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

fin-dim-simple.tex (60877B)

   1 \chapter{Finite-Dimensional Simple Modules}
   2 
   3 In this chapter we classify the finite-dimensional simple
   4 \(\mathfrak{g}\)-modules for a finite-dimensional semisimple Lie algebra
   5 \(\mathfrak{g}\) over \(K\). At the heart of our analysis of
   6 \(\mathfrak{sl}_2(K)\) and \(\mathfrak{sl}_3(K)\) was the decision to consider
   7 the eigenspace decomposition
   8 \begin{equation}\label{sym-diag}
   9   M = \bigoplus_\lambda M_\lambda
  10 \end{equation}
  11 
  12 This was simple enough to do in the case of \(\mathfrak{sl}_2(K)\), but the
  13 rational behind it and the reason why equation (\ref{sym-diag}) holds are
  14 harder to explain in the case of \(\mathfrak{sl}_3(K)\). The eigenspace
  15 decomposition associated with an operator \(M \to M\) is a very well-known
  16 tool, and readers familiarized with basic concepts of linear algebra should be
  17 used to this type of argument. On the other hand, the eigenspace decomposition
  18 of \(M\) with respect to the action of an arbitrary subalgebra \(\mathfrak{h}
  19 \subset \mathfrak{gl}(M)\) is neither well-known nor does it hold in general:
  20 as indicated in the previous chapter, it may very well be that
  21 \[
  22   \bigoplus_{\lambda \in \mathfrak{h}^*} M_\lambda \subsetneq M
  23 \]
  24 
  25 We should note, however, that these two cases are not as different as they may
  26 sound at first glance. Specifically, we can regard the eigenspace decomposition
  27 of a \(\mathfrak{sl}_2(K)\)-module \(M\) with respect to the eigenvalues of the
  28 action of \(h\) as the eigenvalue decomposition of \(M\) with respect to the
  29 action of the subalgebra \(\mathfrak{h} = K h \subset \mathfrak{sl}_2(K)\).
  30 Furthermore, in both cases \(\mathfrak{h} \subset \mathfrak{sl}_n(K)\) is the
  31 subalgebra of diagonal matrices, which is Abelian. The fundamental difference
  32 between these two cases is thus the fact that \(\dim \mathfrak{h} = 1\) for
  33 \(\mathfrak{h} \subset \mathfrak{sl}_2(K)\) while \(\dim \mathfrak{h} > 1\) for
  34 \(\mathfrak{h} \subset \mathfrak{sl}_3(K)\). The question then is: why did we
  35 choose \(\mathfrak{h}\) with \(\dim \mathfrak{h} > 1\) for
  36 \(\mathfrak{sl}_3(K)\)?
  37 
  38 The rational behind fixing an Abelian subalgebra \(\mathfrak{h}\) is a simple
  39 one: we have seen in the previous chapter that representations of Abelian
  40 algebras are generally much simpler to understand than the general case. Thus
  41 it make sense to decompose a given \(\mathfrak{g}\)-module \(M\) of into
  42 subspaces invariant under the action of \(\mathfrak{h}\), and then analyze how
  43 the remaining elements of \(\mathfrak{g}\) act on these subspaces. The bigger
  44 \(\mathfrak{h}\) is, the simpler our problem gets, because there are fewer
  45 elements outside of \(\mathfrak{h}\) left to analyze.
  46 
  47 Hence we are generally interested in maximal Abelian subalgebras \(\mathfrak{h}
  48 \subset \mathfrak{g}\), which leads us to the following definition.
  49 
  50 \begin{definition}\index{Lie subalgebra!Cartan subalgebra}
  51   A subalgebra \(\mathfrak{h} \subset \mathfrak{g}\) is called \emph{a Cartan
  52   subalgebra of \(\mathfrak{g}\)} if is self-normalizing -- i.e. \([X, H] \in
  53   \mathfrak{h}\) for all \(H \in \mathfrak{h}\) if, and only if \(X \in
  54   \mathfrak{h}\) -- and nilpotent. Equivalently for reductive \(\mathfrak{g}\),
  55   \(\mathfrak{h}\) is called \emph{a Cartan subalgebra of \(\mathfrak{g}\)} if
  56   it is Abelian, \(\operatorname{ad}(H)\) is diagonalizable for each \(H \in
  57   \mathfrak{h}\) and if \(\mathfrak{h}\) is maximal with respect to the former
  58   two properties.
  59 \end{definition}
  60 
  61 \begin{proposition}
  62   There exists a Cartan subalgebra \(\mathfrak{h} \subset \mathfrak{g}\).
  63 \end{proposition}
  64 
  65 \begin{proof}
  66   Notice that \(0 \subset \mathfrak{g}\) is an Abelian subalgebra whose
  67   elements act as diagonal operators via the adjoint \(\mathfrak{g}\)-module.
  68   Indeed, \(0\), the only element of \(0 \subset \mathfrak{g}\), is such that
  69   \(\operatorname{ad}(0) = 0\) is a diagonalizable operator. Furthermore, given
  70   a chain of Abelian subalgebras
  71   \[
  72     0 \subset \mathfrak{h}_1 \subset \mathfrak{h}_2 \subset \cdots
  73   \]
  74   such that \(\operatorname{ad}(H)\) is a diagonal operator for each \(H \in
  75   \mathfrak{h}_i\), the subalgebra \(\bigcup_i \mathfrak{h}_i \subset
  76   \mathfrak{g}\) is Abelian, and its elements also act diagonally in
  77   \(\mathfrak{g}\). It then follows from Zorn's Lemma that there exists a
  78   subalgebra \(\mathfrak{h}\) which is maximal with respect to both these
  79   properties, also known as a Cartan subalgebra.
  80 \end{proof}
  81 
  82 We have already seen some concrete examples. Namely\dots
  83 
  84 \begin{example}\label{ex:cartan-of-gl}
  85   The Lie subalgebra
  86   \[
  87     \mathfrak{h} =
  88     \begin{pmatrix}
  89            K &      0 & \cdots &      0 \\
  90            0 &      K & \cdots &      0 \\
  91       \vdots & \vdots & \ddots & \vdots \\
  92            0 &      0 & \cdots &      K
  93     \end{pmatrix}
  94     \subset \mathfrak{gl}_n(K)
  95   \]
  96   of diagonal matrices is a Cartan subalgebra.
  97   Indeed, every pair of diagonal matrices commutes, so that \(\mathfrak{h}\)
  98   is an Abelian -- and hence nilpotent -- subalgebra. A
  99   simple calculation also shows that if \(i \ne j\) then the coefficient of
 100   \(E_{i j}\) in \([E_{i i}, X]\) is the same as the coefficient of \(E_{i j}\)
 101   in \(X\), for all \(X \in \mathfrak{gl}_n(K)\). In particular, if \([E_{i i},
 102   X]\) is diagonal for all \(i\), then so is \(X\) -- i.e. \(\mathfrak{h}\) is
 103   self-normalizing.
 104 \end{example}
 105 
 106 \begin{example}\label{ex:cartan-of-sl}
 107   Let \(\mathfrak{h}\) be as in Example~\ref{ex:cartan-of-gl}. Then the
 108   subalgebra \(\mathfrak{h} \cap \mathfrak{sl}_n(K)\) of traceless diagonal
 109   matrices is a Cartan subalgebra of \(\mathfrak{sl}_n(K)\).
 110 \end{example}
 111 
 112 \begin{example}\label{ex:cartan-of-sp}
 113   It is easy to see from Example~\ref{ex:sp2n} that \(\mathfrak{h} = \{X \in
 114   \mathfrak{sp}_{2n}(K) : X\ \text{is diagonal} \}\) is a Cartan subalgebra.
 115 \end{example}
 116 
 117 \begin{example}\label{ex:cartan-direct-sum}
 118   Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras and
 119   \(\mathfrak{h}_i \subset \mathfrak{g}_i\) be Cartan subalgebras. Then
 120   \(\mathfrak{h}_1 \oplus \mathfrak{h}_2\) is a Cartan subalgebra of
 121   \(\mathfrak{g}_1 \oplus \mathfrak{g}_2\).
 122 \end{example}
 123 
 124 \index{Cartan subalgebra!simultaneous diagonalization}
 125 The intersection of such subalgebra with \(\mathfrak{sl}_n(K)\) -- i.e. the
 126 subalgebra of traceless diagonal matrices -- is a Cartan subalgebra of
 127 \(\mathfrak{sl}_n(K)\). In particular, if \(n = 2\) or \(n = 3\) we get to the
 128 subalgebras described the previous chapter. The remaining question then is: if
 129 \(\mathfrak{h} \subset \mathfrak{g}\) is a Cartan subalgebra and \(M\) is a
 130 \(\mathfrak{g}\)-module, does the eigenspace decomposition
 131 \[
 132   M = \bigoplus_\lambda M_\lambda
 133 \]
 134 of \(M\) hold? The answer to this question turns out to be yes. This is a
 135 consequence of something known as \emph{simultaneous diagonalization}, which is
 136 the primary tool we will use to generalize the results of the previous section.
 137 What is simultaneous diagonalization all about then?
 138 
 139 \begin{definition}\label{def:sim-diag}
 140   Given a \(K\)-vector space \(V\), a set of operators \(\{T_j : V \to V\}_j\)
 141   is called \emph{simultaneously diagonalizable} if there is a basis \(\{v_1,
 142   \ldots, v_n\}\) for \(V\) such that \(T_j v_i\) is a scalar multiple of
 143   \(v_i\), for all \(i, j\).
 144 \end{definition}
 145 
 146 \begin{proposition}
 147   Given a \emph{finite-dimensional} vector space \(V\), a set of diagonalizable
 148   operators \(V \to V\) is simultaneously diagonalizable if, and only if all of
 149   its elements commute with one another.
 150 \end{proposition}
 151 
 152 We should point out that simultaneous diagonalization \emph{only works in the
 153 finite-dimensional setting}. In fact, simultaneous diagonalization is usually
 154 framed as an equivalent statement about diagonalizable \(n \times n\) matrices.
 155 Simultaneous diagonalization implies that to show \(M = \bigoplus_\lambda
 156 M_\lambda\) it suffices to show that \(H\!\restriction_M : M \to M\) is a
 157 diagonalizable operator for each \(H \in \mathfrak{h}\). To that end, we
 158 introduce \emph{the Jordan decomposition of an operator} and \emph{the abstract
 159 Jordan decomposition of a semisimple Lie algebra}.
 160 
 161 \begin{proposition}[Jordan]
 162   Given a finite-dimensional vector space \(V\) and an operator \(T : V \to
 163   V\), there are unique commuting operators \(T_{\operatorname{ss}},
 164   T_{\operatorname{nil}} : V \to V\), with \(T_{\operatorname{ss}}\)
 165   diagonalizable and \(T_{\operatorname{nil}}\) nilpotent, such that \(T =
 166   T_{\operatorname{ss}} + T_{\operatorname{nil}}\). The pair
 167   \((T_{\operatorname{ss}}, T_{\operatorname{nil}})\) is known as \emph{the Jordan
 168   decomposition of \(T\)}.
 169 \end{proposition}
 170 
 171 \begin{proposition}\index{abstract Jordan decomposition}
 172   Given \(\mathfrak{g}\) semisimple and \(X \in \mathfrak{g}\), there are
 173   \(X_{\operatorname{ss}}, X_{\operatorname{nil}} \in \mathfrak{g}\) such that \(X
 174   = X_{\operatorname{ss}} + X_{\operatorname{nil}}\), \([X_{\operatorname{ss}},
 175   X_{\operatorname{nil}}] = 0\), \(\operatorname{ad}(X_{\operatorname{ss}})\) is a
 176   diagonalizable operator and \(\operatorname{ad}(X_{\operatorname{nil}})\) is a
 177   nilpotent operator. The pair \((X_{\operatorname{ss}}, X_{\operatorname{nil}})\)
 178   is known as \emph{the Jordan decomposition of \(X\)}.
 179 \end{proposition}
 180 
 181 It should be clear from the uniqueness of
 182 \(\operatorname{ad}(X)_{\operatorname{ss}}\) and
 183 \(\operatorname{ad}(X)_{\operatorname{nil}}\) that the Jordan decomposition of
 184 \(\operatorname{ad}(X)\) is \(\operatorname{ad}(X) =
 185 \operatorname{ad}(X_{\operatorname{ss}}) +
 186 \operatorname{ad}(X_{\operatorname{nil}})\). What is perhaps more remarkable is
 187 the fact this holds for \emph{any} finite-dimensional \(\mathfrak{g}\)-module.
 188 In other words\dots
 189 
 190 \begin{proposition}\label{thm:preservation-jordan-form}
 191   Let \(M\) be a finite-dimensional \(\mathfrak{g}\)-module and \(X
 192   \in \mathfrak{g}\). Denote by \(X\!\restriction_M\) the action of \(X\) on
 193   \(M\). Then \(X_{\operatorname{ss}}\!\restriction_M =
 194   (X\!\restriction_M)_{\operatorname{ss}}\) and
 195   \(X_{\operatorname{nil}}\!\restriction_M =
 196   (X\!\restriction_M)_{\operatorname{nil}}\).
 197 \end{proposition}
 198 
 199 This last result is known as \emph{the preservation of the Jordan form}, and a
 200 proof can be found in appendix C of \cite{fulton-harris}. As promised this
 201 implies\dots
 202 
 203 \begin{corollary}\label{thm:finite-dim-is-weight-mod}
 204   Let \(\mathfrak{g}\) be a semisimple Lie algebra, \(\mathfrak{h} \subset
 205   \mathfrak{g}\) be a Cartan subalgebra and \(M\) be any finite-dimensional
 206   \(\mathfrak{g}\)-module. Then there is a basis \(\{m_1, \ldots,
 207   m_r\}\) of \(M\) so that each \(m_i\) is simultaneously an eigenvector of all
 208   elements of \(\mathfrak{h}\) -- i.e. each element of \(\mathfrak{h}\) acts as
 209   a diagonal matrix in this basis. In other words, there are linear functionals
 210   \(\lambda_i \in \mathfrak{h}^*\) so that
 211   \(
 212     H \cdot m_i = \lambda_i(H) m_i
 213   \)
 214   for all \(H \in \mathfrak{h}\). In particular,
 215   \[
 216     M = \bigoplus_{\lambda \in \mathfrak{h}^*} M_\lambda
 217   \]
 218 \end{corollary}
 219 
 220 \begin{proof}
 221   Fix some \(H \in \mathfrak{h}\). It suffices to show that \(H\!\restriction_M
 222   : M \to M\) is a diagonalizable operator.
 223 
 224   If we write \(H = H_{\operatorname{ss}} + H_{\operatorname{nil}}\) for the
 225   abstract Jordan decomposition of \(H\), we know
 226   \(\operatorname{ad}(H_{\operatorname{ss}}) =
 227   \operatorname{ad}(H)_{\operatorname{ss}}\). But \(\operatorname{ad}(H)\) is a
 228   diagonalizable operator, so that \(\operatorname{ad}(H)_{\operatorname{ss}} =
 229   \operatorname{ad}(H)\). This implies
 230   \(\operatorname{ad}(H_{\operatorname{nil}}) =
 231   \operatorname{ad}(H)_{\operatorname{nil}} = 0\), so that
 232   \(H_{\operatorname{nil}}\) is a central element of \(\mathfrak{g}\). Since
 233   \(\mathfrak{g}\) is semisimple, \(H_{\operatorname{nil}} = 0\).
 234   Proposition~\ref{thm:preservation-jordan-form} then implies
 235   \((H\!\restriction_M)_{\operatorname{nil}} =
 236   H_{\operatorname{nil}}\!\restriction_M = 0\), so \(H\!\restriction_M =
 237   (H\!\restriction_M)_{\operatorname{ss}}\) is a diagonalizable operator.
 238 \end{proof}
 239 
 240 We should point out that this last proof only works for semisimple Lie
 241 algebras. This is because we rely heavily on
 242 Proposition~\ref{thm:preservation-jordan-form}, as well in the fact that
 243 semisimple Lie algebras are centerless. In fact,
 244 Corollary~\ref{thm:finite-dim-is-weight-mod} fails even for reductive Lie
 245 algebras. For a counterexample, consider the algebra \(\mathfrak{g} = K\): the
 246 Cartan subalgebra of \(\mathfrak{g}\) is \(\mathfrak{g}\) itself, and a
 247 \(\mathfrak{g}\)-module is simply a vector space \(M\) endowed with an operator
 248 \(M \to M\) -- which corresponds to the action of \(1 \in \mathfrak{g}\) on
 249 \(M\). In particular, if we choose an operator \(M \to M\) which is \emph{not}
 250 diagonalizable we find \(M \ne \bigoplus_{\lambda \in \mathfrak{h}^*}
 251 M_\lambda\).
 252 
 253 However, Corollary~\ref{thm:finite-dim-is-weight-mod} does work for reductive
 254 \(\mathfrak{g}\) if we assume that the \(\mathfrak{g}\)-module \(M\) in
 255 question is simple, since central elements of \(\mathfrak{g}\) act on simple
 256 \(\mathfrak{g}\)-modules as scalar operators. The hypothesis of
 257 finite-dimensionality is also of huge importance. For instance, consider\dots
 258 
 259 \begin{example}\label{ex:regular-mod-is-not-weight-mod}
 260   Let \(\mathcal{U}(\mathfrak{g})\) denote the regular \(\mathfrak{g}\)-module.
 261   Notice that \(\mathcal{U}(\mathfrak{g})_\lambda = 0\) for all \(\lambda \in
 262   \mathfrak{h}^*\). Indeed, since \(\mathcal{U}(\mathfrak{g})\) is a domain, if
 263   \((H - \lambda(H)) u = 0\) for some nonzero \(H \in \mathfrak{h}\) then \(u =
 264   0\). In particular,
 265   \[
 266     \bigoplus_{\lambda \in \mathfrak{h}^*} \mathcal{U}(\mathfrak{g})_\lambda
 267     = 0 \neq \mathcal{U}(\mathfrak{g})
 268   \]
 269 \end{example}
 270 
 271 As a first consequence of Corollary~\ref{thm:finite-dim-is-weight-mod} we
 272 show\dots
 273 
 274 \begin{corollary}
 275   The restriction of the Killing form \(\kappa\) to \(\mathfrak{h}\) is
 276   non-degenerate.
 277 \end{corollary}
 278 
 279 \begin{proof}
 280   Consider the root space decomposition \(\mathfrak{g} = \mathfrak{g}_0 \oplus
 281   \bigoplus_\alpha \mathfrak{g}_\alpha\) of the adjoint
 282   \(\mathfrak{g}\)-module, where \(\alpha\) ranges over all nonzero eigenvalues
 283   of the adjoint action of \(\mathfrak{h}\). We claim \(\mathfrak{g}_0 =
 284   \mathfrak{h}\).
 285 
 286   Indeed, since \(\mathfrak{h}\) is Abelian, \(\operatorname{ad}(\mathfrak{h})
 287   \mathfrak{h} = 0\) -- i.e. \(\mathfrak{h} \subset \mathfrak{g}_0\). On the
 288   other hand, since \(\mathfrak{h}\) is self-normalizing, if \([X, H] = 0 \in
 289   \mathfrak{h}\) for all \(H \in \mathfrak{h}\) then \(X \in \mathfrak{h}\) --
 290   i.e. \(\mathfrak{g}_0 \subset \mathfrak{h}\). So the eigenspace decomposition
 291   becomes
 292   \[
 293     \mathfrak{g} = \mathfrak{h} \oplus \bigoplus_\alpha \mathfrak{g}_\alpha
 294   \]
 295 
 296   We furthermore claim that \(\mathfrak{h} = \mathfrak{g}_0\) is orthogonal to
 297   \(\mathfrak{g}_\alpha\) with respect to \(\kappa\) for any \(\alpha \ne 0\).
 298   Indeed, given \(X \in \mathfrak{g}_\alpha\) and \(H_1, H_2 \in \mathfrak{h}\)
 299   with \(\alpha(H_1) \ne 0\) we have
 300   \[
 301     \alpha(H_1) \cdot \kappa(X, H_2)
 302     = \kappa([H_1, X], H_2)
 303     = - \kappa([X, H_1], H_2)
 304     = - \kappa(X, [H_1, H_2])
 305     = 0
 306   \]
 307 
 308   Hence the non-degeneracy of \(\kappa\) implies the non-degeneracy of its
 309   restriction.
 310 \end{proof}
 311 
 312 We should point out that the restriction of \(\kappa\) to \(\mathfrak{h}\) is
 313 \emph{not} the Killing form of \(\mathfrak{h}\). In fact, since
 314 \(\mathfrak{h}\) is Abelian, its Killing form is identically zero -- which is
 315 hardly ever a non-degenerate form.
 316 
 317 \begin{note}
 318   Since \(\kappa\) induces an isomorphism \(\mathfrak{h} \isoto
 319   \mathfrak{h}^*\), it induces a bilinear form \((\kappa(X, \cdot), \kappa(Y,
 320   \cdot)) \mapsto \kappa(X, Y)\) in \(\mathfrak{h}^*\). As in
 321   section~\ref{sec:sl3-reps}, we denote this form by \(\kappa\) as well.
 322 \end{note}
 323 
 324 We now have most of the necessary tools to reproduce the results of the
 325 previous chapter in a general setting. Let \(\mathfrak{g}\) be a
 326 finite-dimensional semisimple algebra with a Cartan subalgebra \(\mathfrak{h}\)
 327 and let \(M\) be a finite-dimensional simple \(\mathfrak{g}\)-module. We will
 328 proceed, as we did before, by generalizing the results of the previous two
 329 sections in order. By now the pattern should be starting to become clear, so we
 330 will mostly omit technical details and proofs analogous to the ones on the
 331 previous sections. Further details can be found in appendix D of
 332 \cite{fulton-harris} and in \cite{humphreys}.
 333 
 334 \section{The Geometry of Roots and Weights}
 335 
 336 We begin our analysis, as we did for \(\mathfrak{sl}_2(K)\) and
 337 \(\mathfrak{sl}_3(K)\), by investigating the locus of roots of and weights of
 338 \(\mathfrak{g}\). Throughout chapter~\ref{ch:sl3} we have seen that the weights
 339 of any given finite-dimensional module of \(\mathfrak{sl}_2(K)\) or
 340 \(\mathfrak{sl}_3(K)\) can only assume very rigid configurations. For instance,
 341 we have seen that the roots of \(\mathfrak{sl}_2(K)\) and
 342 \(\mathfrak{sl}_3(K)\) are symmetric with respect to the origin. In this
 343 chapter we will generalize most results from chapter~\ref{ch:sl3} regarding the
 344 rigidity of the geometry of the set of weights of a given module.
 345 
 346 As for the aforementioned result on the symmetry of roots, this turns out to be
 347 a general fact, which is a consequence of the non-degeneracy of the restriction
 348 of the Killing form to the Cartan subalgebra.
 349 
 350 \begin{proposition}\label{thm:weights-symmetric-span}
 351   The roots \(\alpha\) of \(\mathfrak{g}\) are symmetrical about the origin --
 352   i.e. \(- \alpha\) is also a root -- and they span all of \(\mathfrak{h}^*\).
 353 \end{proposition}
 354 
 355 \begin{proof}
 356   We will start with the first claim. Let \(\alpha\) and \(\beta\) be two
 357   roots. Notice \([\mathfrak{g}_\alpha, \mathfrak{g}_\beta] \subset
 358   \mathfrak{g}_{\alpha + \beta}\). Indeed, if \(X \in \mathfrak{g}_\alpha\) and
 359   \(Y \in \mathfrak{g}_\beta\) then
 360   \[
 361     [H, [X, Y]]
 362     = [X, [H, Y]] - [Y, [H, X]]
 363     = (\alpha + \beta)(H) \cdot [X, Y]
 364   \]
 365   for all \(H \in \mathfrak{h}\).
 366 
 367   This implies that if \(\alpha + \beta \ne 0\) then \(\operatorname{ad}(X)
 368   \operatorname{ad}(Y)\) is nilpotent: if \(Z \in \mathfrak{g}_\gamma\) then
 369   \[
 370     (\operatorname{ad}(X) \operatorname{ad}(Y))^r Z
 371     = [X, [Y, [ \ldots, [X, [Y, Z]]] \ldots ]
 372     \in \mathfrak{g}_{r \alpha + r \beta + \gamma}
 373     = 0
 374   \]
 375   for \(r\) large enough. In particular, \(\kappa(X, Y) =
 376   \operatorname{Tr}(\operatorname{ad}(X) \operatorname{ad}(Y)) = 0\). Now if
 377   \(- \alpha\) is not an eigenvalue we find \(\kappa(X, \mathfrak{g}_\beta) =
 378   0\) for all roots \(\beta\), which contradicts the non-degeneracy of
 379   \(\kappa\). Hence \(- \alpha\) must be an eigenvalue of the adjoint action of
 380   \(\mathfrak{h}\).
 381 
 382   For the second statement, note that if the roots of \(\mathfrak{g}\) do not
 383   span all of \(\mathfrak{h}^*\) then there is some nonzero \(H \in
 384   \mathfrak{h}\) such that \(\alpha(H) = 0\) for all roots \(\alpha\), which is
 385   to say, \(\operatorname{ad}(H) X = [H, X] = 0\) for all \(X \in
 386   \mathfrak{g}\). Another way of putting it is to say \(H\) is an element of
 387   the center \(\mathfrak{z} = 0\) of \(\mathfrak{g}\), a contradiction.
 388 \end{proof}
 389 
 390 Furthermore, as in the case of \(\mathfrak{sl}_2(K)\) and
 391 \(\mathfrak{sl}_3(K)\) one can show\dots
 392 
 393 \begin{proposition}\label{thm:root-space-dim-1}
 394   The root spaces \(\mathfrak{g}_\alpha\) are all \(1\)-dimensional.
 395 \end{proposition}
 396 
 397 The proof of the first statement of
 398 Proposition~\ref{thm:weights-symmetric-span} highlights something interesting:
 399 if we fix some eigenvalue \(\alpha\) of the adjoint action of \(\mathfrak{h}\)
 400 on \(\mathfrak{g}\) and a eigenvector \(X \in \mathfrak{g}_\alpha\), then for
 401 each \(H \in \mathfrak{h}\) and \(m \in M_\lambda\) we find
 402 \[
 403   H \cdot (X \cdot m)
 404   = X H \cdot m + [H, X] \cdot m
 405   = (\lambda + \alpha)(H) X \cdot m
 406 \]
 407 
 408 Thus \(X\) sends \(m\) to \(M_{\lambda + \alpha}\). We have encountered this
 409 formula twice in these notes: again, we find \(\mathfrak{g}_\alpha\) \emph{acts
 410 on \(M\) by translating vectors between eigenspaces}. In particular, if we
 411 denote by \(\Delta\) the set of all roots of \(\mathfrak{g}\) then\dots
 412 
 413 \begin{theorem}\label{thm:weights-congruent-mod-root}\index{weights!root lattice}
 414   The weights of a finite-dimensional simple \(\mathfrak{g}\)-module \(M\) are
 415   all congruent modulo the root lattice \(Q = \mathbb{Z} \Delta\) of
 416   \(\mathfrak{g}\). In other words, all weights of \(M\) lie in the same
 417   \(Q\)-coset \(\xi \in \mfrac{\mathfrak{h}^*}{Q}\).
 418 \end{theorem}
 419 
 420 Again, we may leverage our knowledge of \(\mathfrak{sl}_2(K)\) to obtain
 421 further restrictions on the geometry of the locus of weights of \(M\). Namely,
 422 as in the case of \(\mathfrak{sl}_3(K)\) we show\dots
 423 
 424 \begin{proposition}\label{thm:distinguished-subalgebra}
 425   Given a root \(\alpha\) of \(\mathfrak{g}\) the subspace
 426   \(\mathfrak{s}_\alpha = \mathfrak{g}_\alpha \oplus \mathfrak{g}_{- \alpha}
 427   \oplus [\mathfrak{g}_\alpha, \mathfrak{g}_{- \alpha}]\) is a subalgebra
 428   isomorphic to \(\mathfrak{sl}_2(K)\).
 429 \end{proposition}
 430 
 431 \begin{corollary}\label{thm:distinguished-subalg-rep}
 432   For all weights \(\mu\), the subspace
 433   \[
 434     \bigoplus_k M_{\mu - k \alpha}
 435   \]
 436   is invariant under the action of the subalgebra \(\mathfrak{s}_\alpha\)
 437   and the weight spaces in this string match the eigenspaces of \(h\).
 438 \end{corollary}
 439 
 440 The proof of Proposition~\ref{thm:distinguished-subalgebra} is very technical
 441 in nature and we won't include it here, but the idea behind it is simple:
 442 recall that \(\mathfrak{g}_\alpha\) and \(\mathfrak{g}_{- \alpha}\) are both
 443 \(1\)-dimensional, so that \(\dim [\mathfrak{g}_\alpha, \mathfrak{g}_{-
 444 \alpha}]\) is at most 1. We check that \([\mathfrak{g}_\alpha, \mathfrak{g}_{-
 445 \alpha}] \ne 0\) and that no generator of \([\mathfrak{g}_\alpha,
 446 \mathfrak{g}_{- \alpha}]\) is annihilated by \(\alpha\), so that by adjusting
 447 scalars we can find \(E_\alpha \in \mathfrak{g}_\alpha\) and \(F_\alpha \in
 448 \mathfrak{g}_{- \alpha}\) such that \(H_\alpha = [E_\alpha, F_\alpha]\)
 449 satisfies
 450 \begin{align*}
 451   [H_\alpha, F_\alpha] & = -2 F_\alpha &
 452   [H_\alpha, E_\alpha] & =  2 E_\alpha
 453 \end{align*}
 454 
 455 The elements \(E_\alpha, F_\alpha \in \mathfrak{g}\) are not uniquely
 456 determined by this condition, but \(H_\alpha\) is. As promised, the second
 457 statement of Corollary~\ref{thm:distinguished-subalg-rep} imposes strong
 458 restrictions on the weights of \(M\). Namely, if \(\lambda\) is a weight,
 459 \(\lambda(H_\alpha)\) is an eigenvalue of \(h\) on some
 460 \(\mathfrak{sl}_2(K)\)-module, so it must be an integer. In other words\dots
 461 
 462 \begin{definition}\label{def:weight-lattice}\index{weights!weight lattice}
 463   The lattice \(P = \{ \lambda \in \mathfrak{h}^* : \lambda(H_\alpha) \in
 464   \mathbb{Z} \, \forall \alpha \in \Delta \} \subset \mathfrak{h}^*\) is called
 465   \emph{the weight lattice of \(\mathfrak{g}\)}. We call the elements of \(P\)
 466   \emph{integral}.
 467 \end{definition}
 468 
 469 \begin{proposition}\label{thm:weights-fit-in-weight-lattice}
 470   The weights of a finite-dimensional simple \(\mathfrak{g}\)-module \(M\) of
 471   all lie in the weight lattice \(P\).
 472 \end{proposition}
 473 
 474 Proposition~\ref{thm:weights-fit-in-weight-lattice} is clearly analogous to
 475 Corollary~\ref{thm:sl3-weights-fit-in-weight-lattice}. In fact, the weight
 476 lattice of \(\mathfrak{sl}_3(K)\) -- as in Definition~\ref{def:weight-lattice}
 477 -- is precisely \(\mathbb{Z} \langle \epsilon_1, \epsilon_2, \epsilon_3 \rangle\). To
 478 proceed further, we would like to take \emph{the highest weight of \(M\)} as in
 479 section~\ref{sec:sl3-reps}, but the meaning of \emph{highest} is again unclear
 480 in this situation. We could simply fix a linear function \(\mathbb{Q} P \to
 481 \mathbb{Q}\) -- as we did in section~\ref{sec:sl3-reps} -- and choose a weight
 482 \(\lambda\) of \(M\) that maximizes this functional, but at this point it is
 483 convenient to introduce some additional tools to our arsenal. These tools are
 484 called \emph{basis}.
 485 
 486 \begin{definition}\label{def:basis-of-root}\index{weights!basis}
 487   A subset \(\Sigma = \{\beta_1, \ldots, \beta_r\} \subset \Delta\) of linearly
 488   independent roots is called \emph{a basis for \(\Delta\)} if, given \(\alpha
 489   \in \Delta\), there are unique \(k_1, \ldots, k_r \in \mathbb{N}\) such that
 490   \(\alpha = \pm(k_1 \beta_1 + \cdots + k_r \beta_r)\).
 491 \end{definition}
 492 
 493 \begin{example}\label{ex:sl-canonical-basis}
 494   Suppose \(\mathfrak{g} = \mathfrak{sl}_n(K)\) and \(\mathfrak{h} \subset
 495   \mathfrak{g}\) is the subalgebra of diagonal matrices, as in
 496   Example~\ref{ex:cartan-of-sl}. Consider the linear functionals \(\epsilon_1,
 497   \ldots, \epsilon_n \in \mathfrak{h}^*\) such that \(\epsilon_i(H)\) is the
 498   \(i\)-th entry of the diagonal of \(H\). As observed in
 499   section~\ref{sec:sl3-reps} for \(n = 3\), the roots of \(\mathfrak{sl}_n(K)\)
 500   are \(\epsilon_i - \epsilon_j\) for \(i \ne j\) -- with root vectors given by
 501   \(E_{i j}\) -- and we may take the basis \(\Sigma = \{\beta_1, \ldots,
 502   \beta_{n-1}\}\) with \(\beta_i = \epsilon_i - \epsilon_{i+1}\).
 503 \end{example}
 504 
 505 \begin{example}\label{ex:sp-canonical-basis}
 506   Suppose \(\mathfrak{g} = \mathfrak{sp}_{2n}(K)\) and \(\mathfrak{h} \subset
 507   \mathfrak{g}\) is the subalgebra of diagonal matrices, as in
 508   Example~\ref{ex:cartan-of-sp}. Consider the linear functionals \(\epsilon_1,
 509   \ldots, \epsilon_n \in \mathfrak{h}^*\) such that \(\epsilon_i(H)\) is the
 510   \(i\)-th entry of the diagonal of \(H\). Then the roots of
 511   \(\mathfrak{sp}_{2n}(K)\) are \(\pm \epsilon_i \pm \epsilon_j\) for \(i \ne
 512   j\) and \(\pm 2 \epsilon_i\) -- see \cite[ch.~16]{fulton-harris}. In this
 513   case, we may take the basis \(\Sigma = \{\beta_1, \ldots, \beta_n\}\) with
 514   \(\beta_i = \epsilon_i - \epsilon_{i+1}\) for \(i < n\) and \(\beta_n = 2
 515   \epsilon_n\).
 516 \end{example}
 517 
 518 The interesting thing about basis for \(\Delta\) is that they allow us to
 519 compare weights of a given \(\mathfrak{g}\)-module. At this point the reader
 520 should be asking himself: how? Definition~\ref{def:basis-of-root} isn't exactly
 521 all that intuitive. Well, the thing is that any choice of basis \(\Sigma\)
 522 induces an order in \(Q\), where elements are ordered by their
 523 \emph{\(\Sigma\)-coordinates}.
 524 
 525 \begin{definition}\index{weights!orderings of roots}
 526   Let \(\Sigma = \{\beta_1, \ldots, \beta_r\}\) be a basis for \(\Delta\).
 527   Given \(\alpha = k_1 \beta_1 + \cdots + k_r \beta_r \in Q\) with \(k_1,
 528   \ldots, k_r \in \mathbb{Z}\), we call the vector \(\alpha_\Sigma = (k_1,
 529   \ldots, k_r) \in \mathbb{Z}^r\) \emph{the \(\Sigma\)-coordinate of
 530   \(\alpha\)}. We say that \(\alpha \preceq \beta\) if \(\alpha_\Sigma \le
 531   \beta_\Sigma\) in the lexicographical order.
 532 \end{definition}
 533 
 534 \begin{definition}
 535   Given a basis \(\Sigma\) for \(\Delta\), there is a canonical
 536   partition\footnote{Notice that $\operatorname{ht}(\alpha) = 0$ if, and only
 537   if $\alpha = 0$. Since $0$ is, by definition, not a root, the sets $\Delta^+$
 538   and $\Delta^-$ account for all roots.} \(\Delta^+ \cup \Delta^- = \Delta\),
 539   where \(\Delta^+ = \{ \alpha \in \Delta : \alpha \succ 0 \}\) and \(\Delta^-
 540   = \{ \alpha \in \Delta : \alpha \prec 0 \}\). The elements of \(\Delta^+\)
 541   and \(\Delta^-\) are called \emph{positive} and \emph{negative roots},
 542   respectively.
 543 \end{definition}
 544 
 545 \begin{example}
 546   If \(\mathfrak{g} = \mathfrak{sl}_3(K)\) and \(\Sigma\) is as in
 547   Example~\ref{ex:sl-canonical-basis} then the partition \(\Delta^+ \cup
 548   \Delta^-\) induced by \(\Sigma\) is the same as the one described in
 549   section~\ref{sec:sl3-reps}.
 550 \end{example}
 551 
 552 \begin{definition}\index{Lie subalgebra!Borel subalgebra}\index{Lie subalgebra!parabolic subalgebra}
 553   Let \(\Sigma\) be a basis for \(\Delta\). The subalgebra \(\mathfrak{b} =
 554   \mathfrak{h} \oplus \bigoplus_{\alpha \in \Delta^+} \mathfrak{g}_\alpha\) is
 555   called \emph{the Borel subalgebra associated with \(\mathfrak{h}\) and
 556   \(\Sigma\)}. A subalgebra \(\mathfrak{p} \subset \mathfrak{g}\) is called
 557   \emph{parabolic} if \(\mathfrak{p} \supset \mathfrak{b}\).
 558 \end{definition}
 559 
 560 It should be obvious that the binary relation \(\preceq\) in \(Q\) is a total
 561 order. In addition, we may compare the elements of a given \(Q\)-coset
 562 \(\lambda + Q\) by comparing their difference with \(0 \in Q\). In other words,
 563 given \(\lambda \in \mu + Q\), we say \(\lambda \preceq \mu\) if \(\lambda -
 564 \mu \preceq 0\). In particular, since the weights of \(M\) all lie in a single
 565 \(Q\)-coset, we may compare them in this way. Given a basis \(\Sigma\) for
 566 \(\Delta\) we may take ``the highest weight of \(M\)'' as a maximal weight
 567 \(\lambda\) of \(M\). The obvious question then is: can we always find a basis
 568 for \(\Delta\)?
 569 
 570 \begin{proposition}
 571   There is a basis \(\Sigma\) for \(\Delta\).
 572 \end{proposition}
 573 
 574 The intuition behind the proof of this proposition is similar to our original
 575 idea of fixing a direction in \(\mathfrak{h}^*\) in the case of
 576 \(\mathfrak{sl}_3(K)\). Namely, one can show that \(\kappa(\alpha, \beta) \in
 577 \mathbb{Z}\) for all \(\alpha, \beta \in \Delta\), so that the Killing form
 578 \(\kappa\) restricts to a nondegenerate \(\mathbb{Q}\)-bilinear form
 579 \(\mathbb{Q} \Delta \times \mathbb{Q} \Delta \to \mathbb{Q}\). We can then fix
 580 a nonzero vector \(\gamma \in \mathbb{Q} \Delta\) and consider the orthogonal
 581 projection \(f : \mathbb{Q} \Delta \to \mathbb{Q} \gamma \cong \mathbb{Q}\). We
 582 say a root \(\alpha \in \Delta\) is \emph{positive} if \(f(\alpha) > 0\), and
 583 we call a positive root \(\alpha\) \emph{simple} if it cannot be written as the
 584 sum two other positive roots. The subset \(\Sigma \subset \Delta\) of all
 585 simple roots is a basis for \(\Delta\), and all other basis can be shown to
 586 arise in this way.
 587 
 588 Fix some basis \(\Sigma\) for \(\Delta\), with corresponding decomposition
 589 \(\Delta^+ \cup \Delta^- = \Delta\). Let \(\lambda\) be a maximal weight of
 590 \(M\). We call \(\lambda\) \emph{the highest weight of \(M\)}, and we call any
 591 nonzero \(m \in M_\lambda\) \emph{a highest weight vector}. The strategy then
 592 is to describe all weight spaces of \(M\) in terms of \(\lambda\) and \(m\), as
 593 in Theorem~\ref{thm:sl3-irr-weights-class}. Unsurprisingly we do so by
 594 reproducing the proof of the case of \(\mathfrak{sl}_3(K)\).
 595 
 596 First, we note that any highest weight vector \(m \in M_\lambda\) is
 597 annihilated by all positive root spaces, for if \(\alpha \in \Delta^+\) then
 598 \(E_\alpha \cdot m \in M_{\lambda + \alpha}\) must be zero -- or otherwise we
 599 would have that \(\lambda + \alpha\) is a weight with \(\lambda \prec \lambda +
 600 \alpha\). In particular,
 601 \[
 602   \bigoplus_{k \in \mathbb{Z}}   M_{\lambda - k \alpha}
 603   = \bigoplus_{k \in \mathbb{N}} M_{\lambda - k \alpha}
 604 \]
 605 and \(\lambda(H_\alpha)\) is the right-most eigenvalue of the action of \(h\)
 606 on the \(\mathfrak{sl}_2(K)\)-module \(\bigoplus_k M_{\lambda - k \alpha}\).
 607 
 608 This has a number of important consequences. For instance\dots
 609 
 610 \begin{corollary}
 611   If \(\alpha \in \Delta^+\) and \(\sigma_\alpha : \mathfrak{h}^* \to
 612   \mathfrak{h}^*\) is the reflection in the hyperplane perpendicular to
 613   \(\alpha\) with respect to the Killing form, the weights of \(M\) occurring
 614   in the line joining \(\lambda\) and \(\sigma_\alpha\) are precisely the \(\mu
 615   \in P\) lying between \(\lambda\) and \(\sigma_\alpha(\lambda)\).
 616 \end{corollary}
 617 
 618 \begin{proof}
 619   Notice that any \(\mu \in P\) in the line joining \(\lambda\) and
 620   \(\sigma_\alpha(\lambda)\) has the form \(\mu = \lambda - k \alpha\) for some
 621   \(k\), so that \(M_\mu\) corresponds the eigenspace associated with the
 622   eigenvalue \(\lambda(H_\alpha) - 2k\) of the action of \(h\) on \(\bigoplus_k
 623   M_{\lambda - k \alpha}\). If \(\mu\) lies between \(\lambda\) and
 624   \(\sigma_\alpha(\lambda)\) then \(k\) lies between \(0\) and
 625   \(\lambda(H_\alpha)\), in which case \(M_\mu \neq 0\) and therefore \(\mu\)
 626   is a weight.
 627 
 628   On the other hand, if \(\mu\) does not lie between \(\lambda\) and
 629   \(\sigma_\alpha(\lambda)\) then either \(k < 0\) or \(k >
 630   \lambda(H_\alpha)\). Suppose \(\mu\) is a weight. In the first case \(\mu
 631   \succ \lambda\), a contradiction. On the second case the fact that \(M_\mu
 632   \ne 0\) implies \(M_{\lambda  + (k - \lambda(H_\alpha)) \alpha} =
 633   M_{\sigma_\alpha(\mu)} \ne 0\), which contradicts the fact that \(M_{\lambda
 634   + \ell \alpha} = 0\) for all \(\ell \ge 0\).
 635 \end{proof}
 636 
 637 This is entirely analogous to the situation of \(\mathfrak{sl}_3(K)\), where we
 638 found that the weights of the simple \(\mathfrak{sl}_3(K)\)-modules formed
 639 continuous strings symmetric with respect to the lines \(K \alpha\) with
 640 \(\kappa(\epsilon_i - \epsilon_j, \alpha) = 0\). As in the case of
 641 \(\mathfrak{sl}_3(K)\), the same sort of arguments leads us to the
 642 conclusion\dots
 643 
 644 \begin{definition}\index{Weyl group}
 645   We refer to the (finite) group \(W = \langle \sigma_\alpha : \alpha \in
 646   \Delta \rangle = \langle \sigma_\beta : \beta \in \Sigma \rangle \subset
 647   \operatorname{O}(\mathfrak{h}^*)\) as \emph{the Weyl group of
 648   \(\mathfrak{g}\)}.
 649 \end{definition}
 650 
 651 \begin{theorem}\label{thm:irr-weight-class}
 652   The weights of a simple \(\mathfrak{g}\)-module \(M\) with highest weight
 653   \(\lambda\) are precisely the elements of the weight lattice \(P\) congruent
 654   to \(\lambda\) modulo the root lattice \(Q\) lying inside the convex hull of
 655   the orbit of \(\lambda\) under the action of the Weyl group \(W\).
 656 \end{theorem}
 657 
 658 At this point we are basically done with results regarding the geometry of the
 659 weights of \(M\), but it is convenient to introduce some further notation.
 660 Aside from showing up in the previous theorem, the Weyl group will also play an
 661 important role in chapter~\ref{ch:mathieu} by virtue of the existence of a
 662 canonical action of \(W\) on \(\mathfrak{h}\).
 663 
 664 \begin{definition}\index{Weyl group!natural action}\index{Weyl group!dot action}
 665   The canonical action of \(W\) on \(\mathfrak{h}^*\) given by \(\sigma \cdot
 666   \lambda = \sigma(\lambda)\) is called \emph{the natural action of \(W\)}. We
 667   also consider the equivalent ``shifted'' action \(\sigma \bullet \lambda =
 668   \sigma(\lambda + \rho) - \rho\) of \(W\) on \(\mathfrak{h}^*\), known as
 669   \emph{the dot action of \(W\)} -- here \(\rho =  \sfrac{1}{2} \beta_1 +
 670   \cdots \sfrac{1}{2} \beta_r\).
 671 \end{definition}
 672 
 673 This already allow us to compute some examples of Weyl groups.
 674 
 675 \begin{example}\label{ex:sl-weyl-group}
 676   Suppose \(\mathfrak{g} = \mathfrak{sl}_n(K)\) and \(\mathfrak{h} \subset
 677   \mathfrak{g}\) is as in Example~\ref{ex:cartan-of-sl}. Let \(\epsilon_1,
 678   \ldots, \epsilon_n \in \mathfrak{h}^*\) be as in
 679   Example~\ref{ex:sl-canonical-basis} and take the associated basis \(\Sigma =
 680   \{\beta_1, \ldots, \beta_{n-1}\}\) for \(\Delta\), \(\beta_i = \epsilon_i -
 681   \epsilon_{i + 1}\). Then a simple calculation shows that \(\sigma_{\beta_i}\)
 682   permutes \(\epsilon_i\) and \(\epsilon_{i+1}\) and fixes the other
 683   \(\epsilon_j\). This translates to a canonical isomorphism
 684   \begin{align*}
 685                    W & \isoto  S_n                       \\
 686     \sigma_{\beta_i} & \mapsto \sigma_i = (i \; i\!+\!1)
 687   \end{align*}
 688 \end{example}
 689 
 690 \begin{example}\label{ex:sp-weyl-group}
 691   Suppose \(\mathfrak{g} = \mathfrak{sp}_{2n}(K)\) and \(\mathfrak{h} \subset
 692   \mathfrak{g}\) is as in Example~\ref{ex:cartan-of-sp}. Let \(\epsilon_1,
 693   \ldots, \epsilon_n \in \mathfrak{h}^*\) be as in
 694   Example~\ref{ex:sp-canonical-basis} and take the associated basis \(\Sigma =
 695   \{\beta_1, \ldots, \beta_n\}\) for \(\Delta\). Then a simple calculation
 696   shows that \(\sigma_{\beta_i}\) permutes \(\epsilon_i\) and
 697   \(\epsilon_{i+1}\) for \(i < n\) and \(\sigma_{\beta_n}\) switches the sign
 698   of \(\epsilon_n\). This translates to a canonical isomorphism
 699   \begin{align*}
 700                    W & \isoto  S_n \ltimes (\mathbb{Z}/2\mathbb{Z})^n \\
 701     \sigma_{\beta_i} & \mapsto (\sigma_i, (\bar 0, \ldots, \bar 0))   \\
 702     \sigma_{\beta_n} & \mapsto (1, (\bar 0, \ldots, \bar 0, \bar 1)),
 703   \end{align*}
 704   where \(\sigma_i = (i \ i\!+\!1)\) are the canonical transpositions.
 705 \end{example}
 706 
 707 If we conjugate some \(\sigma \in W\) by the isomorphism \(\mathfrak{h}^*
 708 \isoto \mathfrak{h}\) afforded by the restriction of the Killing for to
 709 \(\mathfrak{h}\) we get a linear action of \(W\) on \(\mathfrak{h}\), which is
 710 given by \(\kappa(\sigma \cdot H, \cdot) = \sigma \cdot \kappa(H, \cdot)\). As
 711 it turns out, this action can be extended to an action of \(W\) on
 712 \(\mathfrak{g}\) by automorphisms of Lie algebras. This translates into the
 713 following results, which we do not prove -- but see
 714 \cite[sec.~14.3]{humphreys}.
 715 
 716 \begin{proposition}\label{thm:weyl-group-action}
 717   Given \(\alpha \in \Delta^+\), there is an automorphism of Lie algebras
 718   \(f_\alpha : \mathfrak{g} \isoto \mathfrak{g}\) such that
 719   \(f_\alpha(H) = \sigma_\alpha \cdot H\) for all \(H \in \mathfrak{h}\). In
 720   addition, these automorphisms can be chosen in such a way that the family
 721   \(\{f_\alpha\}_{\alpha \in \Delta^+}\) defines an action of \(W\) on
 722   \(\mathfrak{g}\) -- which is obviously compatible with the natural action of
 723   \(W\) on \(\mathfrak{h}\).
 724 \end{proposition}
 725 
 726 \begin{note}
 727   We should notice the action of \(W\) on \(\mathfrak{g}\) from
 728   Proposition~\ref{thm:weyl-group-action} is not canonical, since it depends on
 729   the choice of \(E_\alpha\) and \(F_\alpha\). Nevertheless, different choices
 730   of \(E_\alpha\) and \(F_\alpha\) yield isomorphic actions and the restriction
 731   of these actions to \(\mathfrak{h}\) is independent of any choices.
 732 \end{note}
 733 
 734 We should point out that the results in this section regarding the geometry
 735 roots and weights are only the beginning of a well develop axiomatic theory of
 736 the so called \emph{root systems}, which was used by Cartan in the early 20th
 737 century to classify all finite-dimensional simple complex Lie algebras in terms
 738 of Dynking diagrams. This and much more can be found in \cite[III]{humphreys}
 739 and \cite[ch.~21]{fulton-harris}. Having found all of the weights of \(M\), the
 740 only thing we are missing for a complete classification is an existence and
 741 uniqueness theorem analogous to Theorem~\ref{thm:sl2-exist-unique} and
 742 Theorem~\ref{thm:sl3-existence-uniqueness}. This will be the focus of the next
 743 section.
 744 
 745 \section{Highest Weight Modules \& the Highest Weight Theorem}
 746 
 747 It is already clear from the previous discussion that if \(\lambda\) is the
 748 highest weight of \(M\) then \(\lambda(H_\alpha) \ge 0\) for all positive roots
 749 \(\alpha\). Indeed, as in the \(\mathfrak{sl}_3(K)\), for each \(\alpha \in
 750 \Delta^+\) we know \(\lambda(H_\alpha)\) is the highest eigenvalue of the
 751 action of \(h\) in the \(\mathfrak{sl}_2(K)\)-module \(\bigoplus_k M_{\lambda -
 752 k \alpha}\) -- which must be a non-negative integer. This fact may be
 753 summarized in the following proposition.
 754 
 755 \begin{definition}\index{weights!dominant weight}\index{weights!integral weight}
 756   An element \(\lambda\) of \(P\) such that \(\lambda(H_\alpha) \ge 0\) for all
 757   \(\alpha \in \Delta^+\) is referred to as an \emph{dominant integral weight
 758   of \(\mathfrak{g}\)}. The set of all dominant integral weights is denotes by
 759   \(P^+\).
 760 \end{definition}
 761 
 762 \begin{proposition}\label{thm:highes-weight-of-fin-dim-is-dominant}
 763   Suppose \(M\) is a finite-dimensional simple \(\mathfrak{g}\)-module and
 764   \(\lambda\) is its highest weight. Then \(\lambda\) is a dominant integral
 765   weight of \(\mathfrak{g}\).
 766 \end{proposition}
 767 
 768 The condition that \(\lambda \in P^+\) is thus necessary for the existence of a
 769 simple \(\mathfrak{g}\)-module with highest weight given by \(\lambda\). Given
 770 our previous experience with \(\mathfrak{sl}_2(K)\) and \(\mathfrak{sl}_3(K)\),
 771 it is perhaps unsurprising that this condition is also sufficient.
 772 
 773 \begin{theorem}\label{thm:dominant-weight-theo}\index{weights!Highest Weight Theorem}
 774   For each dominant integral \(\lambda \in P^+\) there exists precisely one
 775   finite-dimensional simple \(\mathfrak{g}\)-module \(M\) whose highest weight
 776   is \(\lambda\).
 777 \end{theorem}
 778 
 779 This is known as \emph{the Highest Weight Theorem}, and its proof is the focus
 780 of this section. The ``uniqueness'' part of the theorem follows at once from
 781 the arguments used for \(\mathfrak{sl}_3(K)\). However, the ``existence'' part
 782 of the theorem is more nuanced. Our first instinct is, of course, to try to
 783 generalize the proof used for \(\mathfrak{sl}_3(K)\). Indeed, as in
 784 Proposition~\ref{thm:sl3-mod-is-highest-weight}, one is able to show\dots
 785 
 786 \begin{proposition}\label{thm:fin-dim-simple-mod-has-singular-vector}
 787   Let \(M\) be a finite-dimensional simple \(\mathfrak{g}\)-module. Then there
 788   exists a nonzero weight vector \(m \in M\) which is annihilated by all
 789   positive root spaces of \(\mathfrak{g}\) -- i.e. \(X \cdot m = 0\) for all
 790   \(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\).
 791 \end{proposition}
 792 
 793 \begin{proof}
 794   If \(\lambda\) is the highest weight of \(M\), it suffices to take any \(m
 795   \in M_\lambda\). Indeed, given \(X \in \mathfrak{g}_\alpha\) with \(\alpha
 796   \in \Delta^+\), \(X \cdot m \in M_{\lambda + \alpha} = 0\) because \(\lambda
 797   + \alpha \succ \lambda\).
 798 \end{proof}
 799 
 800 Unfortunately for us, this is where the parallels with
 801 Proposition~\ref{thm:sl3-mod-is-highest-weight} end. The issue is that our
 802 proof relied heavily on our knowledge of the roots of \(\mathfrak{sl}_3(K)\).
 803 It is thus clear that we need a more systematic approach for the general
 804 setting. We begin by asking a simpler question: how can we construct \emph{any}
 805 \(\mathfrak{g}\)-module \(M\) whose highest weight is \(\lambda\)? In the
 806 process of answering this question we will come across a surprisingly elegant
 807 solution to our problem.
 808 
 809 If \(M\) is a finite-dimensional simple module with highest weight \(\lambda\)
 810 and \(m \in M_\lambda\), we already know that \(X \cdot m = 0\) for any \(m \in
 811 M_\lambda\) and \(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\). Since
 812 \(M = \mathcal{U}(\mathfrak{g}) \cdot m\), the restriction of \(M\) to the
 813 Borel subalgebra \(\mathfrak{b} \subset \mathfrak{g}\) has a prescribed action.
 814 On the other hand, we have essentially no information about the action of the
 815 rest of \(\mathfrak{g}\) on \(M\). Nevertheless, given a
 816 \(\mathfrak{b}\)-module we may obtain a \(\mathfrak{g}\)-module by
 817 \emph{freely} extending the action of \(\mathfrak{b}\) via induction. This
 818 leads us to the following definition.
 819 
 820 \begin{definition}\label{def:verma}\index{\(\mathfrak{g}\)-module!(generalized) Verma modules}
 821   Given \(\lambda \in \mathfrak{h}^*\), consider the \(\mathfrak{b}\)-module
 822   \(K m^+\) where \(H \cdot m^+ = \lambda(H) m^+\) for all \(H \in
 823   \mathfrak{h}\) and \(X \cdot m^+ = 0\) for \(X \in \mathfrak{g}_{\alpha}\)
 824   with \(\alpha \in \Delta^+\). The \(\mathfrak{g}\)-module \(M(\lambda) =
 825   \operatorname{Ind}_{\mathfrak{b}}^{\mathfrak{g}} K m^+\) is called \emph{the
 826   Verma module of weight \(\lambda\)}.
 827 \end{definition}
 828 
 829 \begin{example}\label{ex:sl2-verma}
 830   If \(\mathfrak{g} = \mathfrak{sl}_2(K)\), then we can take \(\mathfrak{h} = K
 831   h\) and \(\mathfrak{b} = K e \oplus K h\). In this setting, the linear map
 832   \(g : \mathfrak{h}^* \to K\) defined by \(g(h) = 1\) affords us a canonical
 833   identification \(\mathfrak{h}^* = K g \cong K\), so that given \(\lambda \in
 834   K\) we may denote \(M(\lambda g)\) simply by \(M(\lambda)\). Using this
 835   notation \(M(\lambda) = \bigoplus_{k \ge 0} K f^k \cdot m^+\), and the action
 836   of \(\mathfrak{sl}_2(K)\) on \(M(\lambda)\) is given by
 837   formula (\ref{eq:sl2-verma-formulas}).
 838   \begin{equation}\label{eq:sl2-verma-formulas}
 839     \begin{aligned}
 840       f^k \cdot m^+ & \overset{e}{\mapsto} k(\lambda+1-k) f^{k-1} \cdot m^+ &
 841       f^k \cdot m^+ & \overset{f}{\mapsto} f^{k+1} \cdot m^+                &
 842       f^k \cdot m^+ & \overset{h}{\mapsto} (\lambda - 2k) f^k \cdot m^+     &
 843     \end{aligned}
 844   \end{equation}
 845 \end{example}
 846 
 847 \begin{example}\label{ex:verma-is-not-irr}
 848   Consider the \(\mathfrak{sl}_2(K)\)-module \(M(2)\) as described in
 849   Example~\ref{ex:sl2-verma}. It follows from formula
 850   (\ref{eq:sl2-verma-formulas}) that the action of \(\mathfrak{sl}_2(K)\) on
 851   \(M(2)\) is given by
 852   \begin{center}
 853     \begin{tikzcd}
 854       \cdots      \rar[bend left=60]{-10}
 855       & M(2)_{-6} \rar[bend left=60]{-4}  \lar[bend left=60]{1}
 856       & M(2)_{-4} \rar[bend left=60]{0}   \lar[bend left=60]{1}
 857       & M(2)_{-2} \rar[bend left=60]{2}   \lar[bend left=60]{1}
 858       & M(2)_0    \rar[bend left=60]{2}   \lar[bend left=60]{1}
 859       & M(2)_2                            \lar[bend left=60]{1}
 860     \end{tikzcd}
 861   \end{center}
 862   where \(M(2)_{2 - 2 k} = K f^k \cdot m^+\). Here the top arrows represent the
 863   action of \(e\) and the bottom arrows represent the action of \(f\). The
 864   scalars labeling each arrow indicate to which multiple of \(f^{k \pm 1} \cdot
 865   m^+\) the elements \(e\) and \(f\) send \(f^k \cdot m^+\). The string of
 866   weight spaces to the left of the diagram is infinite. Since \(e \cdot (f^3
 867   \cdot m^+) = 0\), it is easy to see that subspace \(\bigoplus_{k \ge 3} K f^k
 868   \cdot m^+\) is a (maximal) \(\mathfrak{sl}_2(K)\)-submodule, which is
 869   isomorphic to \(M(-4)\).
 870 \end{example}
 871 
 872 These last examples show that, unlike most modules we have so far encountered,
 873 Verma modules are \emph{highly infinite-dimensional}. Indeed, it follows from
 874 the PBW Theorem that the regular module \(\mathcal{U}(\mathfrak{g})\) is a free
 875 \(\mathfrak{b}\)-module of infinite rank -- equal to the codimension of
 876 \(\mathcal{U}(\mathfrak{b})\) in \(\mathcal{U}(\mathfrak{g})\). Hence \(\dim
 877 M(\lambda)\), which is the same as the rank of \(\mathcal{U}(\mathfrak{g})\) as
 878 a \(\mathfrak{b}\)-module, is also infinite. Nevertheless, it turns out that
 879 finite-dimensional modules and Verma module may both be seen as particular
 880 cases of a more general pattern. This leads us to the following definitions.
 881 
 882 \begin{definition}
 883   Let \(M\) be a \(\mathfrak{g}\)-module. A vector \(m \in M\) is called
 884   \emph{singular} if it is annihilated by all positive weight spaces of
 885   \(\mathfrak{g}\) -- i.e. \(X \cdot m = 0\) for all \(X \in
 886   \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\).
 887 \end{definition}
 888 
 889 \begin{definition}\label{def:highest-weight-mod}
 890   A \(\mathfrak{g}\)-module \(M\) is called \emph{a highest weight module} if
 891   there exists some singular weight vector \(m^+ \in M_\lambda\) such that \(M
 892   = \mathcal{U}(\mathfrak{g}) \cdot m^+\). Any such \(m^+\) is called \emph{a
 893   highest weight vector}, while \(\lambda\) is called \emph{the highest weight
 894   of \(M\)}.
 895 \end{definition}
 896 
 897 \begin{example}
 898   Proposition~\ref{thm:fin-dim-simple-mod-has-singular-vector} is equivalent to
 899   the fact that every finite-dimensional simple \(\mathfrak{g}\)-module is a
 900   highest weight module.
 901 \end{example}
 902 
 903 \begin{example}
 904   It should be obvious from the definitions that \(M(\lambda)\) is a highest
 905   weight module of highest weight \(\lambda\) and highest weight vector \(m^+ =
 906   1 \otimes m^+\) as in Definition~\ref{def:verma}. Indeed, \(u \otimes m^+ = u
 907   \cdot m^+\) for all \(u \in \mathcal{U}(\mathfrak{g})\), which already shows
 908   \(M(\lambda)\) is generated by \(m^+\). In particular,
 909   \begin{align*}
 910     H \cdot m^+ & = H \otimes m^+ = 1 \otimes H \cdot m^+ = \lambda(H) m^+ \\
 911     X \cdot m^+ & = X \otimes m^+ = 1 \otimes X \cdot m^+ = 0
 912   \end{align*}
 913   for all \(H \in \mathfrak{h}\) and \(X \in \mathfrak{g}_\alpha\), \(\alpha
 914   \in \Delta^+\).
 915 \end{example}
 916 
 917 While Verma modules show that a highest weight module needs not to be
 918 finite-dimensional, it turns out that highest weight modules enjoy many of the
 919 features we've grown used to in the past chapters. Explicitly, we may establish
 920 the properties described in the following proposition, whose statement should
 921 also explain the nomenclature of Definition~\ref{def:highest-weight-mod}.
 922 
 923 \begin{proposition}\label{thm:high-weight-mod-is-weight-mod}
 924   Let \(M\) be a highest weight \(\mathfrak{g}\)-module with highest weight
 925   vector \(m \in M_\lambda\). The weight spaces decomposition
 926   \[
 927     M = \bigoplus_{\mu \in \mathfrak{h}^*} M_\mu
 928   \]
 929   holds. Furthermore, \(\dim M_\mu < \infty\) for all \(\mu \in
 930   \mathfrak{h}^*\) and \(\dim M_\lambda = 1\) -- i.e. \(M_\lambda = K m\).
 931   Finally, given a weight \(\mu\) of \(M\), \(\lambda \succeq \mu\) -- so that
 932   the highest weight \(\lambda\) of \(M\) is unique and coincides with the
 933   largest of the weights of \(M\).
 934 \end{proposition}
 935 
 936 \begin{proof}
 937   Since \(M = \mathcal{U}(\mathfrak{g}) \cdot m\), the PBW Theorem implies
 938   that \(M\) is spanned by the vectors \(F_{\alpha_{i_1}} F_{\alpha_{i_2}}
 939   \cdots F_{\alpha_{i_s}} \cdot m\) for \(\Delta^+ = \{\alpha_1, \ldots,
 940   \alpha_r\}\) and \(F_{\alpha_i} \in \mathfrak{g}_{- \alpha_i}\) as in the
 941   proof of Proposition~\ref{thm:distinguished-subalgebra}. But
 942   \[
 943     \begin{split}
 944       H \cdot (F_{\alpha_{i_1}} F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}}
 945                \cdot m)
 946       & = ([H, F_{\alpha_{i_1}}] + F_{\alpha_{i_1}} H)
 947           F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m \\
 948       & = - \alpha_{i_1}(H) F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m
 949         + F_{\alpha_{i_1}} ([H, F_{\alpha_{i_2}}] + F_{\alpha_{i_2}} H)
 950           F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m \\
 951       & \;\; \vdots \\
 952       & = (- \alpha_{i_1} - \cdots - \alpha_{i_s})(H)
 953           F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m
 954         + F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} H \cdot m \\
 955       & = (\lambda - \alpha_{i_1} - \cdots - \alpha_{i_s})(H)
 956           F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m \\
 957       & \therefore F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m
 958         \in M_{\lambda - \alpha_{i_1} - \cdots - \alpha_{i_s}}
 959     \end{split}
 960   \]
 961 
 962   Hence \(M \subset \bigoplus_{\mu \in \mathfrak{h}^*} M_\mu\), as desired. In
 963   fact we have established
 964   \[
 965     M
 966     \subset
 967     \bigoplus_{k_i \in \mathbb{N}}
 968     M_{\lambda - k_1 \cdot \alpha_1 - \cdots - k_r \cdot \alpha_r}
 969   \]
 970   where \(\{\alpha_1, \ldots, \alpha_r\} = \Delta^+\), so that all weights of
 971   \(M\) have the form \(\mu = \lambda - k_1 \cdot \alpha_1 - \cdots - k_r \cdot
 972   \alpha_r\). This already gives us that the weights of \(M\) are bounded by
 973   \(\lambda\).
 974 
 975   To see that \(\dim M_\mu < \infty\), simply note that there are only finitely
 976   many monomials \(F_{\alpha_1}^{k_1} F_{\alpha_2}^{k_2} \cdots
 977   F_{\alpha_s}^{k_s}\) such that \(\mu = \lambda + k_1 \cdot \alpha_1 + \cdots
 978   + k_s \cdot \alpha_s\). Since \(M_\mu\) is spanned by the images of \(m\)
 979   under such monomials, we conclude \(\dim M_\mu < \infty\). In particular,
 980   there is a single monomial \(F_{\alpha_1}^{k_1} F_{\alpha_2}^{k_2} \cdots
 981   F_{\alpha_s}^{k_s}\) such that \(\lambda = \lambda + k_1 \cdot \alpha_1 +
 982   \cdots + k_s \cdot \alpha_s\) -- which is, of course, the monomial where
 983   \(k_1 = \cdots = k_n = 0\). Hence \(\dim M_\lambda = 1\).
 984 \end{proof}
 985 
 986 At this point it is important to note that, far from a ``misbehaved'' class of
 987 examples, Verma modules hold a very special place in the theory of highest
 988 weight modules. Intuitively speaking, the Verma module \(M(\lambda)\) should
 989 really be though-of as ``the freest highest weight \(\mathfrak{g}\)-module of
 990 highest weight \(\lambda\)''. In practice, this translates to the following
 991 universal property.
 992 
 993 \begin{proposition}
 994   Let \(M\) be a \(\mathfrak{g}\)-module and \(m \in M_\lambda\) be a singular
 995   vector. Then there exists a unique \(\mathfrak{g}\)-homomorphism \(f :
 996   M(\lambda) \to M\) such that \(f(m^+) = m\). Furthermore, all homomorphisms
 997   \(M(\lambda) \to M\) are given in this fashion.
 998   \[
 999     \operatorname{Hom}_{\mathfrak{g}}(M(\lambda), M)
1000     \cong \{ m \in M_\lambda : m \ \text{is singular}\}
1001   \]
1002 \end{proposition}
1003 
1004 \begin{proof}
1005   The result follows directly from Proposition~\ref{thm:frobenius-reciprocity}.
1006   Indeed, by the Frobenius Reciprocity Theorem, a \(\mathfrak{g}\)-homomorphism
1007   \(f : M(\lambda) \to M\) is the same as a \(\mathfrak{b}\)-homomorphism \(g :
1008   K m^+ \to M = \operatorname{Res}_{\mathfrak{b}}^{\mathfrak{g}} M\). More
1009   specifically, given a \(\mathfrak{b}\)-homomorphism \(g : K m^+ \to M\),
1010   there exists a unique \(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\)
1011   such that \(f(u \otimes m^+) = u \cdot g(m^+)\) for all \(u \in
1012   \mathcal{U}(\mathfrak{g})\), and all \(\mathfrak{g}\)-homomorphism
1013   \(M(\lambda) \to M\) arise in this fashion.
1014 
1015   Any \(K\)-linear map \(g : K m^+ \to M\) is determined by \(m = g(m^+)\).
1016   Finally, notice that \(g\) is a \(\mathfrak{b}\)-homomorphism if, and only if
1017   \(m\) is a singular vector lying in \(M_\lambda\).
1018 \end{proof}
1019 
1020 Why is any of this interesting to us, however? After all, Verma modules are not
1021 specially well suited candidates for a proof of the Highest Weight Theorem.
1022 Indeed, we have seen in Example~\ref{ex:verma-is-not-irr} that in general
1023 \(M(\lambda)\) is not simple, nor is it ever finite-dimensional. Nevertheless,
1024 we may use \(M(\lambda)\) to establish Theorem~\ref{thm:dominant-weight-theo}
1025 as follows.
1026 
1027 Suppose \(M\) is a highest weight \(\mathfrak{g}\)-module of highest weight
1028 \(\lambda\) with highest weight vector \(m\). By the last proposition, there is
1029 a \(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\) such that \(f(m^+) =
1030 m\). Since \(M = \mathcal{U}(\mathfrak{g}) \cdot m\), \(f\) is surjective and
1031 therefore \(M \cong \mfrac{M(\lambda)}{\ker f}\). Hence\dots
1032 
1033 \begin{proposition}
1034   Let \(M\) be a highest weight \(\mathfrak{g}\)-module of highest weight
1035   \(\lambda\). Then \(M\) is quotient of \(M(\lambda)\). If \(M\) is simple
1036   then \(M\) is the quotient of \(M(\lambda)\) by a maximal
1037   \(\mathfrak{g}\)-submodule.
1038 \end{proposition}
1039 
1040 Maximal submodules of Verma modules are thus of primary interest to us. As it
1041 turns out, these can be easily classified.
1042 
1043 \begin{proposition}\label{thm:max-verma-submod-is-weight}
1044   Every submodule \(N \subset M(\lambda)\) is the direct sum of its weight
1045   spaces. In particular, \(M(\lambda)\) has a unique maximal submodule
1046   \(N(\lambda)\) and a unique simple quotient \(L(\lambda) =
1047   \sfrac{M(\lambda)}{N(\lambda)}\). Any simple highest weight
1048   \(\mathfrak{g}\)-module has the form \(L(\lambda)\) for some unique \(\lambda
1049   \in \mathfrak{h}^*\).
1050 \end{proposition}
1051 
1052 \begin{proof}
1053   Let \(N \subset M(\lambda)\) be a submodule and take any nonzero \(n \in N\).
1054   Because of Proposition~\ref{thm:high-weight-mod-is-weight-mod}, we know there
1055   are \(\mu_1, \ldots, \mu_r \in \mathfrak{h}^*\) and nonzero \(m_i \in
1056   M(\lambda)_{\mu_i}\) such that \(n = m_1 + \cdots + m_r\). We want to show
1057   \(m_i \in N\) for all \(i\).
1058 
1059   Fix some \(H_2 \in \mathfrak{h}\) such that \(\mu_1(H_2) \ne \mu_2(H_2)\).
1060   Then
1061   \[
1062     m_1
1063     - \frac{(\mu_3 - \mu_1)(H_2)}{(\mu_2 - \mu_1)(H_2)} \cdot m_3
1064     - \cdots
1065     - \frac{(\mu_r - \mu_1)(H_2)}{(\mu_2 - \mu_1)(H_2)} \cdot m_r
1066     = \left( 1 - \frac{H_2 - \mu_1(H_2)}{(\mu_2 - \mu_1)(H_2)} \right) \cdot n
1067     \in N
1068   \]
1069 
1070   Now take \(H_3 \in \mathfrak{h}\) such that \(\mu_1(H_3) \ne \mu_3(H_3)\). By
1071   applying the same procedure again we get
1072   \begin{multline*}
1073     m_1
1074     -
1075     \frac{(\mu_4 - \mu_3)(H_3) \cdot (\mu_4 - \mu_1)(H_2)}
1076          {(\mu_3 - \mu_1)(H_3) \cdot (\mu_2 - \mu_1)(H_2)} \cdot m_4
1077     - \cdots -
1078     \frac{(\mu_r - \mu_3)(H_3) \cdot (\mu_r - \mu_1)(H_2)}
1079          {(\mu_3 - \mu_1)(H_3) \cdot (\mu_2 - \mu_1)(H_2)} \cdot m_r \\
1080     =
1081     \left(1 - \frac{H_3 - \mu_1(H_3)}{(\mu_3 - \mu_1)(H_3)} \right)
1082     \left(1 - \frac{H_2 - \mu_1(H_2)}{(\mu_2 - \mu_1)(H_2)} \right) \cdot n
1083     \in N
1084   \end{multline*}
1085 
1086   By applying the same procedure over and over again we can see that \(m_1 = u
1087   \cdot n \in N\) for some \(u \in \mathcal{U}(\mathfrak{g})\). Furthermore, if
1088   we reproduce all this for \(m_2 + \cdots + m_r = n - m_1 \in N\) we get that
1089   \(m_2 \in N\). All in all we find \(m_1, \ldots, m_r \in N\). Hence
1090   \[
1091     N = \bigoplus_\mu N_\mu = \bigoplus_\mu M(\lambda)_\mu \cap N
1092   \]
1093 
1094   Since \(M(\lambda) = \mathcal{U}(\mathfrak{g}) \cdot m^+\), if \(N\) is a
1095   proper submodule then \(m^+ \notin N\). Hence any proper submodule lies in
1096   the sum of weight spaces other than \(M(\lambda)_\lambda\), so the sum
1097   \(N(\lambda)\) of all such submodules is still proper. This implies
1098   \(N(\lambda)\) is the unique maximal submodule of \(M(\lambda)\) and
1099   \(L(\lambda) = \sfrac{M(\lambda)}{N(\lambda)}\) is its unique simple
1100   quotient.
1101 \end{proof}
1102 
1103 \begin{corollary}\label{thm:classification-of-simple-high-weight-mods}
1104   Let \(M\) be a simple weight \(\mathfrak{g}\)-module of weight \(\lambda\).
1105   Then \(M \cong L(\lambda)\).
1106 \end{corollary}
1107 
1108 We thus know that \(L(\lambda)\) is the only possible candidate for the
1109 \(\mathfrak{g}\)-module \(M\) in the statement of
1110 Theorem~\ref{thm:dominant-weight-theo}. We should also note that our past
1111 examples indicate that \(L(\lambda)\) does fulfill its required role.
1112 Indeed\dots
1113 
1114 \begin{example}\label{ex:sl2-verma-quotient}
1115   Consider the \(\mathfrak{sl}_2(K)\) module \(M(2)\) as described in
1116   Example~\ref{ex:sl2-verma}. We can see from Example~\ref{ex:verma-is-not-irr}
1117   that \(N(2) = \bigoplus_{k \ge 3} K f^k \cdot m^+\), so that \(L(2)\) is the
1118   \(3\)-dimensional simple \(\mathfrak{sl}_2(K)\)-module -- i.e. the
1119   finite-dimensional simple module with highest weight \(2\) constructed in
1120   chapter~\ref{ch:sl3}.
1121 \end{example}
1122 
1123 All its left to prove the Highest Weight Theorem is verifying that the
1124 situation encountered in Example~\ref{ex:sl2-verma-quotient} holds for any
1125 dominant integral \(\lambda \in P^+\). In other words, we need to show\dots
1126 
1127 \begin{proposition}\label{thm:verma-is-finite-dim}
1128   If \(\mathfrak{g}\) is semisimple and \(\lambda\) is dominant integral then
1129   the unique simple quotient \(L(\lambda)\) of \(M(\lambda)\) is
1130   finite-dimensional.
1131 \end{proposition}
1132 
1133 The proof of Proposition~\ref{thm:verma-is-finite-dim} is very technical and we
1134 won't include it here, but the idea behind it is to show that the set of
1135 weights of \(L(\lambda)\) is stable under the natural action of the Weyl group
1136 \(W\) on \(\mathfrak{h}^*\). One can then show that the every weight
1137 of \(L(\lambda)\) is conjugate to a single dominant integral weight of
1138 \(L(\lambda)\), and that the set of dominant integral weights of \(L(\lambda)\)
1139 is finite. Since \(W\) is finitely generated, this implies the set of
1140 weights of the unique simple quotient of \(M(\lambda)\) is finite. But
1141 each weight space is finite-dimensional. Hence so is the simple quotient
1142 \(L(\lambda)\).
1143 
1144 We refer the reader to \cite[ch. 21]{humphreys} for further details. We are now
1145 ready to prove the Highest Weight Theorem.
1146 
1147 \begin{proof}[Proof of Theorem~\ref{thm:dominant-weight-theo}]
1148   We begin by the ``existence'' part of the theorem. Let \(\lambda\) be a
1149   dominant integral weight of \(\mathfrak{g}\). Since \(\dim L(\lambda) <
1150   \infty\), all its left is to show that \(M = L(\lambda)\) is indeed a highest
1151   weight module of highest weight \(\lambda\). It is clear from the definitions
1152   that \(m^+ + N(\lambda) \in L(\lambda)_\lambda\) is singular and generates
1153   all of \(L(\lambda)\). Hence it suffices to show that \(m^+ + N(\lambda)\) is
1154   nonzero. But this is the same as checking that \(m^+ \notin N(\lambda)\),
1155   which is also clear from the previous definitions. As for the uniqueness of
1156   \(M\), it suffices to apply
1157   Corollary~\ref{thm:classification-of-simple-high-weight-mods}.
1158 \end{proof}
1159 
1160 We would now like to conclude this chapter by describing the situation where
1161 \(\lambda \notin P^+\). We begin by pointing out that
1162 Proposition~\ref{thm:verma-is-finite-dim} fails in the general setting. For
1163 instance, consider\dots
1164 
1165 \begin{example}\label{ex:antidominant-verma}
1166   The action of \(\mathfrak{sl}_2(K)\) on \(M(-4)\) is given by the following
1167   diagram. In general, it is possible to check using formula
1168   (\ref{eq:sl2-verma-formulas}) that \(e\) always maps \(f^{k + 1} \cdot m^+\)
1169   to a nonzero multiple of \(f^k \cdot m^+\), so we can see that \(M(-4)\) has
1170   no proper submodules, \(N(-4) = 0\) and thus \(L(-4) \cong M(-4)\).
1171   \begin{center}
1172     \begin{tikzcd}
1173       \cdots         \rar[bend left=60]{-28}
1174       & M(-4)_{-10}  \rar[bend left=60]{-18} \lar[bend left=60]{1}
1175       & M(-4)_{-8}   \rar[bend left=60]{-10} \lar[bend left=60]{1}
1176       & M(-4)_{-6}   \rar[bend left=60]{-4}  \lar[bend left=60]{1}
1177       & M(-4)_{-4}                           \lar[bend left=60]{1}
1178     \end{tikzcd},
1179   \end{center}
1180 \end{example}
1181 
1182 While \(L(\lambda)\) is always a highest weight module of highest weight
1183 \(\lambda\), we can easily see that if \(\lambda \notin P^+\) then
1184 \(L(\lambda)\) is infinite-dimensional. Indeed, this is precisely the
1185 counterpositive of Proposition~\ref{thm:highes-weight-of-fin-dim-is-dominant}!
1186 If \(\lambda = k_1 \beta_1 + \cdots + k_r \beta_r \in P\) is integral and \(k_i
1187 < 0\) for all \(i\), then one is additionally able to show that \(M(\lambda)
1188 \cong L(\lambda)\) as in Example~\ref{ex:antidominant-verma}. Verma modules can
1189 thus serve as examples of infinite-dimensional simple modules.
1190 
1191 In the next chapter we expand our previous results by exploring the question:
1192 what are \emph{all} the infinite-dimensional simple \(\mathfrak{g}\)-modules?