additional 490 minimality constraints. This is, of course, \emph{far} from a comprehensive 491 account of homological algebra. Nevertheless, this is all we need. We refer the 492 reader to \cite{harder} for a complete exposition, or to part II of 493 \cite{ribeiro} for a more modern account using derived categories. 494 495 We are particularly interested in the case where \(L' = K\) is the trivial 496 \(\mathfrak{g}\)-module. Namely, we may define\dots 497 498 \begin{definition}\index{Lie algebra!cohomology}\index{cohomology of Lie algebras} 499 Given a Lie algebra \(\mathfrak{g}\) and a \(\mathfrak{g}\)-module \(M\), we 500 refer to the Abelian group \(H^i(\mathfrak{g}, M) = \operatorname{Ext}^i(K, 501 M)\) as \emph{the \(i\)-th Lie algebra cohomology group of \(\mathfrak{g}\) 502 with coefficients in \(M\)}. 503 \end{definition} 504 505 \begin{definition}\index{cohomology of Lie algebras!invariants} 506 Given a \(\mathfrak{g}\)-module \(M\), we call the vector space 507 \(M^{\mathfrak{g}} = \{m \in M : X \cdot m = 0 \; \forall X \in 508 \mathfrak{g}\}\) \emph{the space of invariants of \(M\)}. A simple 509 calculations shows that a \(\mathfrak{g}\)-homomorphism \(f : M \to N\) takes 510 invariants to invariants, so that \(f\) restricts to a map \(M^{\mathfrak{g}} 511 \to N^{\mathfrak{g}}\). This construction thus yields a functor 512 \(-^{\mathfrak{g}} : \mathfrak{g}\text{-}\mathbf{Mod} \to 513 K\text{-}\mathbf{Vect}\). 514 \end{definition} 515 516 \begin{example} 517 Let \(M\) be a \(\mathfrak{g}\)-module. Then \(M\) is a direct sum of copies 518 of the trivial \(\mathfrak{g}\)-module if, and only if \(M = 519 M^{\mathfrak{g}}\). 520 \end{example} 521 522 \begin{example}\label{ex:hom-invariants-are-g-homs} 523 Let \(M\) and \(N\) be \(\mathfrak{g}\)-modules. Then \(\operatorname{Hom}(M, 524 N)^{\mathfrak{g}} = \operatorname{Hom}_{\mathfrak{g}}(M, N)\). Indeed, given 525 a \(K\)-linear map \(f : M \to N\) we find 526 \[ 527 \begin{split} 528 f \in \operatorname{Hom}(M, N)^{\mathfrak{g}} 529 & \iff X \cdot f(m) - f(X \cdot m) = (X \cdot f)(m) = 0 530 \; \forall X \in \mathfrak{g}, m \in M \\ 531 & \iff X \cdot f(m) = f(X \cdot m) 532 \; \forall X \in \mathfrak{g}, m \in M \\ 533 & \iff f \in \operatorname{Hom}_{\mathfrak{g}}(M, N) 534 \end{split} 535 \] 536 \end{example} 537 538 The Lie algebra cohomology groups are very much related to invariants of 539 \(\mathfrak{g}\)-modules. Namely, constructing a \(\mathfrak{g}\)-homomorphism 540 \(f : K \to M\) is precisely the same as fixing an invariant of \(M\) -- 541 corresponding to \(f(1)\), which must be an invariant for \(f\) to be a 542 \(\mathfrak{g}\)-homomorphism. Formally, this translates to the existence of a 543 canonical isomorphism of functors 544 \(\operatorname{Hom}_{\mathfrak{g}}(K, -) \isoto {-}^{\mathfrak{g}}\) given by 545 \begin{align*} 546 \operatorname{Hom}_{\mathfrak{g}}(K, M) & \isoto M^{\mathfrak{g}} \\ 547 f & \mapsto f(1) 548 \end{align*} 549 550 This implies\dots 551 552 \begin{corollary} 553 Every short exact sequence of \(\mathfrak{g}\)-modules 554 \begin{center} 555 \begin{tikzcd} 556 0 \rar & N \rar{f} & M \rar{g} & L \rar & 0 557 \end{tikzcd} 558 \end{center} 559 induces a long exact sequence 560 \begin{center} 561 \begin{tikzcd} 562 0 \rar & 563 N^{\mathfrak{g}} \rar{f} 564 \ar[draw=none]{d}[name=X, anchor=center]{} & 565 M^{\mathfrak{g}} \rar{g} & 566 L^{\mathfrak{g}} 567 \ar[rounded corners, 568 to path={ -- ([xshift=2ex]\tikztostart.east) 569 |- (X.center) \tikztonodes 570 -| ([xshift=-2ex]\tikztotarget.west) 571 -- (\tikztotarget)}]{dll}[at end]{} \\ & 572 H^1(\mathfrak{g}, N) \rar\ar[draw=none]{d}[name=Y, anchor=center]{} & 573 H^1(\mathfrak{g}, M) \rar & 574 H^1(\mathfrak{g}, L) 575 \ar[rounded corners, 576 to path={ -- ([xshift=2ex]\tikztostart.east) 577 |- (Y.center) \tikztonodes 578 -| ([xshift=-2ex]\tikztotarget.west) 579 -- (\tikztotarget)}]{dll}[at end]{} \\ & 580 H^2(\mathfrak{g}, N) \rar & 581 H^2(\mathfrak{g}, M) \rar & 582 H^2(\mathfrak{g}, L) \rar[dashed] & 583 \cdots 584 \end{tikzcd} 585 \end{center} 586 \end{corollary} 587 588 \begin{proof} 589 We have an isomorphism of sequences 590 \begin{center} 591 \begin{tikzcd} 592 0 \rar & 593 \operatorname{Hom}_{\mathfrak{g}}(K, N) \rar{f \circ -} \dar & 594 \operatorname{Hom}_{\mathfrak{g}}(K, M) \rar{g \circ -} \dar & 595 \operatorname{Hom}_{\mathfrak{g}}(K, L) \rar \dar & 596 H^1(\mathfrak{g}, N) \rar[dashed]\dar[Rightarrow, no head] & \cdots \\ 597 0 \rar & 598 N^{\mathfrak{g}} \rar[swap]{f} & 599 M^{\mathfrak{g}} \rar[swap]{g} & 600 L^{\mathfrak{g}} \rar & 601 H^1(\mathfrak{g}, N) \rar[dashed] & 602 \cdots 603 \end{tikzcd} 604 \end{center} 605 606 By Theorem~\ref{thm:ext-exacts-seqs} the sequence on the top is exact. Hence 607 so is the sequence on the bottom. 608 \end{proof} 609 610 This is all well and good, but what does any of this have to do with complete 611 reducibility? Well, in general cohomology theories really shine when one is 612 trying to control obstructions of some kind. In our case, the bifunctor 613 \(H^1(\mathfrak{g}, \operatorname{Hom}(-, -)) : 614 \mathfrak{g}\text{-}\mathbf{Mod}^{\operatorname{op}} \times 615 \mathfrak{g}\text{-}\mathbf{Mod} \to K\text{-}\mathbf{Vect}\) classifies 616 obstructions to complete reducibility. Explicitly\dots 617 618 \begin{theorem} 619 There is a natural isomorphism \(\operatorname{Ext}^1 \isoto 620 H^1(\mathfrak{g}, \operatorname{Hom}(-, -))\). In particular, given 621 \(\mathfrak{g}\)-modules \(N\) and \(L\), there is a one-to-one 622 correspondence between elements of \(H^1(\mathfrak{g}, \operatorname{Hom}(L, 623 N))\) and isomorphism classes of short exact sequences 624 \begin{center} 625 \begin{tikzcd} 626 0 \rar & N \rar & M \rar & L \rar & 0 627 \end{tikzcd} 628 \end{center} 629 \end{theorem} 630 631 This is essentially a consequence of Example~\ref{ex:hom-invariants-are-g-homs} 632 and Theorem~\ref{thm:ext-1-classify-short-seqs}, as well as the minimality 633 conditions that characterize \(\operatorname{Ext}^1\). For the readers already 634 familiar with homological algebra: the correspondence between 635 \(H^1(\mathfrak{g}, \operatorname{Hom}(L, N))\) and short exact sequences of 636 \(\mathfrak{g}\)-modules can be described in very concrete terms by considering 637 a canonical free resolution 638 \begin{center} 639 \begin{tikzcd} 640 \cdots \rar[dashed] & 641 \mathcal{U}(\mathfrak{g}) \otimes (\wedge^2 \mathfrak{g}) \rar & 642 \mathcal{U}(\mathfrak{g}) \otimes \mathfrak{g} \rar & 643 \mathcal{U}(\mathfrak{g}) \rar & 644 K \rar & 645 0 646 \end{tikzcd} 647 \end{center} 648 of the trivial \(\mathfrak{g}\)-module \(K\), known as \emph{the 649 Chevalley-Eilenberg resolution}, which provides an explicit construction of the 650 cohomology groups -- see \cite[sec.~1.3C]{cohomologies-lie} or 651 \cite[sec.~24]{symplectic-physics} for further details. 652 653 We will use the previous result implicitly in our proof, but we will not prove 654 it in its full force. Namely, we will show that if \(\mathfrak{g}\) is 655 semisimple then \(H^1(\mathfrak{g}, M) = 0\) for all finite-dimensional \(M\), 656 and that the fact that \(H^1(\mathfrak{g}, \operatorname{Hom}(L, N)) = 0\) for 657 all finite-dimensional \(N\) and \(L\) implies complete reducibility. To that 658 end, we introduce a distinguished element of \(\mathcal{U}(\mathfrak{g})\), 659 known as \emph{the Casimir element of a \(\mathfrak{g}\)-module}. 660 661 \begin{definition}\label{def:casimir-element}\index{Casimir element} 662 Let \(\mathfrak{g}\) be a finite-dimensional semisimple Lie algebra and \(M\) 663 be a finite-dimensional \(\mathfrak{g}\)-module. Let \(\{X_i\}_i\) be a basis 664 for \(\mathfrak{g}\) and denote by \(\{X^i\}_i \subset \mathfrak{g}\) its 665 dual basis with respect to the form \(\kappa_M\) -- i.e. the unique basis for 666 \(\mathfrak{g}\) satisfying \(\kappa_M(X_i, X^j) = \delta_{i j}\), whose 667 existence is a consequence of the non-degeneracy of \(\kappa_M\). We call 668 \[ 669 \Omega_M = X_1 X^1 + \cdots + X_r X^r \in \mathcal{U}(\mathfrak{g}) 670 \] 671 the \emph{Casimir element of \(M\)}. 672 \end{definition} 673 674 \begin{lemma} 675 The definition of \(\Omega_M\) is independent of the choice of basis 676 \(\{X_i\}_i\). 677 \end{lemma} 678 679 \begin{proof} 680 Whatever basis \(\{X_i\}_i\) we choose, the image of \(\Omega_M\) under the 681 canonical isomorphism \(\mathfrak{g} \otimes \mathfrak{g} \isoto \mathfrak{g} 682 \otimes \mathfrak{g}^* \isoto \operatorname{End}(\mathfrak{g})\) is the 683 identity operator\footnote{Here the isomorphism $\mathfrak{g} \otimes 684 \mathfrak{g} \isoto \mathfrak{g} \otimes \mathfrak{g}^*$ is given by 685 tensoring the identity $\mathfrak{g} \to \mathfrak{g}$ with the isomorphism 686 $\mathfrak{g} \isoto \mathfrak{g}^*$ induced by the form $\kappa_M$.}. 687 \end{proof} 688 689 \begin{proposition} 690 The Casimir element \(\Omega_M \in \mathcal{U}(\mathfrak{g})\) is central, so 691 that \(\Omega_M\!\restriction_N : N \to N\) is a 692 \(\mathfrak{g}\)-homomorphism for any \(\mathfrak{g}\)-module \(N\). 693 Furthermore, \(\Omega_M\) acts on \(M\) as a nonzero scalar operator whenever 694 \(M\) is a non-trivial finite-dimensional simple \(\mathfrak{g}\)-module. 695 \end{proposition} 696 697 \begin{proof} 698 To see that \(\Omega_M\) is central fix a basis \(\{X_i\}_i\) for 699 \(\mathfrak{g}\) and denote by \(\{X^i\}_i\) its dual basis with respect to 700 \(\kappa_M\), as in 701 Definition~\ref{def:casimir-element}. Given any \(X \in \mathfrak{g}\), it 702 follows from definition of the \(X^i\) that \(X = \kappa_M(X, X^1) X_1 + 703 \cdots + \kappa_M(X, X^r) X_r = \kappa_M(X, X_1) X^1 + \cdots + \kappa_M(X, 704 X_r) X^r\). 705 706 In particular, it follows from the invariance of \(\kappa_M\) that 707 \[ 708 \begin{split} 709 [X, \Omega_M] 710 & = \sum_i [X, X_i X^i] \\ 711 & = \sum_i [X, X_i] X^i + \sum_i X_i [X, X^i] \\ 712 & = \sum_{i j} \kappa_M([X, X_i], X^j) X_j X^i 713 + \sum_{i j} \kappa_M([X, X^i], X_j) X_i X^j \\ 714 & = \sum_{i j} (\kappa_M([X, X_j], X^i) + \kappa_M(X_j, [X, X^i])) 715 X_i X^j \\ 716 & = 0 717 \end{split}, 718 \] 719 and \(\Omega_M\) is central. This implies that \(\Omega_M\!\restriction_N : N 720 \to N\) is a \(\mathfrak{g}\)-homomorphism for all \(\mathfrak{g}\)-modules 721 \(N\): its action commutes with the action of any other element of 722 \(\mathfrak{g}\). 723 724 In particular, it follows from Schur's Lemma that if \(M\) is 725 finite-dimensional and simple then \(\Omega_M\) acts on \(M\) as a scalar 726 operator. To see that this scalar is nonzero we compute 727 \[ 728 \operatorname{Tr}(\Omega_M\!\restriction_M) 729 = \operatorname{Tr}(X_1\!\restriction_M X^1\!\restriction_M) 730 + \cdots 731 + \operatorname{Tr}(X_r\!\restriction_M X^r\!\restriction_M) 732 = \dim \mathfrak{g}, 733 \] 734 so that \(\Omega_M\!\restriction_M = \lambda \operatorname{Id}\) for 735 \(\lambda = \frac{\dim \mathfrak{g}}{\dim M} \ne 0\). 736 \end{proof} 737 738 As promised, the Casimir element of a \(\mathfrak{g}\)-module can be used to 739 establish\dots 740 741 \begin{proposition}\label{thm:first-cohomology-vanishes} 742 Suppose \(\mathfrak{g}\) is semisimple and let \(M\) be a finite-dimensional 743 \(\mathfrak{g}\)-module. Then \(H^1(\mathfrak{g}, M) = 0\). 744 \end{proposition} 745 746 \begin{proof} 747 We begin by the case where \(M\) is simple. Due to 748 Theorem~\ref{thm:ext-1-classify-short-seqs}, it suffices to show that any 749 exact sequence of the form 750 \begin{equation}\label{eq:exact-seq-h1-vanishes} 751 \begin{tikzcd} 752 0 \rar & M \rar{f} & N \rar{g} & K \rar & 0 753 \end{tikzcd} 754 \end{equation} 755 splits. 756 757 If \(M = K\) is the trivial \(\mathfrak{g}\)-module then the exactness of 758 \begin{equation}\label{eq:trivial-extrems-exact-seq} 759 \begin{tikzcd} 760 0 \rar & K \rar{f} & N \rar{g} & K \rar & 0 761 \end{tikzcd} 762 \end{equation} 763 implies \(N\) is 2-dimensional. Take any nonzero \(n \in N\) outside of the 764 image of \(f\). 765 766 % TODOOOOOOOOO: Fix this 767 % TODO: U(g) w doesn't need to be irreducible a priori. In fact we will show 768 % U(g) w = 0, so this whole argument is inconsistant 769 % TODO: The way to fix this is to prove that rho(g) is both nilpotent -- 770 % because the action of every element of g is strictly upper triangular -- and 771 % semisimple -- because it is a quotient of g, which is semisimple. We thus 772 % have rho(g) = 0, so that W is trivial 773 Since \(\dim N = 2\), the simple component \(\mathcal{U}(\mathfrak{g}) \cdot 774 n\) of \(n\) in \(N\) is either \(K n\) or \(N\) itself. But this component 775 cannot be \(N\), since the image of \(f\) is a \(1\)-dimensional 776 \(\mathfrak{g}\)-module -- i.e. a proper nonzero submodule. Hence \(K n\) is 777 invariant under the action of \(\mathfrak{g}\). In particular, \(X \cdot n = 778 0\) for all \(X \in \mathfrak{g}\). Since \(n\) lies outside the image of 779 \(f\), \(g(n) \ne 0\) -- which is to say, \(n \notin \ker g = 780 \operatorname{im} f\). This implies the map \(K \to N\) that takes \(1\) to 781 \(\sfrac{n}{g(n)}\) is a splitting of (\ref{eq:trivial-extrems-exact-seq}). 782 783 Now suppose that \(M\) is non-trivial, so that \(\Omega_M\) acts on \(M\) as 784 \(\lambda\) for some \(\lambda \ne 0\). Denote by \(N^\mu\) the generalized 785 eigenspace of \(\Omega_M\!\restriction_N : N \to N\) associated with \(\mu 786 \in K\). If we identify \(M\) with \(f(M)\), it is clear that \(M \subset 787 N^\lambda\). The exactness of (\ref{eq:exact-seq-h1-vanishes}) implies \(\dim 788 N = \dim M + 1\), so that either \(N^\lambda = M\) or \(N^\lambda = N\). But 789 if \(N^\lambda = N\) then there is some nonzero \(n \in N^\lambda\) with \(n 790 \notin M = \ker g\) such that 791 \[ 792 0 793 = (\Omega_M - \lambda)^r \cdot n 794 = \sum_{k = 0}^r (-1)^k \binom{r}{k} \lambda^k \Omega_M^{r - k} \cdot n 795 \] 796 for some \(r \ge 1\). 797 798 In particular, 799 \[ 800 (- \lambda)^{r - 1} g(n) 801 = \sum_{k = 0}^{r - 1} (-1)^k \binom{r}{k} \lambda^k 802 g(\Omega_M^{r - k} \cdot n) 803 = \sum_{k = 0}^{r - 1} (-1)^k \binom{r}{k} \lambda^k 804 \underbrace{\Omega_M^{r - k} \cdot g(n)}_{= \; 0} 805 = 0, 806 \] 807 which is a contradiction -- given that neither \((-\lambda)^{r - 1}\) nor 808 \(g(n)\) are nil. Hence \(M = N^\lambda\) and there must be some other 809 eigenvalue \(\mu\) of \(\Omega_M\!\restriction_N\). For any such \(\mu\) and 810 any eigenvector \(n \in N_\mu\), 811 \[ 812 \mu g(n) 813 = g(\mu n) 814 = g(\Omega_M \cdot n) 815 = \Omega_M \cdot g(n) 816 = 0 817 \] 818 implies \(\mu = 0\), so that the eigenvalues of the action of \(\Omega_M\) on 819 \(N\) are precisely \(\lambda\) and \(0\). 820 821 Now notice that \(N^0\) is in fact a submodule of \(N\). Indeed, 822 given \(n \in N^0\) and \(X \in \mathfrak{g}\), it follows from the fact that 823 \(\Omega_M\) is central that 824 \[ 825 \Omega_M^r \cdot (X \cdot n) = X \cdot (\Omega_M^r \cdot n) = X \cdot 0 = 0 826 \] 827 for some \(r\). Hence \(N = M \oplus N^0\) as \(\mathfrak{g}\)-modules. The 828 homomorphism \(g\) thus induces an isomorphism \(N^0 \cong \mfrac{N}{M} 829 \isoto K\), which translates to a splitting of 830 (\ref{eq:exact-seq-h1-vanishes}). 831 832 Finally, we consider the case where \(M\) is not simple. Suppose 833 \(H^1(\mathfrak{g}, N) = 0\) for all \(\mathfrak{g}\)-modules with \(\dim N < 834 \dim M\) and let \(N \subset M\) be a proper nonzero submodule. Then the 835 exact sequence 836 \begin{center} 837 \begin{tikzcd} 838 0 \rar & N \rar & M \rar & \sfrac{M}{N} \rar & 0 839 \end{tikzcd} 840 \end{center} 841 induces a long exact sequence of the form 842 \begin{equation}\label{eq:standard-h1-ext-seq} 843 \begin{tikzcd} 844 \cdots \rar[dashed] & 845 H^1(\mathfrak{g}, N) \rar & 846 H^1(\mathfrak{g}, M) \rar & 847 H^1(\mathfrak{g}, \sfrac{M}{N}) \rar[dashed] & 848 \cdots 849 \end{tikzcd} 850 \end{equation} 851 852 Since \(\dim N < \dim M\), it follows \(H^1(\mathfrak{g}, N) = 0\). In 853 addition, since \(\dim N > 0\), we find \(\dim \mfrac{M}{N} < \dim M\) and 854 thus \(H^1(\mathfrak{g}, \sfrac{M}{N}) = 0\). The exactness of 855 (\ref{eq:standard-h1-ext-seq}) then implies \(H^1(\mathfrak{g}, M) = 0\). 856 Hence by induction in \(\dim M\) we find \(H^1(\mathfrak{g}, M) = 0\) for all 857 finite-dimensional \(M\). We are done. 858 \end{proof} 859 860 We are now finally ready to prove\dots 861 862 \begin{theorem}[Weyl]\label{thm:weyl-theorem} 863 Given a semisimple Lie algebra \(\mathfrak{g}\), every finite-dimensional 864 \(\mathfrak{g}\)-module is semisimple. 865 \end{theorem} 866 867 \begin{proof} 868 Let 869 \begin{equation}\label{eq:generict-exact-sequence} 870 \begin{tikzcd} 871 0 \rar & N \rar{f} & M \rar{g} & L \rar & 0 872 \end{tikzcd} 873 \end{equation} 874 be a short exact sequence of finite-dimensional \(\mathfrak{g}\)-modules. We 875 want to establish that (\ref{eq:generict-exact-sequence}) splits. 876 877 We have an exact sequence 878 \begin{center} 879 \begin{tikzcd} 880 0 \rar & 881 \operatorname{Hom}(L, N) \rar{f \circ -} & 882 \operatorname{Hom}(L, M) \rar{g \circ -} & 883 \operatorname{Hom}(L, L) \rar & 884 0 885 \end{tikzcd} 886 \end{center} 887 of vector spaces. Since all maps involved are \(\mathfrak{g}\)-homomorphisms, 888 this is an exact sequence of \(\mathfrak{g}\)-modules. This then induces a 889 long exact sequence 890 \begin{center} 891 \begin{tikzcd} 892 0 \rar & 893 \operatorname{Hom}(L, N)^{\mathfrak{g}} \rar{f \circ -} 894 \ar[draw=none]{d}[name=X, anchor=center]{} & 895 \operatorname{Hom}(L, M)^{\mathfrak{g}} \rar{g \circ -} & 896 \operatorname{Hom}(L, L)^{\mathfrak{g}} 897 \ar[rounded corners, 898 to path={ -- ([xshift=2ex]\tikztostart.east) 899 |- (X.center) \tikztonodes 900 -| ([xshift=-2ex]\tikztotarget.west) 901 -- (\tikztotarget)}]{dll}[at end]{} \\ & 902 H^1(\mathfrak{g}, \operatorname{Hom}(L, N)) \rar & 903 H^1(\mathfrak{g}, \operatorname{Hom}(L, M)) \rar & 904 H^1(\mathfrak{g}, \operatorname{Hom}(L, L)) \rar[dashed] & 905 \cdots 906 \end{tikzcd} 907 \end{center} 908 of vector spaces. 909 910 But \(H^1(\mathfrak{g}, \operatorname{Hom}(L, N))\) vanishes because of 911 Proposition~\ref{thm:first-cohomology-vanishes}. In addition, recall from 912 Example~\ref{ex:hom-invariants-are-g-homs} that \(\operatorname{Hom}(L, 913 L')^{\mathfrak{g}} = \operatorname{Hom}_{\mathfrak{g}}(L, L')\). We thus have 914 a short exact sequence 915 \begin{center} 916 \begin{tikzcd} 917 0 \rar & 918 \operatorname{Hom}_{\mathfrak{g}}(L, N) \rar{f \circ -} & 919 \operatorname{Hom}_{\mathfrak{g}}(L, M) \rar{g \circ -} & 920 \operatorname{Hom}_{\mathfrak{g}}(L, L) \rar & 921 0 922 \end{tikzcd} 923 \end{center} 924 925 In particular, there is some \(\mathfrak{g}\)-homomorphism \(s : L \to M\) 926 such that \(g \circ s : L \to L\) is the identity operator. In other words 927 \begin{center} 928 \begin{tikzcd} 929 0 \rar & N \rar{f} & M \rar{g} & L \rar \lar[bend left]{s} & 0 930 \end{tikzcd} 931 \end{center} 932 is a splitting of (\ref{eq:generict-exact-sequence}). 933 \end{proof} 934 935 Theorem~\ref{thm:weyl-theorem} typically fails in the infinite-dimensional 936 setting. For instance, consider\dots 937 938 \begin{example}\label{ex:regular-mod-is-not-semisimple} 939 The regular \(\mathfrak{g}\)-module \(\mathcal{U}(\mathfrak{g})\) is an 940 indecomposable module which is not simple. In particular, 941 \(\mathcal{U}(\mathfrak{g})\) is not semisimple. To see this, notice that the 942 submodules of \(\mathcal{U}(\mathfrak{g})\) are precisely its left ideals. If 943 we suppose that \(I, J \normal \mathcal{U}(\mathfrak{g})\) are such that 944 \(\mathcal{U}(\mathfrak{g}) = I \oplus J\) as \(\mathfrak{g}\)-modules, we 945 can find \(u \in I\) and \(v \in J\) such that \(1 = u + v\). The PBW Theorem 946 then implies that \(u\) and \(v\) commute, so that \(uv = vu \in I \cap J = 947 0\). Since \(\mathcal{U}(\mathfrak{g})\) is a domain, either \(u = 0\) or \(v 948 = 0\). Given that \(1 = u + v\), \(u = 1\) or \(v = 1\). Hence either \(I = 949 \mathcal{U}(\mathfrak{g})\) and \(J = 0\) or \(I = 0\) and \(J = 950 \mathcal{U}(\mathfrak{g})\), as required. 951 \end{example} 952 953 We should point out that these last results are just the beginning of a well 954 developed cohomology theory. For example, a similar argument involving the 955 Casimir elements can be used to show that \(H^i(\mathfrak{g}, M) = 0\) for all 956 semisimple \(\mathfrak{g}\) and all non-trivial finite-dimensional simple 957 \(M\), \(i > 0\). For \(K = \mathbb{C}\), the Lie algebra cohomology groups of 958 the algebra \(\mathfrak{g} = \mathbb{C} \otimes \operatorname{Lie}(G)\) are 959 intimately related with the topological cohomologies -- i.e. singular 960 cohomology, de Rham cohomology, etc. -- of \(G\) with coefficients in 961 \(\mathbb{C}\). We refer the reader to \cite{cohomologies-lie} and 962 \cite[sec.~24]{symplectic-physics} for further details. 963 964 Complete reducibility can be generalized for arbitrary -- not necessarily 965 semisimple -- \(\mathfrak{g}\), to a certain extent, by considering the exact 966 sequence 967 \begin{center} 968 \begin{tikzcd} 969 0 \rar & 970 \mathfrak{rad}(\mathfrak{g}) \rar & 971 \mathfrak{g} \rar & 972 \mfrac{\mathfrak{g}}{\mathfrak{rad}(\mathfrak{g})} \rar & 973 0 974 \end{tikzcd} 975 \end{center} 976 977 This sequence always splits for finite-dimensional \(\mathfrak{g}\), which in 978 light of Example~\ref{ex:all-simple-reps-are-tensor-prod} implies we can deduce 979 information about \(\mathfrak{g}\)-modules by studying the modules of its 980 ``semisimple part'' \(\mfrac{\mathfrak{g}}{\mathfrak{rad}(\mathfrak{g})}\) -- 981 see Proposition~\ref{thm:quotients-by-rads}. In practice this translates 982 to\dots 983 984 \begin{proposition}[Lie]\label{thm:lie-thm-solvable-reps} 985 Let \(\mathfrak{g}\) be a solvable Lie algebra. Every finite-dimensional 986 simple \(\mathfrak{g}\)-module is \(1\)-dimensional. 987 \end{proposition} 988 989 \begin{corollary} 990 Let \(\mathfrak{g}\) be a Lie algebra. Every finite-dimensional simple 991 \(\mathfrak{g}\)-module is the tensor product of a simple 992 \(\mfrac{\mathfrak{g}}{\mathfrak{rad}(\mathfrak{g})}\)-module and a 993 \(1\)-dimensional \(\mathfrak{rad}(\mathfrak{g})\)-module. 994 \end{corollary} 995 996 \begin{proof} 997 This follows at once from Proposition~\ref{thm:lie-thm-solvable-reps} and 998 Example~\ref{ex:all-simple-reps-are-tensor-prod}. 999 \end{proof} 1000 1001 Having finally reduced our initial classification problem to that of 1002 classifying the finite-dimensional simple \(\mathfrak{g}\)-modules, we can now 1003 focus exclusively in this particular class of \(\mathfrak{g}\)-modules. 1004 However, there is so far no indication on how we could go about understanding 1005 them. In the next chapter we will explore some concrete examples in the hopes 1006 of finding a solution to our general problem.