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
introduction.tex (54542B)
1 \chapter{Introduction} 2 3 \pagenumbering{arabic} 4 \setcounter{page}{1} 5 6 Associative algebras have proven themselves remarkably useful throughout 7 mathematics. There is no lack of natural and interesting examples coming from a 8 diverse spectrum of different fields: topology, number theory, analysis, you 9 name it. Associative algebras have thus been studied at length, specially the 10 commutative ones. On the other hand, non-associative algebras have never 11 sustained the same degree of scrutiny. To this day, non-associative algebras 12 remain remarkably mysterious. Many have given up on attempting a systematic 13 investigation and focus instead on understanding particular classes of 14 non-associative algebras -- i.e. algebras satisfying 15 \emph{pseudo-associativity} conditions. 16 17 Perhaps the most fascinating class of non-associative algebras are the so 18 called \emph{Lie algebras}, and these will be the focus of these notes. 19 20 \begin{definition}\index{Lie algebra} 21 Given a field \(K\), a Lie algebra over \(K\) is a \(K\)-vector space 22 \(\mathfrak{g}\) endowed with an antisymmetric bilinear map \([\, ,] : 23 \mathfrak{g} \times \mathfrak{g} \to \mathfrak{g}\) -- which we call its 24 \emph{Lie bracket} -- satisfying the Jacobi identity 25 \[ 26 [X, [Y, Z]] + [Y, [Z, X]] + [Z, [X, Y]] = 0 27 \] 28 \end{definition} 29 30 \begin{definition}\index{Lie algebra!homomorphism} 31 Given two Lie algebras \(\mathfrak{g}\) and \(\mathfrak{h}\) over \(K\), a 32 homomorphism of Lie algebras \(\mathfrak{g} \to \mathfrak{h}\) is a 33 \(K\)-linear map \(f : \mathfrak{g} \to \mathfrak{h}\) which \emph{preserves 34 bracket} in the sense that 35 \[ 36 f([X, Y]) = [f(X), f(Y)] 37 \] 38 for all \(X, Y \in \mathfrak{g}\). The dimension \(\dim \mathfrak{g}\) of 39 \(\mathfrak{g}\) is its dimension as a \(K\)-vector space. 40 \end{definition} 41 42 The collection of Lie algebras over a fixed field \(K\) thus form a category, 43 which we call \(K\text{-}\mathbf{LieAlg}\). We are primarily interested in 44 finite-dimensional Lie algebras over algebraically closed fields of 45 characteristic \(0\). Hence from now on we assume \(K\) is algebraically closed 46 and \(\operatorname{char} K = 0\) unless explicitly stated otherwise. 47 Ironically, perhaps the most basic examples of Lie algebras are derived from 48 associative algebras. 49 50 \begin{example}\label{ex:inclusion-alg-in-lie-alg}\index{Lie algebra!Lie algebra of an associative algebra} 51 Given an associative \(K\)-algebra \(A\), we can view \(A\) as a Lie algebra 52 over \(K\) with the Lie bracket given by the commutator \([a, b] = ab - 53 ba\). In particular, given a \(K\)-vector space \(V\) we may view the 54 \(K\)-algebra \(\operatorname{End}(V)\) as a Lie algebra, which we call 55 \(\mathfrak{gl}(V)\). We may also regard the Lie algebra \(\mathfrak{gl}_n(K) 56 = \mathfrak{gl}(K^n)\) as the space of \(n \times n\) matrices with 57 coefficients in \(K\). 58 \end{example} 59 60 \begin{example}\label{ex:gln-inclusions} 61 Let \(n \le m\). Then the map 62 \begin{align*} 63 \mathfrak{gl}_n(K) & \to \mathfrak{gl}_m(K) \\ 64 X & \mapsto \begin{pmatrix} X & 0 \\ 0 & 0 \end{pmatrix} 65 \end{align*} 66 is a homomorphism of Lie algebras. 67 \end{example} 68 69 While straightforward enough, I always found the definition of a Lie algebra 70 unconvincing on its own. Specifically, the Jacobi identity can look very alien 71 to someone who has never ventured outside of the realms of associativity. 72 Traditional abstract algebra courses offer little in the way of a motivation 73 for studying non-associative algebras in general. Why should we drop the 74 assumption of associativity if every example of an algebraic structure we have 75 ever seen is an associative one? Instead, the most natural examples of Lie 76 algebras often come from an entirely different field: geometry. 77 78 Here the meaning of \emph{geometry} is somewhat vague. Topics such as 79 differential and algebraic geometry are prominently featured, but examples 80 from fields such as the theory of differential operators and \(D\)-modules also 81 show up a lot in the theory of representations -- which we will soon discuss. 82 Perhaps one of the most fundamental themes of the study of Lie algebras is 83 their relationship with groups, specially in geometric contexts. We will now 84 provide a brief description of this relationship through a series of examples. 85 86 \begin{example}\index{Lie algebra!Lie algebra of derivations} 87 Let \(A\) be an associative \(K\)-algebra and \(\operatorname{Der}(A)\) be 88 the space of all derivations on \(A\) -- i.e. all linear maps \(D : A \to A\) 89 satisfying the Leibniz rule \(D(a \cdot b) = a \cdot D b + (D a) \cdot b\). 90 The commutator \([D, D']\) of two derivations \(D, D' \in 91 \operatorname{Der}(A)\) in the ring \(\operatorname{End}(A)\) of \(K\)-linear 92 endomorphisms of \(A\) is a derivation. Hence \(\operatorname{Der}(A)\) is a 93 Lie algebra. 94 \end{example} 95 96 One specific instance of this last example is\dots 97 98 \begin{example}\index{Lie algebra!Lie algebra of vector fields} 99 Given a smooth manifold \(M\), the space \(\mathfrak{X}(M)\) of all smooth 100 vector fields is canonically identified with \(\operatorname{Der}(M) = 101 \operatorname{Der}(C^\infty(M))\) -- where a field \(X \in \mathfrak{X}(M)\) 102 is identified with the map \(C^\infty(M) \to C^\infty(M)\) which takes a 103 function \(f \in C^\infty(M)\) to its derivative in the direction of \(X\). 104 This gives \(\mathfrak{X}(M)\) the structure of a Lie algebra over 105 \(\mathbb{R}\). 106 \end{example} 107 108 \begin{example}\label{ex:lie-alg-of-lie-grp}\index{Lie algebra!Lie algebra of a Lie group} 109 Given a Lie group \(G\) -- i.e. a smooth manifold endowed with smooth group 110 operations -- we call \(X \in \mathfrak{X}(G)\) left invariant if \((d 111 \ell_g)_1 X_1 = X_g\) for all \(g \in G\), where \(\ell_g : G \to G\) denotes 112 the left translation by \(g\). The commutator of invariant fields is 113 invariant, so the space \(\mathfrak{g} = \operatorname{Lie}(G)\) of all 114 invariant vector fields has the structure of a Lie algebra over 115 \(\mathbb{R}\) with bracket given by the usual commutator of fields. Notice 116 that an invariant field \(X\) is completely determined by \(X_1 \in T_1 G\). 117 Hence there is a linear isomorphism \(\mathfrak{g} \isoto T_1 G\). In 118 particular, \(\mathfrak{g}\) is finite-dimensional. 119 \end{example} 120 121 We should point out that the Lie algebra \(\mathfrak{g}\) of a complex Lie 122 group \(G\) -- i.e. a complex manifold endowed with holomorphic group 123 operations -- has the natural structure of a complex Lie algebra. Indeed, every 124 left invariant field \(X \in \mathfrak{X}(G)\) is holomorphic, so 125 \(\mathfrak{g}\) is a (complex) subspace of the complex vector space of 126 holomorphic vector fields over \(G\). There is also an algebraic analogue of 127 this last construction. 128 129 \begin{example}\index{Lie algebra!Lie algebra of an algebraic group} 130 Let \(G\) be an affine algebraic \(K\)-group -- i.e. an affine variety over 131 \(K\) with rational group operations -- and \(K[G]\) denote the ring of 132 regular functions \(G \to K\). We call a derivation \(D : K[G] \to K[G]\) 133 left invariant if \(D(g \cdot f) = g \cdot D f\) for all \(g \in G\) and \(f 134 \in K[G]\) -- where the action of \(G\) on \(K[G]\) is given by \((g \cdot 135 f)(h) = f(g^{-1} h)\). The commutator of left invariant derivations is 136 invariant too, so the space \(\operatorname{Lie}(G) = 137 \operatorname{Der}(G)^G\) of invariant derivations in \(K[G]\) has the 138 structure of a Lie algebra over \(K\) with bracket given by the commutator 139 of derivations. Again, \(\operatorname{Lie}(G)\) is isomorphic to the Zariski 140 tangent space \(T_1 G\), which is finite-dimensional. 141 \end{example} 142 143 \begin{example} 144 The Lie algebra \(\operatorname{Lie}(\operatorname{GL}_n(K))\) is canonically 145 isomorphic to the Lie algebra \(\mathfrak{gl}_n(K)\). Likewise, the Lie 146 algebra \(\operatorname{Lie}(\operatorname{SL}_n(K))\) is canonically 147 isomorphic to the Lie algebra \(\mathfrak{sl}_n(K)\) of traceless \(n \times 148 n\) matrices. 149 \[ 150 \mathfrak{sl}_n(K) 151 = \{ X \in \mathfrak{gl}_n(K) : \operatorname{Tr} X = 0 \} 152 \] 153 \end{example} 154 155 \begin{example}\label{ex:sl2-basis} 156 The elements 157 \begin{align*} 158 e & = \begin{pmatrix} 0 & 1 \\ 0 & 0 \end{pmatrix} & 159 f & = \begin{pmatrix} 0 & 0 \\ 1 & 0 \end{pmatrix} & 160 h & = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} 161 \end{align*} 162 form a basis for \(\mathfrak{sl}_2(K)\) and are subject to the following 163 relations. 164 \begin{align*} 165 [e, f] & = h & [h, f] & = -2 f & [h, e] = 2 e 166 \end{align*} 167 \end{example} 168 169 \begin{example}\label{ex:sp2n} 170 The Lie algebra of the affine algebraic group 171 \[ 172 \operatorname{Sp}_{2 n}(K) 173 = \{ 174 g \in \operatorname{GL}_{2 n}(K) : 175 \omega(g \cdot v, g \cdot w) = \omega(v, w) \, \forall v, w \in K^{2n} 176 \} 177 \] 178 is canonically isomorphic to the Lie algebra 179 \[ 180 \mathfrak{sp}_{2 n}(K) = 181 \left\{ 182 \begin{pmatrix} 183 X & Y \\ 184 Z & -X^\top 185 \end{pmatrix} 186 : X, Y, Z \in \mathfrak{gl}_n(K), Y = Y^\top, Z = Z^\top 187 \right\}, 188 \] 189 with bracket given by the usual commutator of matrices -- where 190 \[ 191 \omega( 192 (v_1, \ldots, v_n, \dot v_1, \ldots, \dot v_n), 193 (w_1, \ldots, w_n, \dot w_1, \ldots, \dot w_n) 194 ) 195 = v_1 \dot w_1 + \cdots + v_n \dot w_n 196 - \dot v_1 w_1 - \cdots - \dot v_n w_n 197 \] 198 is, of course, the standard symplectic form of \(K^{2n}\). 199 \end{example} 200 201 It is important to point out that the construction of the Lie algebra 202 \(\mathfrak{g}\) of a Lie group \(G\) in Example~\ref{ex:lie-alg-of-lie-grp} is 203 functorial. Specifically, one can show the derivative \(d f_1 : \mathfrak{g} 204 \cong T_1 G \to T_1 H \cong \mathfrak{h}\) of a smooth group homomorphism \(f : 205 G \to H\) is a homomorphism of Lie algebras, and the chain rule implies \(d (f 206 \circ g)_1 = d f_1 \circ d g_1\). This is known as the \emph{the Lie functor} 207 \(\operatorname{Lie} : \mathbf{LieGrp} \to \mathbb{R}\text{-}\mathbf{LieAlg}\) 208 between the category of Lie groups and smooth group homomorphisms and the 209 category of Lie algebras. 210 211 This goes to show Lie algebras are invariants of Lie groups. What is perhaps 212 more surprising is the fact that, in certain contexts, Lie algebras are perfect 213 invariants. Even more so\dots 214 215 \begin{theorem}[Lie]\label{thm:lie-theorems} 216 The restriction \(\operatorname{Lie} : \mathbf{LieGrp}_{\operatorname{simpl}} 217 \to \mathbb{R}\text{-}\mathbf{LieAlg}\) of the Lie functor to the full 218 subcategory of simply connected Lie groups is an equivalence of categories 219 onto the full subcategory of finite-dimensional real Lie algebras. 220 \end{theorem} 221 222 This last theorem is a direct corollary of the so called \emph{first and third 223 fundamental Lie Theorems}. Lie's first Theorem establishes that if \(G\) is a 224 simply connected Lie group and \(H\) is a connected Lie group then the induced 225 map \(\operatorname{Hom}(G, H) \to \operatorname{Hom}(\mathfrak{g}, 226 \mathfrak{h})\) is bijective, which implies the Lie functor is fully faithful. 227 On the other hand, Lie's third Theorem states that every finite-dimensional 228 real Lie algebra is the Lie algebra of a simply connected Lie group -- i.e. the 229 Lie functor is essentially surjective. 230 231 This goes to show that the relationship between Lie groups and Lie algebras is 232 deeper than the fact they share a name: in a very strong sense, studying simply 233 connected Lie groups is \emph{precisely} the same as studying 234 finite-dimensional Lie algebras. Such a vital connection between apparently 235 distant subjects is bound to produce interesting results. Indeed, the passage 236 from the geometric setting to its algebraic counterpart and vice-versa has 237 proven itself a fruitful one. 238 239 This correspondence can be extended to the complex case too. In other words, 240 the Lie functor \(\mathbf{CLieGrp}_{\operatorname{simpl}} \to 241 \mathbb{C}\text{-}\mathbf{LieAlg}\) is also an equivalence of categories 242 between the category of simply connected complex Lie groups and the full 243 subcategory of finite-dimensional complex Lie algebras. The situation is more 244 delicate in the algebraic case. For instance, consider the complex Lie algebra 245 homomorphism 246 \begin{align*} 247 f : \mathbb{C} & \to \mathfrak{sl}_2(\mathbb{C}) \\ 248 \lambda & \mapsto \lambda h = 249 \begin{pmatrix} \lambda & 0 \\ 0 & - \lambda \end{pmatrix} 250 \end{align*} 251 252 Since \(\mathfrak{sl}_2(\mathbb{C}) = 253 \operatorname{Lie}(\operatorname{SL}_2(\mathbb{C}))\) and 254 \(\operatorname{SL}_2(\mathbb{C})\) is simply connected, we know there exists a 255 unique holomorphic group homomorphism \(g : \mathbb{C} \to 256 \operatorname{SL}_2(\mathbb{C})\) between the affine line \(\mathbb{C}\) and 257 the complex \emph{algebraic} group \(\operatorname{SL}_2(\mathbb{C})\) such 258 that \(f = d g_1\). Indeed, this homomorphism is 259 \begin{align*} 260 g : \mathbb{C} & \to \operatorname{SL}_2(\mathbb{C}) \\ 261 \lambda & \mapsto \operatorname{exp}(\lambda h) = 262 \begin{pmatrix} e^\lambda & 0 \\ 0 & e^{-\lambda} \end{pmatrix}, 263 \end{align*} 264 which is not a rational map. It then follows from the uniqueness of \(g\) that 265 there is no rational group homomorphism \(\mathbb{C} \to 266 \operatorname{SL}_2(\mathbb{C})\) whose derivative at the identity is \(f\). 267 268 In particular, the Lie functor 269 \(\mathbb{C}\text{-}\mathbf{Grp}_{\operatorname{simpl}} \to 270 \mathbb{C}\text{-}\mathbf{LieAlg}\) -- between the category 271 \(\mathbb{C}\text{-}\mathbf{Grp}_{\operatorname{simpl}}\) of simply connected 272 complex algebraic groups and the category of complex Lie algebras -- fails to 273 be full. Similarly, the functor 274 \(\mathbb{C}\text{-}\mathbf{Grp}_{\operatorname{simpl}} \to 275 \mathbb{C}\text{-}\mathbf{LieAlg}\) is \emph{not} essentially surjective onto 276 the subcategory of finite-dimensional algebras: every finite-dimensional 277 complex Lie algebra is isomorphic to the Lie algebra of a unique simply 278 connected complex Lie group, but there are simply connected complex Lie groups 279 which are not algebraic groups. Nevertheless, Lie algebras are still powerful 280 invariants of algebraic groups. An interesting discussion of some of these 281 delicacies can be found in sixth section of \cite[ch.~II]{demazure-gabriel}. 282 283 All in all, there is a profound connection between groups and 284 finite-dimensional Lie algebras throughout multiple fields. While perhaps 285 unintuitive at first, the advantages of working with Lie algebras over their 286 group-theoretic counterparts are numerous. First, Lie algebras allow us to 287 avoid much of the delicacies of geometric objects such as real and complex Lie 288 groups. Even when working without additional geometric considerations, groups 289 can be complicated beasts themselves. They are, after all, nonlinear objects. 290 On the other hand, Lie algebras are linear by nature, which makes them much 291 more flexible than groups. 292 293 Having thus hopefully established that Lie algebras are interesting, we are now 294 ready to dive deeper into them. We begin by analyzing some of their most basic 295 properties. 296 297 \section{Lie Algebras} 298 299 However bizarre Lie algebras may seem at a first glance, they actually share a 300 lot a structural features with their associative counterparts. For instance, it 301 is only natural to define\dots 302 303 \begin{definition}\index{Lie subalgebra}\index{Lie subalgebra!ideals} 304 Given a Lie algebra \(\mathfrak{g}\), a subspace \(\mathfrak{h} \subset 305 \mathfrak{g}\) is called \emph{a subalgebra of \(\mathfrak{g}\)} if \([X, Y] 306 \in \mathfrak{h}\) for all \(X, Y \in \mathfrak{h}\). A subalgebra 307 \(\mathfrak{a} \subset \mathfrak{g}\) is called \emph{an ideal of 308 \(\mathfrak{g}\)} if \([X, Y] \in \mathfrak{a}\) for all \(X \in 309 \mathfrak{g}\) and \(Y \in \mathfrak{a}\), in which case we write 310 \(\mathfrak{a} \normal \mathfrak{g}\). 311 \end{definition} 312 313 \begin{note} 314 In the context of associative algebras, it is usual practice to distinguish 315 between \emph{left ideals} and \emph{right ideals}. This is not necessary 316 when dealing with Lie algebras, however, since any ``left ideal'' of a Lie 317 algebra is also a ``right ideal'': given \(\mathfrak{a} \normal 318 \mathfrak{g}\), \([Y, X] = - [X, Y] \in \mathfrak{a}\) for all \(X \in 319 \mathfrak{g}\) and \(Y \in \mathfrak{a}\). 320 \end{note} 321 322 \begin{example} 323 Let \(f : \mathfrak{g} \to \mathfrak{h}\) be a homomorphism between Lie 324 algebras \(\mathfrak{g}\) and \(\mathfrak{h}\). Then \(\ker f \subset 325 \mathfrak{g}\) and \(\operatorname{im} f \subset \mathfrak{h}\) are 326 subalgebras. Furthermore, \(\ker f \normal \mathfrak{g}\). 327 \end{example} 328 329 \begin{example} 330 Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be a Lie algebras over \(K\). 331 Then the space \(\mathfrak{g}_1 \oplus \mathfrak{g}_2\) is a Lie algebra with 332 bracket 333 \[ 334 [X_1 + X_2, Y_1 + Y_2] = [X_1, Y_1] + [X_2, Y_2], 335 \] 336 and \(\mathfrak{g}_1, \mathfrak{g}_2 \normal \mathfrak{g}_1 \oplus 337 \mathfrak{g}_2\). 338 \end{example} 339 340 \begin{example} 341 Let \(G\) be an affine algebraic \(K\)-group and \(H \subset G\) be a 342 connected closed subgroup. Denote by \(\mathfrak{g}\) and \(\mathfrak{h}\) 343 the Lie algebras of \(G\) and \(H\), respectively. The inclusion \(H \to G\) 344 induces an injective homomorphism \(\mathfrak{h} \to \mathfrak{g}\). We may 345 thus regard \(\mathfrak{h}\) as a subalgebra of \(\mathfrak{g}\). In 346 addition, \(\mathfrak{h} \normal \mathfrak{g}\) if, and only if \(H \normal 347 G\). 348 \end{example} 349 350 There is also a natural analogue of quotients. 351 352 \begin{definition} 353 Given a Lie algebra \(\mathfrak{g}\) and \(\mathfrak{a} \normal 354 \mathfrak{g}\), the space \(\mfrac{\mathfrak{g}}{\mathfrak{a}}\) has the 355 natural structure of a Lie algebra over \(K\), where 356 \[ 357 [X + \mathfrak{a}, Y + \mathfrak{a}] = [X, Y] + \mathfrak{a} 358 \] 359 \end{definition} 360 361 \begin{proposition} 362 Given a Lie algebra \(\mathfrak{g}\) and \(\mathfrak{a} \normal 363 \mathfrak{g}\), every homomorphism of Lie algebras \(f : \mathfrak{g} \to 364 \mathfrak{h}\) such that \(\mathfrak{a} \subset \ker f\) uniquely factors 365 through the projection \(\mathfrak{g} \to 366 \mfrac{\mathfrak{g}}{\mathfrak{a}}\). 367 \begin{center} 368 \begin{tikzcd} 369 \mathfrak{g} \rar{f} \dar & \mathfrak{h} \\ 370 \mfrac{\mathfrak{g}}{\mathfrak{a}} \arrow[dotted]{ur} & 371 \end{tikzcd} 372 \end{center} 373 \end{proposition} 374 375 \begin{definition}\index{Lie algebra!Abelian Lie algebra} 376 A Lie algebra \(\mathfrak{g}\) is called \emph{Abelian} if \([X, Y] = 0\) 377 for all \(X, Y \in \mathfrak{g}\). 378 \end{definition} 379 380 \begin{example} 381 Let \(G\) be a connected algebraic \(K\)-group and \(\mathfrak{g}\) be its 382 Lie algebra. Then \(G\) is Abelian if, and only if \(\mathfrak{g}\) is 383 Abelian. 384 \end{example} 385 386 \begin{note} 387 Notice that an Abelian Lie algebra is determined by its dimension. Indeed, 388 any linear map \(\mathfrak{g} \to \mathfrak{h}\) between Abelian Lie algebras 389 \(\mathfrak{g}\) and \(\mathfrak{h}\) is a homomorphism of Lie algebras. In 390 particular, any linear isomorphism \(\mathfrak{g} \isoto K^n\) -- where 391 \(K^n\) is endowed with the trivial bracket \([v, w] = 0\), \(v, w \in K^n\) 392 -- is an isomorphism of Lie algebras for Abelian \(\mathfrak{g}\). 393 \end{note} 394 395 \begin{example}\index{Lie algebra!center} 396 Let \(\mathfrak{g}\) be a Lie algebra and \(\mathfrak{z} = \{ X \in 397 \mathfrak{g} : [X, Y] = 0, Y \in \mathfrak{g}\}\). Then \(\mathfrak{z}\) is 398 an Abelian ideal of \(\mathfrak{g}\), known as \emph{the center of 399 \(\mathfrak{z}\)}. 400 \end{example} 401 402 Due to their relationship with Lie groups and algebraic groups, Lie algebras 403 also share structural features with groups. For example\dots 404 405 \begin{definition}\index{Lie algebra!solvable Lie algebra} 406 A Lie algebra \(\mathfrak{g}\) is called \emph{solvable} if its derived 407 series 408 \[ 409 \mathfrak{g} 410 \supseteq [\mathfrak{g}, \mathfrak{g}] 411 \supseteq [[\mathfrak{g}, \mathfrak{g}], [\mathfrak{g}, \mathfrak{g}]] 412 \supseteq 413 [ 414 [[\mathfrak{g}, \mathfrak{g}], [\mathfrak{g}, \mathfrak{g}]], 415 [[\mathfrak{g}, \mathfrak{g}], [\mathfrak{g}, \mathfrak{g}]] 416 ] 417 \supseteq \cdots 418 \] 419 converges to \(0\) in finite time. 420 \end{definition} 421 422 \begin{example} 423 Let \(G\) be a connected affine algebraic \(K\)-group and \(\mathfrak{g}\) be 424 its Lie algebra. Then \(G\) is solvable if, and only if \(\mathfrak{g}\) is. 425 \end{example} 426 427 \begin{definition}\index{Lie algebra!nilpotent Lie algebra} 428 A Lie algebra \(\mathfrak{g}\) is called \emph{nilpotent} if its lower 429 central series 430 \[ 431 \mathfrak{g} 432 \supseteq [\mathfrak{g}, \mathfrak{g}] 433 \supseteq [\mathfrak{g}, [\mathfrak{g}, \mathfrak{g}]] 434 \supseteq [\mathfrak{g}, [\mathfrak{g}, [\mathfrak{g}, \mathfrak{g}]]] 435 \supseteq \cdots 436 \] 437 converges to \(0\) in finite time. 438 \end{definition} 439 440 \begin{example} 441 Let \(G\) be a connected affine algebraic \(K\)-group and \(\mathfrak{g}\) be 442 its Lie algebra. Then \(G\) is nilpotent if, and only if \(\mathfrak{g}\) is. 443 \end{example} 444 445 Other interesting classes of Lie algebras are the so called \emph{simple} and 446 \emph{semisimple} Lie algebras. 447 448 \begin{definition}\index{simple!Lie algebra}\index{Lie algebra!simple Lie algebra} 449 A non-Abelian Lie algebra \(\mathfrak{s}\) over \(K\) is called \emph{simple} 450 if its only ideals are \(0\) and \(\mathfrak{s}\). 451 \end{definition} 452 453 \begin{example} 454 The Lie algebra \(\mathfrak{sl}_2(K)\) is simple. To see this, notice that 455 any ideal \(\mathfrak{a} \normal \mathfrak{sl}_2(K)\) must be stable under 456 the operator \(\operatorname{ad}(h) : \mathfrak{sl}_2(K) \to 457 \mathfrak{sl}_2(K)\) given by \(\operatorname{ad}(h) X = [h, X]\). But 458 Example~\ref{ex:sl2-basis} implies \(\operatorname{ad}(h)\) is 459 diagonalizable, with eigenvalues \(0\) and \(\pm 2\). Hence \(\mathfrak{a}\) 460 must be spanned by some of the eigenvectors \(e, f, h\) of 461 \(\operatorname{ad}(h)\). If \(h \in \mathfrak{a}\), then \([e, h] = - 2 e 462 \in \mathfrak{a}\) and \([f, h] = 2 f \in \mathfrak{a}\), so \(\mathfrak{a} = 463 \mathfrak{sl}_2(K)\). If \(e \in \mathfrak{a}\) then \([f, e] = - h \in 464 \mathfrak{a}\), so again \(\mathfrak{a} = \mathfrak{sl}_2(K)\). Similarly, if 465 \(f \in \mathfrak{a}\) then \([e, f] = h \in \mathfrak{a}\) and 466 \(\mathfrak{a} = \mathfrak{sl}_2(K)\). More generally, the Lie algebra 467 \(\mathfrak{sl}_n(K)\) is simple for each \(n > 1\) -- see the section of 468 \cite[ch. 6]{kirillov} on invariant bilinear forms and the semisimplicity of 469 classical Lie algebras. 470 \end{example} 471 472 \begin{example} 473 The Lie algebras \(\mathfrak{sp}_{2n}(K)\) are simple for all \(n \ge 1\) -- 474 agina, see \cite[ch. 6]{kirillov}. 475 \end{example} 476 477 \begin{definition}\label{thm:sesimple-algebra}\index{semisimple!Lie algebra}\index{Lie algebra!semisimple Lie algebra} 478 A Lie algebra \(\mathfrak{g}\) is called \emph{semisimple} if it is the 479 direct sum of simple Lie algebras. Equivalently, a Lie algebra 480 \(\mathfrak{g}\) is called \emph{semisimple} if it has no nonzero solvable 481 ideals. 482 \end{definition} 483 484 \begin{example} 485 Let \(G\) be a connected affine algebraic \(K\)-group. Then \(G\) is 486 semisimple if, and only if \(\mathfrak{g}\) semisimple. 487 \end{example} 488 489 A slight generalization is\dots 490 491 \begin{definition}\index{Lie algebra!reductive Lie algebra} 492 A Lie algebra \(\mathfrak{g}\) is called \emph{reductive} if \(\mathfrak{g}\) 493 is the direct sum of a semisimple Lie algebra and an Abelian Lie algebra. 494 \end{definition} 495 496 \begin{example} 497 The Lie algebra \(\mathfrak{gl}_n(K)\) is reductive. Indeed, 498 \[ 499 X 500 = 501 \begin{pmatrix} 502 a_{1 1} - \frac{\operatorname{Tr}(X)}{n} & \cdots & a_{1 n} \\ 503 \vdots & \ddots & \vdots \\ 504 a_{n 1} & \cdots & a_{n n} - \frac{\operatorname{Tr}(X)}{n} 505 \end{pmatrix} 506 + 507 \begin{pmatrix} 508 \frac{\operatorname{Tr}(X)}{n} & \cdots & 0 \\ 509 \vdots & \ddots & \vdots \\ 510 0 & \cdots & \frac{\operatorname{Tr}(X)}{n} 511 \end{pmatrix} 512 \] 513 for each matrix \(X = (a_{i j})_{i j}\). In other words, 514 \(\mathfrak{gl}_n(K) = \mathfrak{sl}_n(K) \oplus K \operatorname{Id} \cong 515 \mathfrak{sl}_n(K) \oplus K\). 516 \end{example} 517 518 As suggested by their names, simple and semisimple algebras are quite well 519 behaved when compared with the general case. To a lesser degree, reductive 520 algebras are also unusually well behaved. In the next chapter we will explore 521 the question of why this is the case, but for now we note that we can get 522 semisimple and reductive algebras by modding out by certain ideals, known as 523 \emph{radicals}. 524 525 \begin{definition}\index{Lie algebra!radical} 526 Let \(\mathfrak{g}\) be a finite-dimensional Lie algebra. The sum 527 \(\mathfrak{a} + \mathfrak{b}\) of solvable ideals \(\mathfrak{a}, 528 \mathfrak{b} \normal \mathfrak{g}\) is again a solvable ideal. Hence the sum 529 of all solvable ideals of \(\mathfrak{g}\) is a maximal solvable ideal, known 530 as \emph{the radical \(\mathfrak{rad}(\mathfrak{g})\) of \(\mathfrak{g}\)}. 531 \[ 532 \mathfrak{rad}(\mathfrak{g}) 533 = \sum_{\substack{\mathfrak{a} \normal \mathfrak{g}\\\text{solvable}}} 534 \mathfrak{a} 535 \] 536 \end{definition} 537 538 \begin{definition}\index{Lie algebra!nilradical} 539 Let \(\mathfrak{g}\) be a finite-dimensional Lie algebra. The sum of 540 nilpotent ideals is a nilpotent ideal. Hence the sum of all nilpotent ideals 541 of \(\mathfrak{g}\) is a maximal nilpotent ideal, known as \emph{the 542 nilradical \(\mathfrak{nil}(\mathfrak{g})\) of \(\mathfrak{g}\)}. 543 \[ 544 \mathfrak{nil}(\mathfrak{g}) 545 = \sum_{\substack{\mathfrak{a} \normal \mathfrak{g}\\\text{nilpotent}}} 546 \mathfrak{a} 547 \] 548 \end{definition} 549 550 \begin{proposition}\label{thm:quotients-by-rads} 551 Let \(\mathfrak{g}\) be a Lie algebra. Then 552 \(\mfrac{\mathfrak{g}}{\mathfrak{rad}(\mathfrak{g})}\) is semisimple and 553 \(\mfrac{\mathfrak{g}}{\mathfrak{nil}(\mathfrak{g})}\) is reductive. 554 \end{proposition} 555 556 We have seen in Example~\ref{ex:inclusion-alg-in-lie-alg} that we can pass from 557 an associative algebra \(A\) to a Lie algebra by taking its bracket as the 558 commutator \([a, b] = ab - ba\). We should also not that any homomorphism of 559 \(K\)-algebras \(f : A \to B\) preserves commutators, so that \(f\) is also a 560 homomorphism of Lie algebras. Hence we have a functor \(\operatorname{Lie} : 561 K\text{-}\mathbf{Alg} \to K\text{-}\mathbf{LieAlg}\). We can also go the other 562 direction by embedding a Lie algebra \(\mathfrak{g}\) in an associative 563 algebra, known as \emph{the universal enveloping algebra of \(\mathfrak{g}\)}. 564 565 \begin{definition}\index{universal enveloping algebra} 566 Let \(\mathfrak{g}\) be a Lie algebra and \(T \mathfrak{g} = \bigoplus_n 567 \mathfrak{g}^{\otimes n}\) be its tensor algebra -- i.e. the free 568 \(K\)-algebra generated by the elements of \(\mathfrak{g}\). We call the 569 \(K\)-algebra \(\mathcal{U}(\mathfrak{g}) = \mfrac{T \mathfrak{g}}{I}\) 570 \emph{the universal enveloping algebra of \(\mathfrak{g}\)}, where \(I\) is 571 the left ideal of \(T \mathfrak{g}\) generated by the elements \([X, Y] - (X 572 \otimes Y - Y \otimes X)\). 573 \end{definition} 574 575 Notice there is a canonical homomorphism \(\mathfrak{g} \to 576 \mathcal{U}(\mathfrak{g})\) given by the composition 577 \begin{center} 578 \begin{tikzcd} 579 \mathfrak{g} \rar & 580 T \mathfrak{g} \rar & 581 \mfrac{T \mathfrak{g}}{I} = \mathcal{U}(\mathfrak{g}) 582 \end{tikzcd} 583 \end{center} 584 585 Given \(X_1, \ldots, X_n \in \mathfrak{g}\), we identify \(X_i\) with its image 586 under the inclusion \(\mathfrak{g} \to T \mathfrak{g}\) and we write \(X_1 587 \cdots X_n\) for \((X_1 \otimes \cdots \otimes X_n) + I\). This notation 588 suggests the map \(\mathfrak{g} \to \mathcal{U}(\mathfrak{g})\) is injective, 589 but at this point this is not at all clear -- given that the projection \(T 590 \mathfrak{g} \to \mathcal{U}(\mathfrak{g})\) is not injective. However, we will 591 soon see this is the case. Intuitively, \(\mathcal{U}(\mathfrak{g})\) is the 592 smallest associative \(K\)-algebra containing \(\mathfrak{g}\) as a Lie 593 subalgebra. In practice this means\dots 594 595 \begin{proposition}\label{thm:universal-env-uni-prop} 596 Let \(\mathfrak{g}\) be a Lie algebra and \(A\) be an associative 597 \(K\)-algebra. Then every homomorphism of Lie algebras \(f : \mathfrak{g} \to 598 A\) -- where \(A\) is endowed with the structure of a Lie algebra as in 599 Example~\ref{ex:inclusion-alg-in-lie-alg} -- can be uniquely extended to a 600 homomorphism of algebras \(\mathcal{U}(\mathfrak{g}) \to A\). 601 \begin{center} 602 \begin{tikzcd} 603 \mathfrak{g} \rar{f} \dar & A \\ 604 \mathcal{U}(\mathfrak{g}) \urar[dotted] & 605 \end{tikzcd} 606 \end{center} 607 \end{proposition} 608 609 \begin{proof} 610 Let \(f : \mathfrak{g} \to A\) be a homomorphism of Lie algebras. By the 611 universal property of free algebras, there is a homomorphism of algebras 612 \(\tilde f : T \mathfrak{g} \to A\) such that 613 \begin{center} 614 \begin{tikzcd} 615 \mathfrak{g} \dar \rar{f} & A \\ 616 T \mathfrak{g} \urar[swap, dotted]{\tilde f} & 617 \end{tikzcd} 618 \end{center} 619 620 Since \(f\) is a homomorphism of Lie algebras, 621 \[ 622 \tilde f([X, Y]) 623 = f([X, Y]) 624 = [f(X), f(Y)] 625 = [\tilde f(X), \tilde f(Y)] 626 = \tilde f(X \otimes Y - Y \otimes X) 627 \] 628 for all \(X, Y \in \mathfrak{g}\). Hence \(I = ([X, Y] - (X \otimes Y - Y 629 \otimes X) : X, Y \in \mathfrak{g}) \subset \ker \tilde f\) and therefore 630 \(\tilde f\) factors through the quotient \(\mathcal{U}(\mathfrak{g}) = 631 \mfrac{T \mathfrak{g}}{I}\). 632 \begin{center} 633 \begin{tikzcd} 634 T \mathfrak{g} \rar{\tilde f} \dar & A \\ 635 \mathcal{U}(\mathfrak{g}) \arrow[swap, dotted]{ur}{\bar{\tilde f}} & 636 \end{tikzcd} 637 \end{center} 638 639 Combining the two previous diagrams, we can see that \(\bar{\tilde f}\) is 640 indeed an extension of \(f\). The uniqueness of the extension then follows 641 from the uniqueness of \(\tilde f\) and \(\bar{\tilde f}\). 642 \end{proof} 643 644 We should point out this construction is functorial. Indeed, if 645 \(f : \mathfrak{g} \to \mathfrak{h}\) is a homomorphism of Lie algebras then 646 Proposition~\ref{thm:universal-env-uni-prop} implies there is a homomorphism of 647 algebras \(\mathcal{U}(f) : \mathcal{U}(\mathfrak{g}) \to 648 \mathcal{U}(\mathfrak{h})\) satisfying 649 \begin{center} 650 \begin{tikzcd} 651 \mathfrak{g} \rar{f} \dar & 652 \mathfrak{h} \rar & 653 \mathcal{U}(\mathfrak{h}) \\ 654 \mathcal{U}(\mathfrak{g}) \arrow[swap, dotted]{urr}{\mathcal{U}(f)} & & 655 \end{tikzcd} 656 \end{center} 657 658 It is important to note, however, that \(\mathcal{U} : K\text{-}\mathbf{LieAlg} 659 \to K\text{-}\mathbf{Alg}\) is not the ``inverse'' of our functor 660 \(K\text{-}\mathbf{Alg} \to K\text{-}\mathbf{LieAlg}\). For instance, if 661 \(\mathfrak{g} = K\) is the \(1\)-dimensional Abelian Lie algebra then 662 \(\mathcal{U}(\mathfrak{g}) \cong K[x]\), which is infinite-dimensional. 663 Nevertheless, Proposition~\ref{thm:universal-env-uni-prop} may be restated 664 using the language of adjoint functors -- as described in \cite{maclane} for 665 instance. 666 667 \begin{corollary} 668 If \(\operatorname{Lie} : K\text{-}\mathbf{Alg} \to 669 K\text{-}\mathbf{LieAlg}\) is the functor described in 670 Example~\ref{ex:inclusion-alg-in-lie-alg}, there is an adjunction 671 \(\operatorname{Lie} \vdash \mathcal{U}\). 672 \end{corollary} 673 674 The structure of \(\mathcal{U}(\mathfrak{g})\) can often be described in terms 675 of the structure of \(\mathfrak{g}\). For instance, \(\mathfrak{g}\) is Abelian 676 if, and only if \(\mathcal{U}(\mathfrak{g})\) is commutative, in which case any 677 basis \(\{X_i\}_i\) for \(\mathfrak{g}\) induces an isomorphism 678 \(\mathcal{U}(\mathfrak{g}) \cong K[x_1, x_2, \ldots, x_i, \ldots]\). More 679 generally, we find\dots 680 681 \begin{theorem}[Poincaré-Birkoff-Witt]\index{PBW Theorem} 682 Let \(\mathfrak{g}\) be a Lie algebra over \(K\) and \(\{X_i\}_i \subset 683 \mathfrak{g}\) be an ordered basis for \(\mathfrak{g}\) -- i.e. a basis 684 indexed by an ordered set. Then \(\{X_{i_1} \cdot X_{i_2} \cdots X_{i_n} : n 685 \ge 0, i_1 \le i_2 \le \cdots \le i_n\}\) is a basis for 686 \(\mathcal{U}(\mathfrak{g})\). 687 \end{theorem} 688 689 This last result is known as \emph{the PBW Theorem}. It is hugely important and 690 will come up again and again throughout these notes. Among other things, it 691 implies\dots 692 693 \begin{corollary} 694 Let \(\mathfrak{g}\) be a Lie algebra over \(K\). Then 695 \(\mathcal{U}(\mathfrak{g})\) is a domain and the inclusion \(\mathfrak{g} 696 \to \mathcal{U}(\mathfrak{g})\) is injective. 697 \end{corollary} 698 699 The PBW Theorem can also be used to compute a series of 700 examples. 701 702 \begin{example} 703 Consider the Lie algebra \(\mathfrak{gl}_n(K)\) and its canonical basis 704 \(\{E_{i j}\}_{i j}\). Even though \(E_{i j} E_{j k} = E_{i k}\) in the 705 associative algebra \(\operatorname{End}(K^n)\), the PBW 706 Theorem implies \(E_{i j} E_{j k} \ne E_{i k}\) in 707 \(\mathcal{U}(\mathfrak{gl}_n(K))\). In general, if \(A\) is an associative 708 \(K\)-algebra then the elements in the image of the inclusion \(A \to 709 \mathcal{U}(A)\) do not satisfy the same relations as the elements of \(A\). 710 \end{example} 711 712 \begin{example} 713 Let \(\mathfrak{g}\) be an Abelian Lie algebra. As previously stated, any 714 choice of basis \(\{X_i\}_i \subset \mathfrak{g}\) induces an isomorphism of 715 algebras \(\mathcal{U}(\mathfrak{g}) \isoto K[x_1, x_2, \ldots, x_i, 716 \ldots]\) which takes \(X_i \in \mathfrak{g}\) to the variable \(x_i \in 717 K[x_1, x_2, \ldots, x_i, \ldots]\). 718 \end{example} 719 720 \begin{example}\label{ex:univ-enveloping-of-sum-is-tensor} 721 Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras over \(K\). We 722 claim that the natural map 723 \begin{align*} 724 f: \mathcal{U}(\mathfrak{g}_1) \otimes_K \mathcal{U}(\mathfrak{g}_2) & 725 \to \mathcal{U}(\mathfrak{g}_1 \oplus \mathfrak{g}_2) \\ 726 u \otimes v & \mapsto u \cdot v 727 \end{align*} 728 is an isomorphism of algebras. Since the elements of \(\mathfrak{g}_1\) 729 commute with the elements of \(\mathfrak{g}_2\) in \(\mathfrak{g}_1 \oplus 730 \mathfrak{g}_2\), a simple calculation shows that \(f\) is indeed a 731 homomorphism of algebras. In addition, the PBW Theorem implies that \(f\) is 732 a linear isomorphism. 733 \end{example} 734 735 The construction of \(\mathcal{U}(\mathfrak{g})\) may seem like a purely 736 algebraic affair, but the universal enveloping algebra of the Lie algebra of a 737 Lie group \(G\) is in fact intimately related with the algebra 738 \(\operatorname{Diff}(G)\) of differential operators \(C^\infty(G) \to 739 C^\infty(G)\) -- i.e. \(\mathbb{R}\)-linear endomorphisms \(C^\infty(G) \to 740 C^\infty(G)\) of finite order, as defined in \cite[ch.~3]{coutinho} for 741 example. Algebras of differential operators and their modules are the subject 742 of the theory of \(D\)-modules, which has seen remarkable progress in the past 743 century. Specifically, we find\dots 744 745 \begin{proposition}\label{thm:geometric-realization-of-uni-env} 746 Let \(G\) be a Lie group and \(\mathfrak{g}\) be its Lie algebra. Denote by 747 \(\operatorname{Diff}(G)^G\) the algebra of \(G\)-invariant differential 748 operators in \(G\) -- i.e. the algebra of all differential operators \(P : 749 C^\infty(G) \to C^\infty(G)\) such that \(g \cdot P f = P (g \cdot f)\) for 750 all \(f \in C^\infty(G)\) and \(g \in G\). There is a canonical isomorphism 751 of algebras \(\mathcal{U}(\mathfrak{g}) \isoto \operatorname{Diff}(G)^G\). 752 \end{proposition} 753 754 \begin{proof} 755 An order \(0\) \(G\)-invariant differential operator in \(G\) is simply 756 multiplication by a constant in \(\mathbb{R}\). A homogeneous order \(1\) 757 \(G\)-invariant differential operator in \(G\) is simply a left invariant 758 derivation \(C^\infty(G) \to C^\infty(G)\). All other \(G\)-invariant 759 differential operators are generated by invariant operators of order \(0\) 760 and \(1\). Hence \(\operatorname{Diff}(G)^G\) is generated by 761 \(\operatorname{Der}(G)^G + \mathbb{R}\) -- here \(\operatorname{Der}(G)^G 762 \subset \operatorname{Der}(G)\) denotes the Lie subalgebra of invariant 763 derivations. 764 765 Now recall that there is a canonical isomorphism of Lie algebras 766 \(\mathfrak{X}(G) \isoto \operatorname{Der}(G)\). This isomorphism takes left 767 invariant fields to left invariant derivations, so it restricts to an 768 isomorphism \(f : \mathfrak{g} \isoto \operatorname{Der}(G)^G \subset 769 \operatorname{Diff}(G)^G\). Since \(f\) is a homomorphism of Lie algebras, it 770 can be extended to an algebra homomorphism \(\tilde f : 771 \mathcal{U}(\mathfrak{g}) \to \operatorname{Diff}(G)^G\). We claim \(\tilde 772 f\) is an isomorphism. 773 774 To see that \(\tilde f\) is injective, it suffices to notice 775 \[ 776 \tilde f(X_1 \cdots X_n) 777 = \tilde f(X_1) \cdots \tilde f(X_n) 778 = f(X_1) \cdots f(X_n) 779 \ne 0 780 \] 781 for all nonzero \(X_1, \ldots, X_n \in \mathfrak{g}\) -- the product of 782 operators of positive order has positive order and is therefore nonzero. 783 Since \(\mathcal{U}(\mathfrak{g})\) is generated by the image of the 784 inclusion \(\mathfrak{g} \to \mathcal{U}(\mathfrak{g})\), this implies \(\ker 785 \tilde f = 0\). Given that \(\operatorname{Diff}(G)^G\) is generated by 786 \(\operatorname{Der}(G)^G + \mathbb{R}\), this also goes to show \(\tilde f\) 787 is surjective. 788 \end{proof} 789 790 As one would expect, the same holds for complex Lie groups and algebraic groups 791 too -- if we replace \(C^\infty(G)\) by \(\mathcal{O}(G)\) and \(K[G]\), 792 respectively. This last proposition has profound implications. For example, it 793 affords us an analytic proof of certain particular cases of the 794 PBW Theorem. Most surprising of all, 795 Proposition~\ref{thm:geometric-realization-of-uni-env} implies 796 \(\mathcal{U}(\mathfrak{g})\)-modules are \emph{precisely} the same as modules 797 over the ring of \(G\)-invariant differential operators -- i.e. 798 \(\operatorname{Diff}(G)^G\)-modules. We can thus use 799 \(\mathcal{U}(\mathfrak{g})\) and its modules to understand the geometry of 800 \(G\). 801 802 Proposition~\ref{thm:geometric-realization-of-uni-env} is in fact only the 803 beginning of a profound connection between the theory of \(D\)-modules and 804 \emph{representation theory}, the latter of which we now explore in the 805 following section. 806 807 \section{Representation Theory} 808 809 First introduced in 1896 by Georg Frobenius in his paper \citetitle{frobenius} 810 \cite{frobenius} in the context of group theory, representation theory is now 811 one of the cornerstones of modern mathematics. In this section we provide a 812 brief overview of basic concepts of the representation theory of Lie algebras. 813 We should stress, however, that the representation theory of Lie algebras is 814 only a small fragment of what is today known as ``representation theory'', 815 which is in general concerned with a diverse spectrum of algebraic and 816 combinatorial structures -- such as groups, quivers and associative algebras. 817 An introductory exploration of some of these structures can be found in 818 \cite{etingof}. 819 820 We begin by noting that any \(\mathcal{U}(\mathfrak{g})\)-module \(M\) may be 821 regarded as a \(K\)-vector space endowed with a ``linear action'' of 822 \(\mathfrak{g}\). Indeed, by restricting the action map 823 \(\mathcal{U}(\mathfrak{g}) \to \operatorname{End}(M)\) to \(\mathfrak{g} 824 \subset \mathcal{U}(\mathfrak{g})\) yields a homomorphism of Lie algebras 825 \(\mathfrak{g} \to \mathfrak{gl}(M) = \operatorname{End}(M)\). In fact 826 Proposition~\ref{thm:universal-env-uni-prop} implies that given a vector space 827 \(M\) there is a one-to-one correspondence between 828 \(\mathcal{U}(\mathfrak{g})\)-module structures for \(M\) and homomorphisms 829 \(\mathfrak{g} \to \mathfrak{gl}(M)\). This leads us to the following 830 definition. 831 832 \begin{definition} 833 Given a Lie algebra \(\mathfrak{g}\) over \(K\), \emph{a representation \(V\) 834 of \(\mathfrak{g}\)} is a \(K\)-vector space endowed with a homomorphism of 835 Lie algebras \(\rho : \mathfrak{g} \to \mathfrak{gl}(V)\). 836 \end{definition} 837 838 Hence there is a one-to-one correspondence between representations of 839 \(\mathfrak{g}\) and \(\mathcal{U}(\mathfrak{g})\)-modules. 840 841 \begin{example}\index{\(\mathfrak{g}\)-module!trivial module} 842 Given a Lie algebra \(\mathfrak{g}\), the zero map \(0 : \mathfrak{g} \to K\) 843 gives \(K\) the structure of a representation of \(\mathfrak{g}\), known as 844 \emph{the trivial representation}. 845 \end{example} 846 847 \begin{example}\index{\(\mathfrak{g}\)-module!adjoint module} 848 Given a Lie algebra \(\mathfrak{g}\), consider the homomorphism 849 \(\operatorname{ad} : \mathfrak{g} \to \mathfrak{gl}(\mathfrak{g})\) given by 850 \(\operatorname{ad}(X) Y = [X, Y]\). This gives \(\mathfrak{g}\) the 851 structure of a representation of \(\mathfrak{g}\), known as \emph{the adjoint 852 representation}. 853 \end{example} 854 855 \begin{example}\index{\(\mathfrak{g}\)-module!regular module} 856 Given a Lie algebra \(\mathfrak{g}\), the map \(\rho : \mathfrak{g} \to 857 \mathfrak{gl}(\mathcal{U}(\mathfrak{g}))\) given by left multiplication 858 endows \(\mathcal{U}(\mathfrak{g})\) with the structure of a representation 859 of \(\mathfrak{g}\), known as \emph{the regular representation of 860 \(\mathfrak{g}\)}. 861 \[ 862 \arraycolsep=1.4pt 863 \begin{array}[t]{rl} 864 \rho : \mathfrak{g} & \to \mathfrak{gl}(\mathcal{U}(\mathfrak{g})) \\ 865 X & \mapsto 866 \begin{array}[t]{rl} 867 \rho(X) : \mathcal{U}(\mathfrak{g}) & \to \mathcal{U}(\mathfrak{g}) \\ 868 u & \mapsto X \cdot u 869 \end{array} 870 \end{array} 871 \] 872 \end{example} 873 874 \begin{example}\index{\(\mathfrak{g}\)-module!natural module} 875 Given a subalgebra \(\mathfrak{g} \subset \mathfrak{gl}_n(K)\), the inclusion 876 \(\mathfrak{g} \to \mathfrak{gl}_n(K)\) endows \(K^n\) with the structure of 877 a representation of \(\mathfrak{g}\), known as \emph{the natural 878 representation of \(\mathfrak{g}\)}. 879 \end{example} 880 881 When the map \(\rho : \mathfrak{g} \to \mathfrak{gl}(V)\) is clear from context 882 it is usual practice to denote the \(K\)-endomorphism \(\rho(X) : V \to V\), 883 \(X \in \mathfrak{g}\), simply by \(X\!\restriction_V\). This leads us to the 884 natural notion of \emph{transformations} between representations. 885 886 \begin{definition} 887 Given a Lie algebra \(\mathfrak{g}\) and two representations \(V\) and \(W\) 888 of \(\mathfrak{g}\), we call a \(K\)-linear map \(f : V \to W\) \emph{an 889 intertwining operator}, or \emph{an intertwiner}, if it commutes with the 890 action of \(\mathfrak{g}\) on \(V\) and \(W\), in the sense that the diagram 891 \begin{center} 892 \begin{tikzcd} 893 V \rar{f} \dar[swap]{X\!\restriction_V} & W \dar{X\!\restriction_W} \\ 894 V \rar[swap]{f} & W 895 \end{tikzcd} 896 \end{center} 897 commutes for all \(X \in \mathfrak{g}\). We denote the space of all 898 intertwiners \(V \to W\) by \(\operatorname{Hom}_{\mathfrak{g}}(V, W)\) -- as 899 opposed the space \(\operatorname{Hom}(V, W)\) of all \(K\)-linear maps 900 \(V \to W\). 901 \end{definition} 902 903 The collection of representations of a fixed Lie algebra \(\mathfrak{g}\) thus 904 forms a category, which we call \(\mathbf{Rep}(\mathfrak{g})\). This allow us 905 formulate the correspondence between representations of \(\mathfrak{g}\) and 906 \(\mathcal{U}(\mathfrak{g})\)-modules in more precise terms. 907 908 \begin{proposition} 909 There is a natural isomorphism of categories \(\mathbf{Rep}(\mathfrak{g}) 910 \isoto \mathcal{U}(\mathfrak{g})\text{-}\mathbf{Mod}\). 911 \end{proposition} 912 913 \begin{proof} 914 We have seen that given a \(K\)-vector space \(M\) there is a one-to-one 915 correspondence between \(\mathfrak{g}\)-representation structures for \(M\) 916 -- i.e. homomorphisms \(\mathfrak{g} \to \mathfrak{gl}(M)\) -- and 917 \(\mathcal{U}(\mathfrak{g})\)-module structures for \(M\) -- i.e. 918 homomorphisms \(\mathcal{U}(\mathfrak{g}) \to \operatorname{End}(M)\). This 919 gives us a surjective map that takes objects in 920 \(\mathbf{Rep}(\mathfrak{g})\) to objects in 921 \(\mathcal{U}(\mathfrak{g})\text{-}\mathbf{Mod}\). 922 923 As for the corresponding maps \(\operatorname{Hom}_{\mathfrak{g}}(M, N) \to 924 \operatorname{Hom}_{\mathcal{U}(\mathfrak{g})}(M, N)\), it suffices to note 925 that a \(K\)-linear map between representations \(M\) and \(N\) is an 926 intertwiner if, and only if it is a homomorphism of 927 \(\mathcal{U}(\mathfrak{g})\)-modules. Our functor thus takes an intertwiner 928 \(M \to N\) to itself. It should then be clear that our functor 929 \(\mathbf{Rep}(\mathfrak{g}) \to \mathfrak{g}\text{-}\mathbf{Mod}\) is 930 invertible. 931 \end{proof} 932 933 The language of representation is thus equivalent to that of 934 \(\mathcal{U}(\mathfrak{g})\)-modules, which we call 935 \emph{\(\mathfrak{g}\)-modules}. Correspondingly, we refer to the category 936 \(\mathcal{U}(\mathfrak{g})\text{-}\mathbf{Mod}\) as 937 \(\mathfrak{g}\text{-}\mathbf{Mod}\). The terms 938 \emph{\(\mathfrak{g}\)-submodule} and \emph{\(\mathfrak{g}\)-homomorphism} 939 should also be self-explanatory. To avoid any confusion, we will, for the most 940 part, exclusively use the language of \(\mathfrak{g}\)-modules. It should be 941 noted, however, that both points of view are profitable. 942 943 For starters, the notation for \(\mathfrak{g}\)-modules is much cleaner than 944 that of representations: it is much easier to write ``\(X \cdot m\)'' than 945 ``\((\rho(X))(m)\)'' or even ``\(X\!\restriction_M(m)\)''. By using the 946 language of \(\mathfrak{g}\)-modules we can also rely on the general theory of 947 modules over associative algebras -- which we assume the reader is already 948 familiarized with. On the other hand, it is usually easier to express geometric 949 considerations in terms of the language representations, particularly in group 950 representation theory. 951 952 Often times it is easier to define a \(\mathfrak{g}\)-module \(M\) in terms of 953 the corresponding map \(\mathfrak{g} \to \mathfrak{gl}(M)\) -- this is 954 technique we will use throughout the text. In general, the equivalence between 955 both languages makes it clear that to understand the action of 956 \(\mathcal{U}(\mathfrak{g})\) on \(M\) it suffices to understand the action of 957 \(\mathfrak{g} \subset \mathcal{U}(\mathfrak{g})\). For instance, for defining 958 a \(\mathfrak{g}\)-module \(M\) it suffices to define the action of each \(X 959 \in \mathfrak{g}\) and verify this action respects the commutator relations of 960 \(\mathfrak{g}\) -- indeed, \(\mathfrak{g}\) generates 961 \(\mathcal{U}(\mathfrak{g})\) as an algebra, and the only relations between 962 elements of \(\mathfrak{g}\) are the ones derived from the commutator 963 relations. 964 965 \begin{example}\label{ex:sl2-polynomial-rep} 966 The space \(K[x, y]\) is a \(\mathfrak{sl}_2(K)\)-module with 967 \begin{align*} 968 e \cdot p & = x \frac{\mathrm{d}}{\mathrm{d}y} p & 969 f \cdot p & = y \frac{\mathrm{d}}{\mathrm{d}x} p & 970 h \cdot p & = 971 \left( 972 x \frac{\mathrm{d}}{\mathrm{d}x} - y \frac{\mathrm{d}}{\mathrm{d}y} 973 \right) p 974 \end{align*} 975 \end{example} 976 977 \begin{example} 978 Given a Lie algebra \(\mathfrak{g}\) and \(\mathfrak{g}\)-modules \(M\) and 979 \(N\), the space \(\operatorname{Hom}(M, N)\) of \(K\)-linear maps \(M \to 980 N\) is a \(\mathfrak{g}\)-module where \((X \cdot f)(m) = X \cdot f(m) - f(X 981 \cdot m)\) for all \(X \in \mathfrak{g}\) and \(f \in \operatorname{Hom}(M, 982 N)\). In particular, if we take \(N = K\) the trivial 983 \(\mathfrak{g}\)-module, we can view \(M^*\) -- the dual of \(M\) in the 984 category of \(K\)-vector spaces -- as a \(\mathfrak{g}\)-module where \((X 985 \cdot f)(m) = - f(X \cdot m)\) for all \(f : M \to K\). 986 \end{example} 987 988 The fundamental problem of representation theory is a simple one: classifying 989 all representations of a given Lie algebra up to isomorphism. However, 990 understanding the relationship between representations is also of huge 991 importance. In other words, to understand the whole of 992 \(\mathfrak{g}\text{-}\mathbf{Mod}\) we need to study the collective behavior 993 of representations -- as opposed to individual examples. For instance, we may 994 consider \(\mathfrak{g}\)-submodules, quotients and tensor products. 995 996 \begin{example}\label{ex:sl2-polynomial-subrep} 997 Let \(K[x, y]\) be the \(\mathfrak{sl}_2(K)\)-module as in 998 Example~\ref{ex:sl2-polynomial-rep}. Since \(e\), \(f\) and \(h\) all 999 preserve the degree of monomials, the space \(K[x, y]^{(d)} = \bigoplus_{k + 1000 \ell = d} K x^k y^\ell\) of homogeneous polynomials of degree \(d\) is a 1001 finite-dimensional \(\mathfrak{sl}_2(K)\)-submodule of \(K[x, y]\). 1002 \end{example} 1003 1004 \begin{example} 1005 Given a Lie algebra \(\mathfrak{g}\), a \(\mathfrak{g}\)-module \(M\) and \(m 1006 \in M\), the subspace \(\mathcal{U}(\mathfrak{g}) \cdot m = \{ u \cdot m : u 1007 \in \mathcal{U}(\mathfrak{g}) \}\) is a \(\mathfrak{g}\)-submodule of \(M\), 1008 which we call \emph{the submodule generated by \(m\)}. 1009 \end{example} 1010 1011 \begin{example}\index{\(\mathfrak{g}\)-module!tensor product} 1012 Given a Lie algebra \(\mathfrak{g}\) and \(\mathfrak{g}\)-modules \(M\) and 1013 \(N\), the space \(M \otimes N = M \otimes_K N\) is a \(\mathfrak{g}\)-module 1014 where \(X \cdot (m \otimes n) = X \cdot m \otimes n + m \otimes X \cdot n\). 1015 The exterior and symmetric products \(M \wedge N\) and \(M \odot N\) are both 1016 quotients of \(M \otimes N\) by \(\mathfrak{g}\)-submodules. In particular, 1017 the exterior and symmetric powers \(\wedge^r M\) and \(\operatorname{Sym}^r 1018 M\) are \(\mathfrak{g}\)-modules. 1019 \end{example} 1020 1021 \begin{note} 1022 We would like to stress that the monoidal structure of 1023 \(\mathfrak{g}\text{-}\mathbf{Mod}\) we've just described is \emph{not} given 1024 by the usual tensor product of modules. In other words, \(M \otimes N\) is 1025 not the same as \(M \otimes_{\mathcal{U}(\mathfrak{g})} N\). 1026 \end{note} 1027 1028 It is also interesting to consider the relationship between representations of 1029 separate algebras. In particular, we may define\dots 1030 1031 \begin{example}\index{\(\mathfrak{g}\)-module!restriction} 1032 Let \(\mathfrak{g}\) be a Lie algebra and \(\mathfrak{h}\) be a subalgebra. 1033 Given a \(\mathfrak{g}\)-module \(M\), denote by 1034 \(\operatorname{Res}_{\mathfrak{h}}^{\mathfrak{g}} M = M\) the 1035 \(\mathfrak{h}\)-module where the action of \(\mathfrak{h}\) is given by 1036 restricting the map \(\mathfrak{g} \to \mathfrak{gl}(M)\) to 1037 \(\mathfrak{h}\). Any homomorphism of \(\mathfrak{g}\)-modules \(M \to N\) is 1038 also a homomorphism of \(\mathfrak{h}\)-modules and this construction is 1039 clearly functorial. 1040 \[ 1041 \operatorname{Res}_{\mathfrak{h}}^{\mathfrak{g}} : 1042 \mathfrak{g}\text{-}\mathbf{Mod} \to 1043 \mathfrak{h}\text{-}\mathbf{Mod} 1044 \] 1045 \end{example} 1046 1047 \begin{example} 1048 Given a Lie algebra \(\mathfrak{g}\), the adjoint \(\mathfrak{g}\)-module is 1049 a submodule of the restriction of the adjoint 1050 \(\mathcal{U}(\mathfrak{g})\)-module -- where we consider 1051 \(\mathcal{U}(\mathfrak{g})\) a Lie algebra as in 1052 Example~\ref{ex:inclusion-alg-in-lie-alg}, not as an associative algebra -- 1053 to \(\mathfrak{g}\). 1054 \end{example} 1055 1056 Surprisingly, this functor has a right adjoint. 1057 1058 \begin{example}\index{\(\mathfrak{g}\)-module!induction} 1059 Let \(\mathfrak{g}\) be a Lie algebra and \(\mathfrak{h}\) be a subalgebra. 1060 Given a \(\mathfrak{h}\)-module \(M\), let 1061 \(\operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} M = 1062 \mathcal{U}(\mathfrak{g}) \otimes_{\mathcal{U}(\mathfrak{h})} M\) -- where 1063 the right \(\mathfrak{h}\)-module structure of \(\mathcal{U}(\mathfrak{g})\) 1064 is given by right multiplication. Any \(\mathfrak{h}\)-homomorphism \(f : M 1065 \to N\) induces a \(\mathfrak{g}\)-homomorphism 1066 \(\operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} f = \operatorname{id} 1067 \otimes f : \operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} M \to 1068 \operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} N\) and this construction is 1069 clearly functorial. 1070 \[ 1071 \operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} : 1072 \mathfrak{h}\text{-}\mathbf{Mod} \to \mathfrak{g}\text{-}\mathbf{Mod} 1073 \] 1074 \end{example} 1075 1076 \begin{proposition}\label{thm:frobenius-reciprocity} 1077 Given a Lie algebra \(\mathfrak{g}\), a subalgebra \(\mathfrak{h} \subset 1078 \mathfrak{g}\), a \(\mathfrak{h}\)-module \(M\) and a \(\mathfrak{g}\)-module 1079 \(N\), the map 1080 \[ 1081 \arraycolsep=1.4pt 1082 \begin{array}[t]{rl} 1083 \alpha : 1084 \operatorname{Hom}_{\mathfrak{g}}( 1085 \operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} M, 1086 N 1087 ) & \to 1088 \operatorname{Hom}_{\mathfrak{h}}( 1089 M, 1090 \operatorname{Res}_{\mathfrak{h}}^{\mathfrak{g}} N 1091 ) \\ 1092 f & \mapsto 1093 \begin{array}[t]{rl} 1094 \alpha(f) : M & \to \operatorname{Res}_{\mathfrak{h}}^{\mathfrak{g}} N \\ 1095 m & \mapsto f(1 \otimes m) 1096 \end{array} 1097 \end{array} 1098 \] 1099 is a \(K\)-linear isomorphism. In other words, there is an adjunction 1100 \(\operatorname{Res}_{\mathfrak{h}}^{\mathfrak{g}} \vdash 1101 \operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}}\). 1102 \end{proposition} 1103 1104 \begin{proof} 1105 It suffices to note that the map 1106 \[ 1107 \arraycolsep=1.4pt 1108 \begin{array}[t]{rl} 1109 \beta : 1110 \operatorname{Hom}_{\mathfrak{h}}( 1111 M, 1112 \operatorname{Res}_{\mathfrak{h}}^{\mathfrak{g}} N 1113 ) & \to 1114 \operatorname{Hom}_{\mathfrak{g}}( 1115 \operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} M, 1116 N 1117 ) \\ 1118 f & \mapsto 1119 \begin{array}[t]{rl} 1120 \beta(f) : \operatorname{Ind}_{\mathfrak{h}}^{\mathfrak{g}} M & \to N \\ 1121 u \otimes m & \mapsto u \cdot f(m) 1122 \end{array} 1123 \end{array} 1124 \] 1125 is the inverse of \(\alpha\). 1126 \end{proof} 1127 1128 This last proposition is known as \emph{Frobenius reciprocity}, and was first 1129 proved by Frobenius himself in the context of finite groups. Another 1130 interesting construction is\dots 1131 1132 \begin{example}\label{ex:tensor-prod-separate-algs}\index{\(\mathfrak{g}\)-module!tensor product} 1133 Given two \(K\)-algebras \(A\) and \(B\), an \(A\)-module \(M\) and a 1134 \(B\)-module \(N\), \(M \otimes N = M \otimes_K N\) has the natural structure 1135 of an \(A \otimes_K B\)-module. In light of 1136 Example~\ref{ex:univ-enveloping-of-sum-is-tensor}, this implies that given 1137 Lie algebras \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\), a 1138 \(\mathfrak{g}_1\)-module \(M_1\) and a \(\mathfrak{g}_2\)-module \(M_2\), 1139 the space \(M_1 \otimes M_2\) has the natural structure of a \(\mathfrak{g}_1 1140 \oplus \mathfrak{g}_2\)-module, where the action of \(\mathfrak{g}_1 \oplus 1141 \mathfrak{g}_2\) is given by 1142 \[ 1143 (X_1 + X_2) \cdot (m \otimes n) 1144 = X_1 \cdot m \otimes n + m \otimes X_2 \cdot n 1145 \] 1146 \end{example} 1147 1148 Example~\ref{ex:tensor-prod-separate-algs} thus provides a way to describe 1149 representations of \(\mathfrak{g}_1 \oplus \mathfrak{g}_2\) in terms of the 1150 representations of \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\). We will soon see 1151 that in many cases \emph{all} (simple) \(\mathfrak{g}_1 \oplus 1152 \mathfrak{g}_2\)-modules can be constructed in such a manner. This concludes 1153 our initial remarks on \(\mathfrak{g}\)-modules. In the following chapters we 1154 will explore the fundamental problem of representation theory: that of 1155 classifying all representations of a given algebra up to isomorphism.