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

Commit
9bbab87c16919043a2db660c996645b7bbfd745e
Parent
7abeda5bc9d129a3946bb2cb668dcf4cae777625
Author
Pablo <pablo-escobar@riseup.net>
Date

Renamed some files

Diffstats

5 files changed, 2657 insertions, 2657 deletions

Status Name Changes Insertions Deletions
Added sections/fin-dim-simple.tex 1 file changed 1044 0
Deleted sections/mathieu.tex 1 file changed 0 1611
Deleted sections/semisimple-algebras.tex 1 file changed 0 1044
Added sections/simple-weight.tex 1 file changed 1611 0
Modified tcc.tex 2 files changed 2 2
diff --git /dev/null b/sections/fin-dim-simple.tex
@@ -0,0 +1,1044 @@
+\chapter{Finite-Dimensional Simple Modules}
+
+In this chapter we classify the finite-dimensional simple
+\(\mathfrak{g}\)-modules for a finite-dimensional semisimple Lie algebra
+\(\mathfrak{g}\) over \(K\). At the heart of our analysis of
+\(\mathfrak{sl}_2(K)\) and \(\mathfrak{sl}_3(K)\) was the decision to consider
+the eigenspace decomposition
+\begin{equation}\label{sym-diag}
+  M = \bigoplus_\lambda M_\lambda
+\end{equation}
+
+This was simple enough to do in the case of \(\mathfrak{sl}_2(K)\), but the
+rational behind it and the reason why equation (\ref{sym-diag}) holds are
+harder to explain in the case of \(\mathfrak{sl}_3(K)\). The eigenspace
+decomposition associated with an operator \(M \to M\) is a very well-known
+tool, and readers familiarized with basic concepts of linear algebra should be
+used to this type of argument. On the other hand, the eigenspace decomposition
+of \(M\) with respect to the action of an arbitrary subalgebra \(\mathfrak{h}
+\subset \mathfrak{gl}(M)\) is neither well-known nor does it hold in general:
+as indicated in the previous chapter, it may very well be that
+\[
+  \bigoplus_{\lambda \in \mathfrak{h}^*} M_\lambda \subsetneq M
+\]
+
+We should note, however, that these two cases are not as different as they may
+sound at first glance. Specifically, we can regard the eigenspace decomposition
+of a \(\mathfrak{sl}_2(K)\)-module \(M\) with respect to the eigenvalues of the
+action of \(h\) as the eigenvalue decomposition of \(M\) with respect to the
+action of the subalgebra \(\mathfrak{h} = K h \subset \mathfrak{sl}_2(K)\).
+Furthermore, in both cases \(\mathfrak{h} \subset \mathfrak{sl}_n(K)\) is the
+subalgebra of diagonal matrices, which is Abelian. The fundamental difference
+between these two cases is thus the fact that \(\dim \mathfrak{h} = 1\) for
+\(\mathfrak{h} \subset \mathfrak{sl}_2(K)\) while \(\dim \mathfrak{h} > 1\) for
+\(\mathfrak{h} \subset \mathfrak{sl}_3(K)\). The question then is: why did we
+choose \(\mathfrak{h}\) with \(\dim \mathfrak{h} > 1\) for
+\(\mathfrak{sl}_3(K)\)?
+
+The rational behind fixing an Abelian subalgebra \(\mathfrak{h}\) is a simple
+one: we have seen in the previous chapter that representations of Abelian
+algebras are generally much simpler to understand than the general case. Thus
+it make sense to decompose a given \(\mathfrak{g}\)-module \(M\) of into
+subspaces invariant under the action of \(\mathfrak{h}\), and then analyze how
+the remaining elements of \(\mathfrak{g}\) act on these subspaces. The bigger
+\(\mathfrak{h}\) is, the simpler our problem gets, because there are fewer
+elements outside of \(\mathfrak{h}\) left to analyze.
+
+Hence we are generally interested in maximal Abelian subalgebras \(\mathfrak{h}
+\subset \mathfrak{g}\), which leads us to the following definition.
+
+\begin{definition}\index{Cartan subalgebra}
+  A subalgebra \(\mathfrak{h} \subset \mathfrak{g}\) is called \emph{a Cartan
+  subalgebra of \(\mathfrak{g}\)} if is self-normalizing -- i.e. \([X, H] \in
+  \mathfrak{h}\) for all \(H \in \mathfrak{h}\) if, and only if \(X \in
+  \mathfrak{h}\) -- and nilpotent. Equivalently for reductive \(\mathfrak{g}\),
+  \(\mathfrak{h}\) is called \emph{a Cartan subalgebra of \(\mathfrak{g}\)} if
+  it is Abelian, \(\operatorname{ad}(H)\) is diagonalizable for each \(H \in
+  \mathfrak{h}\) and if \(\mathfrak{h}\) is maximal with respect to the former
+  two properties.
+\end{definition}
+
+\begin{proposition}
+  There exists a Cartan subalgebra \(\mathfrak{h} \subset \mathfrak{g}\).
+\end{proposition}
+
+\begin{proof}
+  Notice that \(0 \subset \mathfrak{g}\) is an Abelian subalgebra whose
+  elements act as diagonal operators via the adjoint \(\mathfrak{g}\)-module.
+  Indeed, \(0\), the only element of \(0 \subset \mathfrak{g}\), is such that
+  \(\operatorname{ad}(0) = 0\) is a diagonalizable operator. Furthermore, given
+  a chain of Abelian subalgebras
+  \[
+    0 \subset \mathfrak{h}_1 \subset \mathfrak{h}_2 \subset \cdots
+  \]
+  such that \(\operatorname{ad}(H)\) is a diagonal operator for each \(H \in
+  \mathfrak{h}_i\), the subalgebra \(\bigcup_i \mathfrak{h}_i \subset
+  \mathfrak{g}\) is Abelian, and its elements also act diagonally in
+  \(\mathfrak{g}\). It then follows from Zorn's Lemma that there exists a
+  subalgebra \(\mathfrak{h}\) which is maximal with respect to both these
+  properties, also known as a Cartan subalgebra.
+\end{proof}
+
+We have already seen some concrete examples. Namely\dots
+
+\begin{example}\label{ex:cartan-of-gl}
+  The Lie subalgebra
+  \[
+    \mathfrak{h} =
+    \begin{pmatrix}
+           K &      0 & \cdots &      0 \\
+           0 &      K & \cdots &      0 \\
+      \vdots & \vdots & \ddots & \vdots \\
+           0 &      0 & \cdots &      K
+    \end{pmatrix}
+    \subset \mathfrak{gl}_n(K)
+  \]
+  of diagonal matrices is a Cartan subalgebra.
+  Indeed, every pair of diagonal matrices commutes, so that \(\mathfrak{h}\)
+  is an Abelian -- and hence nilpotent -- subalgebra. A
+  simple calculation also shows that if \(i \ne j\) then the coefficient of
+  \(E_{i j}\) in \([E_{i i}, X]\) is the same as the coefficient of \(E_{i j}\)
+  in \(X\), for all \(X \in \mathfrak{gl}_n(K)\). In particular, if \([E_{i i},
+  X]\) is diagonal for all \(i\), then so is \(X\) -- i.e. \(\mathfrak{h}\) is
+  self-normalizing.
+\end{example}
+
+\begin{example}
+  Let \(\mathfrak{h}\) be as in Example~\ref{ex:cartan-of-gl}. Then the
+  subalgebra \(\mathfrak{h} \cap \mathfrak{sl}_n(K)\) of traceless diagonal
+  matrices is a Cartan subalgebra of \(\mathfrak{sl}_n(K)\).
+\end{example}
+
+\begin{example}\label{ex:cartan-direct-sum}
+  Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras and
+  \(\mathfrak{h}_i \subset \mathfrak{g}_i\) be Cartan subalgebras. Then
+  \(\mathfrak{h}_1 \oplus \mathfrak{h}_2\) is a Cartan subalgebra of
+  \(\mathfrak{g}_1 \oplus \mathfrak{g}_2\).
+\end{example}
+
+\index{Cartan subalgebra!simultaneous diagonalization}
+The intersection of such subalgebra with \(\mathfrak{sl}_n(K)\) -- i.e. the
+subalgebra of traceless diagonal matrices -- is a Cartan subalgebra of
+\(\mathfrak{sl}_n(K)\). In particular, if \(n = 2\) or \(n = 3\) we get to the
+subalgebras described the previous chapter. The remaining question then is: if
+\(\mathfrak{h} \subset \mathfrak{g}\) is a Cartan subalgebra and \(M\) is a
+\(\mathfrak{g}\)-module, does the eigenspace decomposition
+\[
+  M = \bigoplus_\lambda M_\lambda
+\]
+of \(M\) hold? The answer to this question turns out to be yes. This is a
+consequence of something known as \emph{simultaneous diagonalization}, which is
+the primary tool we will use to generalize the results of the previous section.
+What is simultaneous diagonalization all about then?
+
+\begin{definition}\label{def:sim-diag}
+  Given a \(K\)-vector space \(V\), a set of operators \(\{T_j : V \to V\}_j\)
+  is called \emph{simultaneously diagonalizable} if there is a basis \(\{v_1,
+  \ldots, v_n\}\) for \(V\) such that \(T_j v_i\) is a scalar multiple of
+  \(v_i\), for all \(i, j\).
+\end{definition}
+
+\begin{proposition}
+  Given a \emph{finite-dimensional} vector space \(V\), a set of diagonalizable
+  operators \(V \to V\) is simultaneously diagonalizable if, and only if all of
+  its elements commute with one another.
+\end{proposition}
+
+We should point out that simultaneous diagonalization \emph{only works in the
+finite-dimensional setting}. In fact, simultaneous diagonalization is usually
+framed as an equivalent statement about diagonalizable \(n \times n\) matrices.
+Simultaneous diagonalization implies that to show \(M = \bigoplus_\lambda
+M_\lambda\) it suffices to show that \(H\!\restriction_M : M \to M\) is a
+diagonalizable operator for each \(H \in \mathfrak{h}\). To that end, we
+introduce \emph{the Jordan decomposition of an operator} and \emph{the abstract
+Jordan decomposition of a semisimple Lie algebra}.
+
+\begin{proposition}[Jordan]
+  Given a finite-dimensional vector space \(V\) and an operator \(T : V \to
+  V\), there are unique commuting operators \(T_{\operatorname{ss}},
+  T_{\operatorname{nil}} : V \to V\), with \(T_{\operatorname{ss}}\)
+  diagonalizable and \(T_{\operatorname{nil}}\) nilpotent, such that \(T =
+  T_{\operatorname{ss}} + T_{\operatorname{nil}}\). The pair
+  \((T_{\operatorname{ss}}, T_{\operatorname{nil}})\) is known as \emph{the Jordan
+  decomposition of \(T\)}.
+\end{proposition}
+
+\begin{proposition}\index{abstract Jordan decomposition}
+  Given \(\mathfrak{g}\) semisimple and \(X \in \mathfrak{g}\), there are
+  \(X_{\operatorname{ss}}, X_{\operatorname{nil}} \in \mathfrak{g}\) such that \(X
+  = X_{\operatorname{ss}} + X_{\operatorname{nil}}\), \([X_{\operatorname{ss}},
+  X_{\operatorname{nil}}] = 0\), \(\operatorname{ad}(X_{\operatorname{ss}})\) is a
+  diagonalizable operator and \(\operatorname{ad}(X_{\operatorname{nil}})\) is a
+  nilpotent operator. The pair \((X_{\operatorname{ss}}, X_{\operatorname{nil}})\)
+  is known as \emph{the Jordan decomposition of \(X\)}.
+\end{proposition}
+
+It should be clear from the uniqueness of
+\(\operatorname{ad}(X)_{\operatorname{ss}}\) and
+\(\operatorname{ad}(X)_{\operatorname{nil}}\) that the Jordan decomposition of
+\(\operatorname{ad}(X)\) is \(\operatorname{ad}(X) =
+\operatorname{ad}(X_{\operatorname{ss}}) +
+\operatorname{ad}(X_{\operatorname{nil}})\). What is perhaps more remarkable is
+the fact this holds for \emph{any} finite-dimensional \(\mathfrak{g}\)-module.
+In other words\dots
+
+\begin{proposition}\label{thm:preservation-jordan-form}
+  Let \(M\) be a finite-dimensional \(\mathfrak{g}\)-module and \(X
+  \in \mathfrak{g}\). Denote by \(X\!\restriction_M\) the action of \(X\) on
+  \(M\). Then \(X_{\operatorname{ss}}\!\restriction_M =
+  (X\!\restriction_M)_{\operatorname{ss}}\) and
+  \(X_{\operatorname{nil}}\!\restriction_M =
+  (X\!\restriction_M)_{\operatorname{nil}}\).
+\end{proposition}
+
+This last result is known as \emph{the preservation of the Jordan form}, and a
+proof can be found in appendix C of \cite{fulton-harris}. As promised this
+implies\dots
+
+\begin{corollary}\label{thm:finite-dim-is-weight-mod}
+  Let \(\mathfrak{g}\) be a semisimple Lie algebra, \(\mathfrak{h} \subset
+  \mathfrak{g}\) be a Cartan subalgebra and \(M\) be any finite-dimensional
+  \(\mathfrak{g}\)-module. Then there is a basis \(\{m_1, \ldots,
+  m_r\}\) of \(M\) so that each \(m_i\) is simultaneously an eigenvector of all
+  elements of \(\mathfrak{h}\) -- i.e. each element of \(\mathfrak{h}\) acts as
+  a diagonal matrix in this basis. In other words, there are linear functionals
+  \(\lambda_i \in \mathfrak{h}^*\) so that
+  \(
+    H \cdot m_i = \lambda_i(H) m_i
+  \)
+  for all \(H \in \mathfrak{h}\). In particular,
+  \[
+    M = \bigoplus_{\lambda \in \mathfrak{h}^*} M_\lambda
+  \]
+\end{corollary}
+
+\begin{proof}
+  Fix some \(H \in \mathfrak{h}\). It suffices to show that \(H\!\restriction_M
+  : M \to M\) is a diagonalizable operator.
+
+  If we write \(H = H_{\operatorname{ss}} + H_{\operatorname{nil}}\) for the
+  abstract Jordan decomposition of \(H\), we know
+  \(\operatorname{ad}(H_{\operatorname{ss}}) =
+  \operatorname{ad}(H)_{\operatorname{ss}}\). But \(\operatorname{ad}(H)\) is a
+  diagonalizable operator, so that \(\operatorname{ad}(H)_{\operatorname{ss}} =
+  \operatorname{ad}(H)\). This implies
+  \(\operatorname{ad}(H_{\operatorname{nil}}) =
+  \operatorname{ad}(H)_{\operatorname{nil}} = 0\), so that
+  \(H_{\operatorname{nil}}\) is a central element of \(\mathfrak{g}\). Since
+  \(\mathfrak{g}\) is semisimple, \(H_{\operatorname{nil}} = 0\).
+  Proposition~\ref{thm:preservation-jordan-form} then implies
+  \((H\!\restriction_M)_{\operatorname{nil}} =
+  H_{\operatorname{nil}}\!\restriction_M = 0\), so \(H\!\restriction_M =
+  (H\!\restriction_M)_{\operatorname{ss}}\) is a diagonalizable operator.
+\end{proof}
+
+We should point out that this last proof only works for semisimple Lie
+algebras. This is because we rely heavily on
+Proposition~\ref{thm:preservation-jordan-form}, as well in the fact that
+semisimple Lie algebras are centerless. In fact,
+Corollary~\ref{thm:finite-dim-is-weight-mod} fails even for reductive Lie
+algebras. For a counterexample, consider the algebra \(\mathfrak{g} = K\): the
+Cartan subalgebra of \(\mathfrak{g}\) is \(\mathfrak{g}\) itself, and a
+\(\mathfrak{g}\)-module is simply a vector space \(M\) endowed with an operator
+\(M \to M\) -- which corresponds to the action of \(1 \in \mathfrak{g}\) on
+\(M\). In particular, if we choose an operator \(M \to M\) which is \emph{not}
+diagonalizable we find \(M \ne \bigoplus_{\lambda \in \mathfrak{h}^*}
+M_\lambda\).
+
+However, Corollary~\ref{thm:finite-dim-is-weight-mod} does work for reductive
+\(\mathfrak{g}\) if we assume that the \(\mathfrak{g}\)-module \(M\) in
+question is simple, since central elements of \(\mathfrak{g}\) act on simple
+\(\mathfrak{g}\)-modules as scalar operators. The hypothesis of
+finite-dimensionality is also of huge importance. For instance, consider\dots
+
+\begin{example}\label{ex:regular-mod-is-not-weight-mod}
+  Let \(\mathcal{U}(\mathfrak{g})\) denote the regular \(\mathfrak{g}\)-module.
+  Notice that \(\mathcal{U}(\mathfrak{g})_\lambda = 0\) for all \(\lambda \in
+  \mathfrak{h}^*\). Indeed, since \(\mathcal{U}(\mathfrak{g})\) is a domain, if
+  \((H - \lambda(H)) u = 0\) for some nonzero \(H \in \mathfrak{h}\) then \(u =
+  0\). In particular,
+  \[
+    \bigoplus_{\lambda \in \mathfrak{h}^*} \mathcal{U}(\mathfrak{g})_\lambda
+    = 0 \neq \mathcal{U}(\mathfrak{g})
+  \]
+\end{example}
+
+As a first consequence of Corollary~\ref{thm:finite-dim-is-weight-mod} we
+show\dots
+
+\begin{corollary}
+  The restriction of the Killing form \(\kappa\) to \(\mathfrak{h}\) is
+  non-degenerate.
+\end{corollary}
+
+\begin{proof}
+  Consider the root space decomposition \(\mathfrak{g} = \mathfrak{g}_0 \oplus
+  \bigoplus_\alpha \mathfrak{g}_\alpha\) of the adjoint
+  \(\mathfrak{g}\)-module, where \(\alpha\) ranges over all nonzero eigenvalues
+  of the adjoint action of \(\mathfrak{h}\). We claim \(\mathfrak{g}_0 =
+  \mathfrak{h}\).
+
+  Indeed, since \(\mathfrak{h}\) is Abelian, \(\operatorname{ad}(\mathfrak{h})
+  \mathfrak{h} = 0\) -- i.e. \(\mathfrak{h} \subset \mathfrak{g}_0\). On the
+  other hand, since \(\mathfrak{h}\) is self-normalizing, if \([X, H] = 0 \in
+  \mathfrak{h}\) for all \(H \in \mathfrak{h}\) then \(X \in \mathfrak{h}\) --
+  i.e. \(\mathfrak{g}_0 \subset \mathfrak{h}\). So the eigenspace decomposition
+  becomes
+  \[
+    \mathfrak{g} = \mathfrak{h} \oplus \bigoplus_\alpha \mathfrak{g}_\alpha
+  \]
+
+  We furthermore claim that \(\mathfrak{h} = \mathfrak{g}_0\) is orthogonal to
+  \(\mathfrak{g}_\alpha\) with respect to \(\kappa\) for any \(\alpha \ne 0\).
+  Indeed, given \(X \in \mathfrak{g}_\alpha\) and \(H_1, H_2 \in \mathfrak{h}\)
+  with \(\alpha(H_1) \ne 0\) we have
+  \[
+    \alpha(H_1) \cdot \kappa(X, H_2)
+    = \kappa([H_1, X], H_2)
+    = - \kappa([X, H_1], H_2)
+    = - \kappa(X, [H_1, H_2])
+    = 0
+  \]
+
+  Hence the non-degeneracy of \(\kappa\) implies the non-degeneracy of its
+  restriction.
+\end{proof}
+
+We should point out that the restriction of \(\kappa\) to \(\mathfrak{h}\) is
+\emph{not} the Killing form of \(\mathfrak{h}\). In fact, since
+\(\mathfrak{h}\) is Abelian, its Killing form is identically zero -- which is
+hardly ever a non-degenerate form.
+
+\begin{note}
+  Since \(\kappa\) induces an isomorphism \(\mathfrak{h} \isoto
+  \mathfrak{h}^*\), it induces a bilinear form \((\kappa(X, \cdot), \kappa(Y,
+  \cdot)) \mapsto \kappa(X, Y)\) in \(\mathfrak{h}^*\). As in
+  section~\ref{sec:sl3-reps}, we denote this form by \(\kappa\) as well.
+\end{note}
+
+We now have most of the necessary tools to reproduce the results of the
+previous chapter in a general setting. Let \(\mathfrak{g}\) be a
+finite-dimensional semisimple algebra with a Cartan subalgebra \(\mathfrak{h}\)
+and let \(M\) be a finite-dimensional simple \(\mathfrak{g}\)-module. We will
+proceed, as we did before, by generalizing the results of the previous two
+sections in order. By now the pattern should be starting to become clear, so we
+will mostly omit technical details and proofs analogous to the ones on the
+previous sections. Further details can be found in appendix D of
+\cite{fulton-harris} and in \cite{humphreys}.
+
+\section{The Geometry of Roots and Weights}
+
+We begin our analysis, as we did for \(\mathfrak{sl}_2(K)\) and
+\(\mathfrak{sl}_3(K)\), by investigating the locus of roots of and weights of
+\(\mathfrak{g}\). Throughout chapter~\ref{ch:sl3} we have seen that the weights
+of any given finite-dimensional module of \(\mathfrak{sl}_2(K)\) or
+\(\mathfrak{sl}_3(K)\) can only assume very rigid configurations. For instance,
+we have seen that the roots of \(\mathfrak{sl}_2(K)\) and
+\(\mathfrak{sl}_3(K)\) are symmetric with respect to the origin. In this
+chapter we will generalize most results from chapter~\ref{ch:sl3} regarding the
+rigidity of the geometry of the set of weights of a given module.
+
+As for the aforementioned result on the symmetry of roots, this turns out to be
+a general fact, which is a consequence of the non-degeneracy of the restriction
+of the Killing form to the Cartan subalgebra.
+
+\begin{proposition}\label{thm:weights-symmetric-span}
+  The roots \(\alpha\) of \(\mathfrak{g}\) are symmetrical about the origin --
+  i.e. \(- \alpha\) is also a root -- and they span all of \(\mathfrak{h}^*\).
+\end{proposition}
+
+\begin{proof}
+  We will start with the first claim. Let \(\alpha\) and \(\beta\) be two
+  roots. Notice \([\mathfrak{g}_\alpha, \mathfrak{g}_\beta] \subset
+  \mathfrak{g}_{\alpha + \beta}\). Indeed, if \(X \in \mathfrak{g}_\alpha\) and
+  \(Y \in \mathfrak{g}_\beta\) then
+  \[
+    [H [X, Y]]
+    = [X, [H, Y]] - [Y, [H, X]]
+    = (\alpha + \beta)(H) \cdot [X, Y]
+  \]
+  for all \(H \in \mathfrak{h}\).
+
+  This implies that if \(\alpha + \beta \ne 0\) then \(\operatorname{ad}(X)
+  \operatorname{ad}(Y)\) is nilpotent: if \(Z \in \mathfrak{g}_\gamma\) then
+  \[
+    (\operatorname{ad}(X) \operatorname{ad}(Y))^r Z
+    = [X, [Y, [ \ldots, [X, [Y, Z]]] \ldots ]
+    \in \mathfrak{g}_{r \alpha + r \beta + \gamma}
+    = 0
+  \]
+  for \(r\) large enough. In particular, \(\kappa(X, Y) =
+  \operatorname{Tr}(\operatorname{ad}(X) \operatorname{ad}(Y)) = 0\). Now if
+  \(- \alpha\) is not an eigenvalue we find \(\kappa(X, \mathfrak{g}_\beta) =
+  0\) for all roots \(\beta\), which contradicts the non-degeneracy of
+  \(\kappa\). Hence \(- \alpha\) must be an eigenvalue of the adjoint action of
+  \(\mathfrak{h}\).
+
+  For the second statement, note that if the roots of \(\mathfrak{g}\) do not
+  span all of \(\mathfrak{h}^*\) then there is some nonzero \(H \in
+  \mathfrak{h}\) such that \(\alpha(H) = 0\) for all roots \(\alpha\), which is
+  to say, \(\operatorname{ad}(H) X = [H, X] = 0\) for all \(X \in
+  \mathfrak{g}\). Another way of putting it is to say \(H\) is an element of
+  the center \(\mathfrak{z} = 0\) of \(\mathfrak{g}\), a contradiction.
+\end{proof}
+
+Furthermore, as in the case of \(\mathfrak{sl}_2(K)\) and
+\(\mathfrak{sl}_3(K)\) one can show\dots
+
+\begin{proposition}\label{thm:root-space-dim-1}
+  The root spaces \(\mathfrak{g}_\alpha\) are all \(1\)-dimensional.
+\end{proposition}
+
+The proof of the first statement of
+Proposition~\ref{thm:weights-symmetric-span} highlights something interesting:
+if we fix some eigenvalue \(\alpha\) of the adjoint action of \(\mathfrak{h}\)
+on \(\mathfrak{g}\) and a eigenvector \(X \in \mathfrak{g}_\alpha\), then for
+each \(H \in \mathfrak{h}\) and \(m \in M_\lambda\) we find
+\[
+  H \cdot (X \cdot m)
+  = X H \cdot m + [H, X] \cdot m
+  = (\lambda + \alpha)(H) X \cdot m
+\]
+
+Thus \(X\) sends \(m\) to \(M_{\lambda + \alpha}\). We have encountered this
+formula twice in these notes: again, we find \(\mathfrak{g}_\alpha\) \emph{acts
+on \(M\) by translating vectors between eigenspaces}. In particular, if we
+denote by \(\Delta\) the set of all roots of \(\mathfrak{g}\) then\dots
+
+\begin{theorem}\label{thm:weights-congruent-mod-root}\index{weights!root lattice}
+  The weights of a finite-dimensional simple \(\mathfrak{g}\)-module \(M\) are
+  all congruent modulo the root lattice \(Q = \mathbb{Z} \Delta\) of
+  \(\mathfrak{g}\). In other words, all weights of \(M\) lie in the same
+  \(Q\)-coset \(\xi \in \mfrac{\mathfrak{h}^*}{Q}\).
+\end{theorem}
+
+Again, we may leverage our knowledge of \(\mathfrak{sl}_2(K)\) to obtain
+further restrictions on the geometry of the locus of weights of \(M\). Namely,
+as in the case of \(\mathfrak{sl}_3(K)\) we show\dots
+
+\begin{proposition}\label{thm:distinguished-subalgebra}
+  Given a root \(\alpha\) of \(\mathfrak{g}\) the subspace
+  \(\mathfrak{s}_\alpha = \mathfrak{g}_\alpha \oplus \mathfrak{g}_{- \alpha}
+  \oplus [\mathfrak{g}_\alpha, \mathfrak{g}_{- \alpha}]\) is a subalgebra
+  isomorphic to \(\mathfrak{sl}_2(K)\).
+\end{proposition}
+
+\begin{corollary}\label{thm:distinguished-subalg-rep}
+  For all weights \(\mu\), the subspace
+  \[
+    \bigoplus_k M_{\mu - k \alpha}
+  \]
+  is invariant under the action of the subalgebra \(\mathfrak{s}_\alpha\)
+  and the weight spaces in this string match the eigenspaces of \(h\).
+\end{corollary}
+
+The proof of Proposition~\ref{thm:distinguished-subalgebra} is very technical
+in nature and we won't include it here, but the idea behind it is simple:
+recall that \(\mathfrak{g}_\alpha\) and \(\mathfrak{g}_{- \alpha}\) are both
+\(1\)-dimensional, so that \(\dim [\mathfrak{g}_\alpha, \mathfrak{g}_{-
+\alpha}]\) is at most 1. We check that \([\mathfrak{g}_\alpha, \mathfrak{g}_{-
+\alpha}] \ne 0\) and that no generator of \([\mathfrak{g}_\alpha,
+\mathfrak{g}_{- \alpha}]\) is annihilated by \(\alpha\), so that by adjusting
+scalars we can find \(E_\alpha \in \mathfrak{g}_\alpha\) and \(F_\alpha \in
+\mathfrak{g}_{- \alpha}\) such that \(H_\alpha = [E_\alpha, F_\alpha]\)
+satisfies
+\begin{align*}
+  [H_\alpha, F_\alpha] & = -2 F_\alpha &
+  [H_\alpha, E_\alpha] & =  2 E_\alpha
+\end{align*}
+
+The elements \(E_\alpha, F_\alpha \in \mathfrak{g}\) are not uniquely
+determined by this condition, but \(H_\alpha\) is. As promised, the second
+statement of Corollary~\ref{thm:distinguished-subalg-rep} imposes strong
+restrictions on the weights of \(M\). Namely, if \(\lambda\) is a weight,
+\(\lambda(H_\alpha)\) is an eigenvalue of \(h\) on some
+\(\mathfrak{sl}_2(K)\)-module, so it must be an integer. In other words\dots
+
+\begin{definition}\label{def:weight-lattice}\index{weights!weight lattice}
+  The lattice \(P = \{ \lambda \in \mathfrak{h}^* : \lambda(H_\alpha) \in
+  \mathbb{Z} \, \forall \alpha \in \Delta \} \subset \mathfrak{h}^*\) is called
+  \emph{the weight lattice of \(\mathfrak{g}\)}. We call the elements of \(P\)
+  \emph{integral}.
+\end{definition}
+
+\begin{proposition}\label{thm:weights-fit-in-weight-lattice}
+  The weights of a finite-dimensional simple \(\mathfrak{g}\)-module \(M\) of
+  all lie in the weight lattice \(P\).
+\end{proposition}
+
+Proposition~\ref{thm:weights-fit-in-weight-lattice} is clearly analogous to
+Corollary~\ref{thm:sl3-weights-fit-in-weight-lattice}. In fact, the weight
+lattice of \(\mathfrak{sl}_3(K)\) -- as in Definition~\ref{def:weight-lattice}
+-- is precisely \(\mathbb{Z} \langle \alpha_1, \alpha_2, \alpha_3 \rangle\). To
+proceed further, we would like to take \emph{the highest weight of \(M\)} as in
+section~\ref{sec:sl3-reps}, but the meaning of \emph{highest} is again unclear
+in this situation. We could simply fix a linear function \(\mathbb{Q} P \to
+\mathbb{Q}\) -- as we did in section~\ref{sec:sl3-reps} -- and choose a weight
+\(\lambda\) of \(M\) that maximizes this functional, but at this point it is
+convenient to introduce some additional tools to our arsenal. These tools are
+called \emph{basis}.
+
+\begin{definition}\label{def:basis-of-root}\index{weights!basis}
+  A subset \(\Sigma = \{\beta_1, \ldots, \beta_r\} \subset \Delta\) of linearly
+  independent roots is called \emph{a basis for \(\Delta\)} if, given \(\alpha
+  \in \Delta\), there are \(k_1, \ldots, k_r \in \mathbb{N}\) such that
+  \(\alpha = \pm(k_1 \beta_1 + \cdots + k_r \beta_r)\).
+\end{definition}
+
+The interesting thing about basis for \(\Delta\) is that they allow us to
+compare weights of a given \(\mathfrak{g}\)-module. At this point the reader
+should be asking himself: how? Definition~\ref{def:basis-of-root} isn't exactly
+all that intuitive. Well, the thing is that any choice of basis induces a
+partial order in \(Q\), where elements are ordered by their \emph{heights}.
+
+\begin{definition}\index{weights!orderings of roots}
+  Let \(\Sigma = \{\beta_1, \ldots, \beta_r\}\) be a basis for \(\Delta\).
+  Given \(\alpha = k_1 \beta_1 + \cdots + k_r \beta_r \in Q\) with \(k_1,
+  \ldots, k_r \in \mathbb{Z}\), we call the number \(\operatorname{ht}(\alpha)
+  = k_1 + \cdots + k_r \in \mathbb{Z}\) \emph{the height of \(\alpha\)}. We say
+  that \(\alpha \preceq \beta\) if \(\operatorname{ht}(\alpha) \le
+  \operatorname{ht}(\beta)\).
+\end{definition}
+
+\begin{definition}
+  Given a basis \(\Sigma\) for \(\Delta\), there is a canonical
+  partition\footnote{Notice that $\operatorname{ht}(\alpha) = 0$ if, and only
+  if $\alpha = 0$. Since $0$ is, by definition, not a root, the sets $\Delta^+$
+  and $\Delta^-$ account for all roots.} \(\Delta^+ \cup \Delta^- = \Delta\),
+  where \(\Delta^+ = \{ \alpha \in \Delta : \alpha \succ 0 \}\) and \(\Delta^-
+  = \{ \alpha \in \Delta : \alpha \prec 0 \}\). The elements of \(\Delta^+\)
+  and \(\Delta^-\) are called \emph{positive} and \emph{negative roots},
+  respectively.
+\end{definition}
+
+\begin{definition}
+  Let \(\Sigma\) be a basis for \(\Delta\). The subalgebra \(\mathfrak{b} =
+  \mathfrak{h} \oplus \bigoplus_{\alpha \in \Delta^+} \mathfrak{g}_\alpha\) is
+  called \emph{the Borel subalgebra associated with \(\mathfrak{h}\) and
+  \(\Sigma\)}.
+\end{definition}
+
+It should be obvious that the binary relation \(\preceq\) in \(Q\) is a partial
+order. In addition, we may compare the elements of a given \(Q\)-coset
+\(\lambda + Q\) by comparing their difference with \(0 \in Q\). In other words,
+given \(\lambda \in \mu + Q\), we say \(\lambda \preceq \mu\) if \(\lambda -
+\mu \preceq 0\). In particular, since the weights of \(M\) all lie in a single
+\(Q\)-coset, we may compare them in this way. Given a basis \(\Sigma\) for
+\(\Delta\) we may take ``the highest weight of \(M\)'' as a maximal weight
+\(\lambda\) of \(M\). The obvious question then is: can we always find a basis
+for \(\Delta\)?
+
+\begin{proposition}
+  There is a basis \(\Sigma\) for \(\Delta\).
+\end{proposition}
+
+The intuition behind the proof of this proposition is similar to our original
+idea of fixing a direction in \(\mathfrak{h}^*\) in the case of
+\(\mathfrak{sl}_3(K)\). Namely, one can show that \(\kappa(\alpha, \beta) \in
+\mathbb{Z}\) for all \(\alpha, \beta \in \Delta\), so that the Killing form
+\(\kappa\) restricts to a nondegenerate \(\mathbb{Q}\)-bilinear form
+\(\mathbb{Q} \Delta \times \mathbb{Q} \Delta \to \mathbb{Q}\). We can then fix
+a nonzero vector \(\gamma \in \mathbb{Q} \Delta\) and consider the orthogonal
+projection \(f : \mathbb{Q} \Delta \to \mathbb{Q} \gamma \cong \mathbb{Q}\). We
+say a root \(\alpha \in \Delta\) is \emph{positive} if \(f(\alpha) > 0\), and
+we call a positive root \(\alpha\) \emph{simple} if it cannot be written as the
+sum two other positive roots. The subset \(\Sigma \subset \Delta\) of all
+simple roots is a basis for \(\Delta\), and all other basis can be shown to
+arise in this way.
+
+Fix some basis \(\Sigma\) for \(\Delta\), with corresponding decomposition
+\(\Delta^+ \cup \Delta^- = \Delta\). Let \(\lambda\) be a maximal weight of
+\(M\). We call \(\lambda\) \emph{the highest weight of \(M\)}, and we call any
+nonzero \(m \in M_\lambda\) \emph{a highest weight vector}. The strategy then
+is to describe all weight spaces of \(M\) in terms of \(\lambda\) and \(m\), as
+in Theorem~\ref{thm:sl3-irr-weights-class}. Unsurprisingly we do so by
+reproducing the proof of the case of \(\mathfrak{sl}_3(K)\).
+
+First, we note that any highest weight vector \(m \in M_\lambda\) is
+annihilated by all positive root spaces, for if \(\alpha \in \Delta^+\) then
+\(E_\alpha \cdot m \in M_{\lambda + \alpha}\) must be zero -- or otherwise we
+would have that \(\lambda + \alpha\) is a weight with \(\lambda \prec \lambda +
+\alpha\). In particular,
+\[
+  \bigoplus_{k \in \mathbb{Z}}   M_{\lambda - k \alpha}
+  = \bigoplus_{k \in \mathbb{N}} M_{\lambda - k \alpha}
+\]
+and \(\lambda(H_\alpha)\) is the right-most eigenvalue of the action of \(h\)
+on the \(\mathfrak{sl}_2(K)\)-module \(\bigoplus_k M_{\lambda - k \alpha}\).
+
+This has a number of important consequences. For instance\dots
+
+\begin{corollary}
+  If \(\alpha \in \Delta^+\) and \(\sigma_\alpha : \mathfrak{h}^* \to
+  \mathfrak{h}^*\) is the reflection in the hyperplane perpendicular to
+  \(\alpha\) with respect to the Killing form, the weights of \(M\) occurring
+  in the line joining \(\lambda\) and \(\sigma_\alpha\) are precisely the \(\mu
+  \in P\) lying between \(\lambda\) and \(\sigma_\alpha(\lambda)\).
+\end{corollary}
+
+\begin{proof}
+  Notice that any \(\mu \in P\) in the line joining \(\lambda\) and
+  \(\sigma_\alpha(\lambda)\) has the form \(\mu = \lambda - k \alpha\) for some
+  \(k\), so that \(M_\mu\) corresponds the eigenspace associated with the
+  eigenvalue \(\lambda(H_\alpha) - 2k\) of the action of \(h\) on \(\bigoplus_k
+  M_{\lambda - k \alpha}\). If \(\mu\) lies between \(\lambda\) and
+  \(\sigma_\alpha(\lambda)\) then \(k\) lies between \(0\) and
+  \(\lambda(H_\alpha)\), in which case \(M_\mu \neq 0\) and therefore \(\mu\)
+  is a weight.
+
+  On the other hand, if \(\mu\) does not lie between \(\lambda\) and
+  \(\sigma_\alpha(\lambda)\) then either \(k < 0\) or \(k >
+  \lambda(H_\alpha)\). Suppose \(\mu\) is a weight. In the first case \(\mu
+  \succ \lambda\), a contradiction. On the second case the fact that \(M_\mu
+  \ne 0\) implies \(M_{\lambda  + (k - \lambda(H_\alpha)) \alpha} =
+  M_{\sigma_\alpha(\mu)} \ne 0\), which contradicts the fact that \(M_{\lambda
+  + \ell \alpha} = 0\) for all \(\ell \ge 0\).
+\end{proof}
+
+This is entirely analogous to the situation of \(\mathfrak{sl}_3(K)\), where we
+found that the weights of the simple \(\mathfrak{sl}_3(K)\)-modules formed
+continuous strings symmetric with respect to the lines \(K \alpha\) with
+\(\kappa(\alpha_i - \alpha_j, \alpha) = 0\). As in the case of
+\(\mathfrak{sl}_3(K)\), the same class of arguments leads us to the
+conclusion\dots
+
+\begin{definition}\index{Weyl group}
+  We refer to the group \(W = \langle \sigma_\alpha : \alpha \in
+  \Delta^+ \rangle \subset \operatorname{O}(\mathfrak{h}^*)\) as \emph{the Weyl
+  group of \(\mathfrak{g}\)}.
+\end{definition}
+
+\begin{theorem}\label{thm:irr-weight-class}
+  The weights of a simple \(\mathfrak{g}\)-module \(M\) with highest weight
+  \(\lambda\) are precisely the elements of the weight lattice \(P\) congruent
+  to \(\lambda\) modulo the root lattice \(Q\) lying inside the convex hull of
+  the orbit of \(\lambda\) under the action of the Weyl group \(W\).
+\end{theorem}
+
+\index{Weyl group!actions}
+Aside from showing up in the previous theorem, the Weyl group will also play an
+important role in chapter~\ref{ch:mathieu} by virtue of the existence of a
+canonical action of \(W\) on \(\mathfrak{h}\). By its very nature,
+\(W\) acts in \(\mathfrak{h}^*\). If we conjugate the action
+\(\sigma\!\restriction_{\mathfrak{h}^*} : \mathfrak{h}^* \isoto
+\mathfrak{h}^*\) of some \(\sigma \in W\) by the isomorphism
+\(\mathfrak{h}^* \isoto \mathfrak{h}\) afforded by the restriction of the
+Killing for to \(\mathfrak{h}\) we get a linear automorphism
+\(\sigma\!\restriction_{\mathfrak{h}} : \mathfrak{h} \isoto \mathfrak{h}\). As
+it turns out, the \(\sigma\!\restriction_{\mathfrak{h}}\) can be extended to an
+automorphism of Lie algebras \(\mathfrak{g} \isoto \mathfrak{g}\). This
+translates into the following results, which we do not prove -- but see
+\cite[sec.~14.3]{humphreys}.
+
+\begin{proposition}\label{thm:weyl-group-action}
+  Given \(\alpha \in \Delta^+\), let\footnote{Notice that since $\mathfrak{g}$
+  is finite-dimensional, $\operatorname{ad}(X)$ is nilpotent for each root
+  vector $X \in \mathfrak{g}$, so that the linear automorphism
+  $e^{\operatorname{ad}(X)} = \operatorname{Id} + \operatorname{ad}(X) +
+  \frac{\operatorname{ad}(X)^2}{2!} + \cdots : \mathfrak{g} \isoto
+  \mathfrak{g}$ is well defined.} \(\tilde{\sigma}_\alpha =
+  e^{\operatorname{ad}(E_\alpha)} e^{- \operatorname{ad}(F_\alpha)}
+  e^{\operatorname{ad}(E_\alpha)} : \mathfrak{g} \isoto \mathfrak{g}\). Then
+  \(\tilde\sigma_\alpha\) is an automorphism of Lie algebras, and this defines
+  an action of \(W\) on \(\mathfrak{g}\) which is compatible with the
+  canonical action of \(W\) on \(\mathfrak{h}\) -- i.e.
+  \(\tilde\sigma\!\restriction_{\mathfrak{h}} =
+  \sigma\!\restriction_{\mathfrak{h}}\) for all \(\sigma \in W\).
+\end{proposition}
+
+\begin{note}
+  Notice that the action of \(W\) on \(\mathfrak{g}\) from
+  Proposition~\ref{thm:weyl-group-action} is not canonical, since it depends on
+  the choice of \(E_\alpha\) and \(F_\alpha\). Nevertheless, \(\mathfrak{h}\)
+  is stable under the action of \(W\) -- i.e. \(W \cdot
+  \mathfrak{h} \subset \mathfrak{h}\) -- and the restriction of this action to
+  \(\mathfrak{h}\) is independent of any choices.
+\end{note}
+
+We should point out that the results in this section regarding the geometry
+roots and weights are only the beginning of a well develop axiomatic theory of
+the so called \emph{root systems}, which was used by Cartan in the early 20th
+century to classify all finite-dimensional simple complex Lie algebras in terms
+of Dynking diagrams. This and much more can be found in \cite[III]{humphreys}
+and \cite[ch.~21]{fulton-harris}. Having found all of the weights of \(M\), the
+only thing we are missing for a complete classification is an existence and
+uniqueness theorem analogous to Theorem~\ref{thm:sl2-exist-unique} and
+Theorem~\ref{thm:sl3-existence-uniqueness}. This will be the focus of the next
+section.
+
+\section{Verma Modules \& the Highest Weight Theorem}
+
+It is already clear from the previous discussion that if \(\lambda\) is the
+highest weight of \(M\) then \(\lambda(H_\alpha) \ge 0\) for all positive roots
+\(\alpha\). In other words, having \(\lambda(H_\alpha) \ge 0\), for all
+\(\alpha \in \Delta^+\), is a necessary condition for the existence of a simple
+\(\mathfrak{g}\)-module with highest weight given by \(\lambda\). Surprisingly,
+this condition is also sufficient. In other words\dots
+
+\begin{definition}\index{weights!dominant weight}\index{weights!integral weight}
+  An element \(\lambda\) of \(P\) such that \(\lambda(H_\alpha) \ge 0\) for all
+  \(\alpha \in \Delta^+\) is referred to as an \emph{dominant integral weight
+  of \(\mathfrak{g}\)}.
+\end{definition}
+
+\begin{theorem}\label{thm:dominant-weight-theo}
+  For each dominant integral \(\lambda \in P\) there exists precisely one
+  finite-dimensional simple \(\mathfrak{g}\)-module \(M\) whose highest weight
+  is \(\lambda\).
+\end{theorem}
+
+\index{weights!Highest Weight Theorem} This is known as \emph{the Highest
+Weight Theorem}, and its proof is the focus of this section. The ``uniqueness''
+part of the theorem follows at once from the arguments used for
+\(\mathfrak{sl}_3(K)\). However, the ``existence'' part of the theorem is more
+nuanced.
+
+Our first instinct is, of course, to try to generalize the proof used for
+\(\mathfrak{sl}_3(K)\). The issue is that our proof relied heavily on our
+knowledge of the roots of \(\mathfrak{sl}_3(K)\). It is thus clear that we need
+a more systematic approach for the general setting. We begin by asking a
+simpler question: how can we construct \emph{any} -- potentially
+infinite-dimensional -- \(\mathfrak{g}\)-module \(M\) of highest weight
+\(\lambda\)? In the process of answering this question we will come across a
+surprisingly elegant solution to our problem.
+
+If \(M\) is a module with highest weight vector \(m^+ \in M_\lambda\), we
+already know \(H \cdot m^+ = \lambda(H) m^+\) for all \(\mathfrak{h}\) and \(X
+\cdot m^+ = 0\) for \(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\). If
+\(M\) is simple we find \(M = \mathcal{U}(\mathfrak{g}) \cdot m^+\), which
+implies the restriction of \(M\) to the Borel subalgebra \(\mathfrak{b} \subset
+\mathfrak{g}\) has a prescribed action. On the other hand, we have essentially
+no information about the action of the rest of \(\mathfrak{g}\) on \(M\).
+Nevertheless, given a \(\mathfrak{b}\)-module we may obtain a
+\(\mathfrak{g}\)-module by formally extending the action of \(\mathfrak{b}\)
+via induction. This leads us to the following definition.
+
+\begin{definition}\label{def:verma}\index{\(\mathfrak{g}\)-module!(generalized) Verma modules}
+  The \(\mathfrak{g}\)-module \(M(\lambda) =
+  \operatorname{Ind}_{\mathfrak{b}}^{\mathfrak{g}} K m^+\), where the action of
+  \(\mathfrak{b}\) on \(K m^+\) is given by \(H \cdot m^+ = \lambda(H) m^+\)
+  for all \(H \in \mathfrak{h}\) and \(X \cdot m^+ = 0\) for \(X \in
+  \mathfrak{g}_{\alpha}\), \(\alpha \in \Delta^+\), is called \emph{the Verma
+  module of weight \(\lambda\)}.
+\end{definition}
+
+It turns out that \(M(\lambda)\) enjoys many of the features we've grown used
+to in the past chapters. Explicitly\dots
+
+\begin{proposition}\label{thm:verma-is-weight-mod}
+  The Verma module \(M(\lambda)\) is generated \(m^+ = 1 \otimes m^+ \in
+  M(\lambda)\) as in Definition~\ref{def:verma}. The weight spaces
+  decomposition
+  \[
+    M(\lambda) = \bigoplus_{\mu \in \mathfrak{h}^*} M(\lambda)_\mu
+  \]
+  holds. Furthermore, \(\dim M(\lambda)_\mu < \infty\) for all \(\mu \in
+  \mathfrak{h}^*\) and \(\dim M(\lambda)_\lambda = 1\). Finally, \(\lambda\) is
+  the highest weight of \(M(\lambda)\), with highest weight vector given by
+  \(m^+ \in M(\lambda)\).
+\end{proposition}
+
+\begin{proof}
+  The PBW Theorem implies that \(M(\lambda)\) is spanned by the vectors
+  \(F_{\alpha_{i_1}} F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+\) for
+  \(\Delta^+ = \{\alpha_1, \ldots, \alpha_r\}\) and \(F_{\alpha_i} \in
+  \mathfrak{g}_{- \alpha_i}\) as in the proof of
+  Proposition~\ref{thm:distinguished-subalgebra}. But
+  \[
+    \begin{split}
+      H \cdot (F_{\alpha_{i_1}} F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+)
+      & = ([H, F_{\alpha_{i_1}}] + F_{\alpha_{i_1}} H)
+          F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+ \\
+      & = - \alpha_{i_1}(H) F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+
+        + F_{\alpha_{i_1}} ([H, F_{\alpha_{i_2}}] + F_{\alpha_{i_2}} H)
+          F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+ \\
+      & \;\; \vdots \\
+      & = (- \alpha_{i_1} - \cdots - \alpha_{i_s})(H)
+          F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+
+        + F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} H \cdot m^+ \\
+      & = (\lambda - \alpha_{i_1} - \cdots - \alpha_{i_s})(H)
+          F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+ \\
+      & \therefore F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+
+        \in M(\lambda)_{\lambda - \alpha_{i_1} - \cdots - \alpha_{i_s}}
+    \end{split}
+  \]
+
+  Hence \(M(\lambda) \subset \bigoplus_{\mu \in \mathfrak{h}^*}
+  M(\lambda)_\mu\), as desired. In fact we have established
+  \[
+    M(\lambda)
+    \subset
+    \bigoplus_{k_i \in \mathbb{N}}
+    M(\lambda)_{\lambda - k_1 \cdot \alpha_1 - \cdots - k_r \cdot \alpha_r}
+  \]
+  where \(\{\alpha_1, \ldots, \alpha_r\} = \Delta^+\), so that all weights of
+  \(M(\lambda)\) have the form \(\mu = \lambda - k_1 \cdot \alpha_1 - \cdots -
+  k_r \cdot \alpha_r\).
+
+  This already gives us that the weights of \(M(\lambda)\) are bounded by
+  \(\lambda\). To see that \(\lambda\) is indeed a weight, we show that
+  \(m^+\) is nonzero weight vector. Clearly \(m^+ \in M_\lambda\). The
+  Poincaré-Birkhoff-Witt Theorem implies \(\mathcal{U}(\mathfrak{g})\) is a
+  free \(\mathfrak{b}\)-module, so that
+  \[
+    M(\lambda)
+    \cong \left(\bigoplus_i \mathcal{U}(\mathfrak{b}) \right)
+    \otimes_{\mathcal{U}(\mathfrak{b})} K m^+
+    \cong \bigoplus_i \mathcal{U}(\mathfrak{b})
+    \otimes_{\mathcal{U}(\mathfrak{b})} K m^+
+    \cong \bigoplus_i K m^+
+    \ne 0
+  \]
+  as \(\mathfrak{b}\)-modules. We then conclude \(m^+ \ne 0\) in
+  \(M(\lambda)\), for if this was not the case we would find \(M(\lambda) =
+  \mathcal{U}(\mathfrak{g}) \cdot m^+ = 0\). Hence \(M_\lambda \ne 0\) and
+  therefore \(\lambda\) is the highest weight of \(M(\lambda)\), with highest
+  weight vector \(m^+\).
+
+  To see that \(\dim M(\lambda)_\mu < \infty\), simply note that there are only
+  finitely many monomials \(F_{\alpha_1}^{k_1} F_{\alpha_2}^{k_2} \cdots
+  F_{\alpha_s}^{k_s}\) such that \(\mu = \lambda + k_1 \cdot \alpha_1 + \cdots
+  + k_s \cdot \alpha_s\). Since \(M(\lambda)_\mu\) is spanned by the images of
+  \(m^+\) under such monomials, we conclude \(\dim M(\lambda) < \infty\). In
+  particular, there is a single monomials \(F_{\alpha_1}^{k_1}
+  F_{\alpha_2}^{k_2} \cdots F_{\alpha_s}^{k_s}\) such that \(\lambda = \lambda
+  + k_1 \cdot \alpha_1 + \cdots + k_s \cdot \alpha_s\) -- which is, of course,
+  the monomial where \(k_1 = \cdots = k_n = 0\). Hence \(\dim M_\lambda = 1\).
+\end{proof}
+
+\begin{example}\label{ex:sl2-verma}
+  If \(\mathfrak{g} = \mathfrak{sl}_2(K)\), then we can take \(\mathfrak{h} = K
+  h\) and \(\mathfrak{b} = K e \oplus K h\). If \(\lambda \in
+  \mathfrak{h}^*\) is the map \(h \mapsto 2\) then \(M(\lambda) =
+  \bigoplus_{k \ge 0} K f^k \cdot m^+\), and the action of
+  \(\mathfrak{sl}_2(K)\) on \(M(\lambda)\) is given by the formulas in
+  (\ref{eq:sl2-verma-formulas}). Visually,
+  \begin{center}
+    \begin{tikzcd}
+      \cdots            \rar[bend left=60]{-10}
+      & M(\lambda)_{-6} \rar[bend left=60]{-4} \lar[bend left=60]{1}
+      & M(\lambda)_{-4} \rar[bend left=60]{0}  \lar[bend left=60]{1}
+      & M(\lambda)_{-2} \rar[bend left=60]{2}  \lar[bend left=60]{1}
+      & M(\lambda)_0    \rar[bend left=60]{2}  \lar[bend left=60]{1}
+      & M(\lambda)_2                           \lar[bend left=60]{1}
+    \end{tikzcd}
+  \end{center}
+  where \(M(\lambda)_{2 - 2 k} = K f^k \cdot m^+\). Here the top arrows
+  represent the action of \(e\) and the bottom arrows represent the action of
+  \(f\). The scalars labeling each arrow indicate to which multiple of \(f^{k
+  \pm 1} \cdot m^+\) the elements \(e\) and \(f\) send \(f^k \cdot m^+\). The
+  string of weight spaces to the left of the diagram is infinite.
+  \begin{equation}\label{eq:sl2-verma-formulas}
+    \begin{aligned}
+      f^k \cdot m^+ & \overset{e}{\mapsto} (2 - k(k + 1)) f^{k - 1} \cdot m^+ &
+      f^k \cdot m^+ & \overset{f}{\mapsto} f^{k + 1} \cdot m^+                &
+      f^k \cdot m^+ & \overset{h}{\mapsto} (2 - 2k) f^k \cdot m^+             &
+    \end{aligned}
+  \end{equation}
+\end{example}
+
+The Verma module \(M(\lambda)\) should really be though-of as ``the freest
+highest weight \(\mathfrak{g}\)-module of weight \(\lambda\)''. Unfortunately
+for us, this is not a proof of Theorem~\ref{thm:dominant-weight-theo}, since in
+general \(M(\lambda)\) is neither simple nor finite-dimensional. Indeed, the
+dimension of \(M(\lambda)\) is the same as the codimension of
+\(\mathcal{U}(\mathfrak{b})\) in \(\mathcal{U}(\mathfrak{g})\), which is always
+infinite. Nevertheless, we may use \(M(\lambda)\) to prove
+Theorem~\ref{thm:dominant-weight-theo} as follows.
+
+Given a \(\mathfrak{g}\)-module \(M\), any \(\mathfrak{g}\)-homomorphism \(f :
+M(\lambda) \to M\) is determined by the image of \(m^+\). Indeed, \(f(u \cdot
+m^+) = u \cdot f(m^+)\) for all \(u \in \mathcal{U}(\mathfrak{g})\). In
+addition, it is clear that
+\[
+  H \cdot f(m^+) = f(H \cdot m^+) = f(\lambda(H) m^+) = \lambda(H) f(m^+)
+\]
+for all \(H \in \mathfrak{h}\) and, similarly, \(X \cdot f(m^+) = 0\) for all
+\(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\). This leads us to the
+universal property of \(M(\lambda)\).
+
+\begin{definition}
+  Let \(M\) be a \(\mathfrak{g}\)-module and \(m \in M\). If \(X \cdot m = 0\)
+  for all \(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\), then \(m\) is
+  called \emph{a singular vector of \(M\)}.
+\end{definition}
+
+\begin{proposition}
+  Let \(M\) be a \(\mathfrak{g}\)-module and \(m \in M_\lambda\) be a singular
+  vector. Then there exists a unique \(\mathfrak{g}\)-homomorphism \(f :
+  M(\lambda) \to M\) such that \(f(m^+) = m\). Furthermore, all homomorphisms
+  \(M(\lambda) \to M\) are given in this fashion.
+  \[
+    \operatorname{Hom}_{\mathfrak{g}}(M(\lambda), M)
+    \cong \{ m \in M_\lambda : m \ \text{is singular}\}
+  \]
+\end{proposition}
+
+\begin{proof}
+  The result follows directly from Proposition~\ref{thm:frobenius-reciprocity}.
+  Indeed, by the Frobenius Reciprocity Theorem, a \(\mathfrak{g}\)-homomorphism
+  \(f : M(\lambda) \to M\) is the same as a \(\mathfrak{b}\)-homomorphism \(g :
+  K m^+ \to M = \operatorname{Res}_{\mathfrak{b}}^{\mathfrak{g}} M\). More
+  specifically, given a \(\mathfrak{b}\)-homomorphism \(g : K m^+ \to M\),
+  there exists a unique \(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\)
+  such that \(f(u \otimes m) = u \cdot g(m)\) for all \(m \in K m^+\), and all
+  \(\mathfrak{g}\)-homomorphism \(M(\lambda) \to M\) arise in this fashion.
+
+  Any \(K\)-linear map \(g : K m^+ \to M\) is determined by the image
+  \(g(m^+)\) of \(m^+\) and such an image is a singular vector if, and only if
+  \(g\) is a \(\mathfrak{b}\)-homomorphism.
+\end{proof}
+
+Notice that any highest weight vector is a singular vector. Now suppose \(M\)
+is a simple finite-dimensional \(\mathfrak{g}\)-module of highest weight vector
+\(m \in M_\lambda\). By the last proposition, there is a
+\(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\) such that \(f(m^+) = m\).
+Since \(M\) is simple, \(M = \mathcal{U}(\mathfrak{g}) \cdot m\) and therefore
+\(M \cong \mfrac{M(\lambda)}{\ker f}\). It then follows from the simplicity of
+\(M\) that \(\ker f \subset M(\lambda)\) is a maximal
+\(\mathfrak{g}\)-submodule. Maximal submodules of Verma modules are thus of
+primary interest to us. As it turns out, these can be easily classified.
+
+\begin{proposition}\label{thm:max-verma-submod-is-weight}
+  Every submodule \(N \subset M(\lambda)\) is the direct sum of its weight
+  spaces. In particular, \(M(\lambda)\) has a unique maximal submodule
+  \(N(\lambda)\) and a unique simple quotient \(L(\lambda) =
+  \sfrac{M(\lambda)}{N(\lambda)}\).
+\end{proposition}
+
+\begin{proof}
+  Let \(N \subset M(\lambda)\) be a submodule and take any nonzero \(n \in N\).
+  Because of Proposition~\ref{thm:verma-is-weight-mod}, we know there are
+  \(\mu_1, \ldots, \mu_r \in \mathfrak{h}^*\) and nonzero \(m_i \in
+  M(\lambda)_{\mu_i}\) such that \(n = m_1 + \cdots + m_r\). We want to show
+  \(m_i \in N\) for all \(i\).
+
+  Fix some \(H_2 \in \mathfrak{h}\) such that \(\mu_1(H_2) \ne \mu_2(H_2)\).
+  Then
+  \[
+    m_1
+    - \frac{(\mu_3 - \mu_1)(H_2)}{(\mu_2 - \mu_1)(H_2)} \cdot m_3
+    - \cdots
+    - \frac{(\mu_r - \mu_1)(H_2)}{(\mu_2 - \mu_1)(H_2)} \cdot m_r
+    = \left( 1 - \frac{H_2 - \mu_1(H_2)}{(\mu_2 - \mu_1)(H_2)} \right) \cdot n
+    \in N
+  \]
+
+  Now take \(H_3 \in \mathfrak{h}\) such that \(\mu_1(H_3) \ne \mu_3(H_3)\). By
+  applying the same procedure again we get
+  \begin{multline*}
+    m_1
+    -
+    \frac{(\mu_4 - \mu_3)(H_3) \cdot (\mu_4 - \mu_1)(H_2)}
+         {(\mu_3 - \mu_1)(H_3) \cdot (\mu_2 - \mu_1)(H_2)} \cdot m_4
+    - \cdots -
+    \frac{(\mu_r - \mu_3)(H_3) \cdot (\mu_r - \mu_1)(H_2)}
+         {(\mu_3 - \mu_1)(H_3) \cdot (\mu_2 - \mu_1)(H_2)} \cdot m_r \\
+    =
+    \left(1 - \frac{H_3 - \mu_1(H_3)}{(\mu_3 - \mu_1)(H_3)} \right)
+    \left(1 - \frac{H_2 - \mu_1(H_2)}{(\mu_2 - \mu_1)(H_2)} \right) \cdot n
+    \in N
+  \end{multline*}
+
+  By applying the same procedure over and over again we can see that \(m_1 = u
+  \cdot n \in N\) for some \(u \in \mathcal{U}(\mathfrak{g})\). Furthermore, if
+  we reproduce all this for \(m_2 + \cdots + m_r = n - m_1 \in N\) we get that
+  \(m_2 \in N\). All in all we find \(m_1, \ldots, m_r \in N\). Hence
+  \[
+    N = \bigoplus_\mu N_\mu = \bigoplus_\mu M(\lambda)_\mu \cap N
+  \]
+
+  Since \(M(\lambda) = \mathcal{U}(\mathfrak{g}) \cdot m^+\), if \(N\) is a
+  proper submodule then \(m^+ \notin N\). Hence any proper submodule lies in
+  the sum of weight spaces other than \(M(\lambda)_\lambda\), so the sum
+  \(N(\lambda)\) of all such submodules is still proper. This implies
+  \(N(\lambda)\) is the unique maximal submodule of \(M(\lambda)\) and
+  \(L(\lambda) = \sfrac{M(\lambda)}{N(\lambda)}\) is its unique simple
+  quotient.
+\end{proof}
+
+\begin{example}\label{ex:sl2-verma-quotient}
+  If \(\mathfrak{g} = \mathfrak{sl}_2(K)\) and \(\lambda : h \mapsto 2\), we
+  can see from Example~\ref{ex:sl2-verma} that \(N(\lambda) = \bigoplus_{k \ge
+  3} K f^k \cdot m^+\), so that \(L(\lambda)\) is the \(3\)-dimensional simple
+  \(\mathfrak{sl}_2(K)\)-module -- i.e. the finite-dimensional simple module
+  with highest weight \(\lambda\) constructed in chapter~\ref{ch:sl3}.
+\end{example}
+
+All its left to prove the Highest Weight Theorem is verifying that the
+situation encountered in Example~\ref{ex:sl2-verma-quotient} holds for any
+\(\lambda \in P\). In other words, we need to show\dots
+
+\begin{proposition}\label{thm:verma-is-finite-dim}
+  If \(\mathfrak{g}\) is semisimple and \(\lambda\) is dominant integral then
+  the unique simple quotient \(L(\lambda)\) of \(M(\lambda)\) is
+  finite-dimensional.
+\end{proposition}
+
+The proof of Proposition~\ref{thm:verma-is-finite-dim} is very technical and we
+won't include it here, but the idea behind it is to show that the set of
+weights of \(L(\lambda)\) is stable under the natural action of the Weyl group
+\(W\) on \(\mathfrak{h}^*\). One can then show that the every weight
+of \(L(\lambda)\) is conjugate to a single dominant integral weight of
+\(L(\lambda)\), and that the set of dominant integral weights of \(L(\lambda)\)
+is finite. Since \(W\) is finitely generated, this implies the set of
+weights of the unique simple quotient of \(M(\lambda)\) is finite. But
+each weight space is finite-dimensional. Hence so is the simple quotient
+\(L(\lambda)\).
+
+We refer the reader to \cite[ch. 21]{humphreys} for further details. We are now
+ready to prove the Highest Weight Theorem.
+
+\begin{proof}[Proof of Theorem~\ref{thm:dominant-weight-theo}]
+  We begin with the ``existence'' part of the theorem by showing that
+  \(L(\lambda)\) is indeed a finite-dimensional simple module whose
+  highest-weight is \(\lambda\). It suffices to show that the highest weight of
+  \(L(\lambda)\) is \(\lambda\). We have already seen that \(m^+ \in
+  M(\lambda)_\lambda\) is a highest weight vector. Now since \(m^+\) lies
+  outside of the maximal submodule of \(M(\lambda)\), the projection \(m^+ +
+  N(\lambda) \in L(\lambda)\) is nonzero.
+
+  We now claim that \(m^+ + N(\lambda) \in L(\lambda)_\lambda\). Indeed,
+  \[
+    H \cdot (m^+ + N(\lambda))
+    = H \cdot m^+ + N(\lambda)
+    = \lambda(H) (m^+ + N(\lambda))
+  \]
+  for all \(H \in \mathfrak{h}\). Hence \(\lambda\) is a weight of
+  \(L(\lambda)\), with weight vector \(m^+ + N(\lambda)\). Finally, we remark
+  that \(\lambda\) is the highest weight of \(L(\lambda)\), for if this was not
+  the case we could find a weight \(\mu\) of \(M(\lambda)\) with \(\mu \succ
+  \lambda\).
+
+  Now suppose \(M\) is some other finite-dimensional simple
+  \(\mathfrak{g}\)-module with highest weight vector \(m \in M_\lambda\). By
+  the universal property of the Verma module, there is a
+  \(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\) such that \(f(m^+) =
+  m\). As indicated before, since \(M\) is simple, \(M =
+  \mathcal{U}(\mathfrak{g}) \cdot m\) and therefore \(f\) is surjective. It
+  then follows \(M \cong \mfrac{M(\lambda)}{\ker f}\).
+
+  Since \(M\) is simple, \(\ker f \subset M(\lambda)\) is maximal and therefore
+  \(\ker f = N(\lambda)\). In other words, \(M \cong \mfrac{M(\lambda)}{\ker f}
+  = L(\lambda)\). We are done.
+\end{proof}
+
+We should point out that Proposition~\ref{thm:verma-is-finite-dim} fails for
+non-dominant \(\lambda \in P\). While \(\lambda\) is always a maximal weight of
+\(M(\lambda)\), one can show that if \(\lambda \in P\) is not dominant then
+\(N(\lambda) = 0\) and \(M(\lambda)\) is simple. For instance, if
+\(\mathfrak{g} = \mathfrak{sl}_2(K)\) and \(\lambda : h \mapsto -2\) then the
+action of \(\mathfrak{g}\) on \(M(\lambda)\) is given by
+\begin{center}
+  \begin{tikzcd}
+    \cdots             \rar[bend left=60]{-20}
+    & M(\lambda)_{-8}  \rar[bend left=60]{-12} \lar[bend left=60]{1}
+    & M(\lambda)_{-6}  \rar[bend left=60]{-6}  \lar[bend left=60]{1}
+    & M(\lambda)_{-4}  \rar[bend left=60]{-2}  \lar[bend left=60]{1}
+    & M(\lambda)_{-2}                          \lar[bend left=60]{1}
+  \end{tikzcd},
+\end{center}
+so we can see that \(M(\lambda)\) has no proper submodules. Verma modules can
+thus serve as examples of infinite-dimensional simple modules. Our next
+question is: what are \emph{all} the infinite-dimensional simple
+\(\mathfrak{g}\)-modules?
diff --git a/sections/mathieu.tex /dev/null
@@ -1,1611 +0,0 @@
-\chapter{Simple Weight Modules}\label{ch:mathieu}
-
-In this chapter we will expand our results on finite-dimensional simple modules
-of semisimple Lie algebras by considering \emph{infinite-dimensional}
-\(\mathfrak{g}\)-modules, which introduces numerous complications to our
-analysis. 
-
-For instance, in the infinite-dimensional setting we can no longer take
-complete-reducibility for granted. Indeed, we have seen that even if
-\(\mathfrak{g}\) is a semisimple Lie algebra, there are infinite-dimensional
-\(\mathfrak{g}\)-modules which are not semisimple. For a counterexample look no
-further than Example~\ref{ex:regular-mod-is-not-semisimple}: the regular
-\(\mathfrak{g}\)-module \(\mathcal{U}(\mathfrak{g})\) is never semisimple.
-Nevertheless, for simplicity -- or shall we say \emph{semisimplicity} -- we
-will focus exclusively on \emph{semisimple} \(\mathfrak{g}\)-modules. Our
-strategy is, once again, that of classifying simple modules. The regular
-\(\mathfrak{g}\)-module hides further unpleasant surprises, however: recall
-from Example~\ref{ex:regular-mod-is-not-weight-mod} that
-\[
-  \bigoplus_\lambda \mathcal{U}(\mathfrak{g})_\lambda
-  = 0
-  \subsetneq \mathcal{U}(\mathfrak{g})
-\]
-and the weight space decomposition fails for \(\mathcal{U}(\mathfrak{g})\).
-
-Indeed, our proof of the weight space decomposition in the finite-dimensional
-case relied heavily in the simultaneous diagonalization of commuting operators
-in a finite-dimensional space. Even if we restrict ourselves to simple modules,
-there is still a diverse spectrum of counterexamples to
-Corollary~\ref{thm:finite-dim-is-weight-mod} in the infinite-dimensional
-setting. For instance, any \(\mathfrak{g}\)-module \(M\) whose restriction to
-\(\mathfrak{h}\) is a free module satisfies \(M_\lambda = 0\) for all
-\(\lambda\) as in Example~\ref{ex:regular-mod-is-not-weight-mod}. These are
-called \emph{\(\mathfrak{h}\)-free \(\mathfrak{g}\)-modules}, and rank \(1\)
-simple \(\mathfrak{h}\)-free \(\mathfrak{sp}_{2 n}(K)\)-modules where first
-classified by Nilsson in \cite{nilsson}. Dimitar's construction of the so
-called \emph{exponential tensor \(\mathfrak{sl}_n(K)\)-modules} in
-\cite{dimitar-exp} is also an interesting source of counterexamples.
-
-Since the weight space decomposition was perhaps the single most instrumental
-ingredient of our previous analysis, it is only natural to restrict ourselves
-to the case it holds. This brings us to the following definition.
-
-\begin{definition}\label{def:weight-mod}\index{\(\mathfrak{g}\)-module!weight modules}\index{weights!weight modules}\index{\(\mathfrak{g}\)-module!(essential) support}
-  A \(\mathfrak{g}\)-module \(M\) is called a \emph{weight
-  \(\mathfrak{g}\)-module} if \(M = \bigoplus_{\lambda \in \mathfrak{h}^*}
-  M_\lambda\) and \(\dim M_\lambda < \infty\) for all \(\lambda \in
-  \mathfrak{h}^*\). The \emph{support of \(M\)} is the set
-  \(\operatorname{supp} M = \{\lambda \in \mathfrak{h}^* : M_\lambda \ne 0\}\).
-\end{definition}
-
-\begin{example}
-  Corollary~\ref{thm:finite-dim-is-weight-mod} is equivalent to the fact that
-  every finite-dimensional module of a semisimple Lie algebra is a weight
-  module. More generally, every finite-dimensional simple module of a reductive
-  Lie algebra is a weight module.
-\end{example}
-
-\begin{example}\label{ex:reductive-alg-equivalence}
-  We have seen that every finite-dimensional \(\mathfrak{g}\)-module is a
-  weight module for semisimple \(\mathfrak{g}\). In particular, if
-  \(\mathfrak{g}\) is finite-dimensional then the adjoint
-  \(\mathfrak{g}\)-module \(\mathfrak{g}\) is a weight module. More generally,
-  a finite-dimensional Lie algebra \(\mathfrak{g}\) is reductive if, and only
-  if the adjoint \(\mathfrak{g}\)-module \(\mathfrak{g}\) is a weight module,
-  in which case its weight spaces are given by the root spaces of
-  \(\mathfrak{g}\)
-\end{example}
-
-\begin{example}\label{ex:submod-is-weight-mod}
-  Proposition~\ref{thm:verma-is-weight-mod} and
-  Proposition~\ref{thm:max-verma-submod-is-weight} imply that the Verma module
-  \(M(\lambda)\) and its maximal submodule are both weight modules. In
-  fact, the proof of Proposition~\ref{thm:max-verma-submod-is-weight} is
-  actually a proof of the fact that every submodule \(N \subset M\) of
-  a weight module \(M\) is a weight module, and \(N_\lambda = M_\lambda \cap
-  N\) for all \(\lambda \in \mathfrak{h}^*\).
-\end{example}
-
-\begin{example}\label{ex:quotient-is-weight-mod}
-  Given a weight module \(M\), a submodule \(N \subset M\) and \(\lambda \in
-  \mathfrak{h}^*\), it is clear that \(\mfrac{M_\lambda}{N} \subset
-  \left(\mfrac{M}{N}\right)_\lambda\). In addition, \(\mfrac{M}{N} =
-  \bigoplus_{\lambda \in \mathfrak{h}^*} \mfrac{M_\lambda}{N}\). Hence
-  \(\mfrac{M}{N}\) is weight \(\mathfrak{g}\)-module with
-  \(\left(\mfrac{M}{N}\right)_\lambda = \mfrac{M_\lambda}{N} \cong
-  \mfrac{M_\lambda}{N_\lambda}\).
-\end{example}
-
-\begin{example}\label{ex:tensor-prod-of-weight-is-weight}
-  Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras, \(M_1\) be a
-  weight \(\mathfrak{g}_1\)-module and \(M_2\) a weight
-  \(\mathfrak{g}_2\)-module. Recall from Example~\ref{ex:cartan-direct-sum}
-  that if \(\mathfrak{h}_i \subset \mathfrak{g}_i\) are Cartan subalgebras then
-  \(\mathfrak{h} = \mathfrak{h}_1 \oplus \mathfrak{h}_2\) is a Cartan
-  subalgebra of \(\mathfrak{g} = \mathfrak{g}_1 \oplus \mathfrak{g}_2\) with
-  \(\mathfrak{h}^* = \mathfrak{h}_1^* \oplus \mathfrak{h}_2^*\). In this
-  setting, one can readily check that \(M_1 \otimes M_2\) is a weight
-  \(\mathfrak{g}\)-module with
-  \[
-    (M_1 \otimes M_2)_{\lambda_1 + \lambda_2}
-    = (M_1)_{\lambda_1} \otimes (M_2)_{\lambda_2}
-  \]
-  for all \(\lambda_i \in \mathfrak{h}_i^*\) and \(\operatorname{supp} M_1
-  \otimes M_2 = \operatorname{supp} M_1 \oplus \operatorname{supp} M_2 = \{
-  \lambda_1 + \lambda_2 : \lambda_i \in \operatorname{supp} M_i \subset
-  \mathfrak{h}_i^*\}\).
-\end{example}
-
-\begin{example}\label{thm:simple-weight-mod-is-tensor-prod}
-  Let \(\mathfrak{g} = \mathfrak{z} \oplus \mathfrak{s}_1 \oplus \cdots \oplus
-  \mathfrak{s}_r\) be a reductive Lie algebra, where \(\mathfrak{z}\) is the
-  center of \(\mathfrak{g}\) and \(\mathfrak{s}_1, \ldots, \mathfrak{s}_r\) are
-  its simple components. As in
-  Example~\ref{ex:all-simple-reps-are-tensor-prod}, any simple weight
-  \(\mathfrak{g}\)-module \(M\) can be decomposed as
-  \[
-    M \cong Z \otimes M_1 \otimes \cdots \otimes M_r
-  \]
-  where \(Z\) is a \(1\)-dimensional representation of \(\mathfrak{z}\) and
-  \(M_i\) is a simple weight \(\mathfrak{s}_i\)-module. The modules \(Z\) and
-  \(M_i\) are uniquely determined up to isomorphism.
-\end{example}
-
-\begin{example}\label{ex:adjoint-action-in-universal-enveloping-is-weight}
-  We would like to show that the requirement of finite-dimensionality in
-  Definition~\ref{def:weight-mod} is not redundant. Let \(\mathfrak{g}\) be a
-  finite-dimensional reductive Lie algebra and consider the adjoint
-  \(\mathfrak{g}\)-module \(\mathcal{U}(\mathfrak{g})\) -- where \(X \in
-  \mathfrak{g}\) acts by taking commutators. Given \(\alpha \in Q\), a simple
-  computation shows \(K \langle X_1 \cdots X_n H_1 \cdots H_m : X_i \in
-  \mathfrak{g}_{\alpha_i}, H_i \in \mathfrak{h}, \alpha_i \in \Delta, \alpha =
-  \alpha_1 + \cdots + \alpha_n \rangle \subset
-  \mathcal{U}(\mathfrak{g})_\alpha\). The PBW Theorem and
-  Example~\ref{ex:reductive-alg-equivalence} thus imply that
-  \(\mathcal{U}(\mathfrak{g}) = \bigoplus_{\alpha \in Q}
-  \mathcal{U}(\mathfrak{g})_\alpha\) where \(\mathcal{U}(\mathfrak{g})_\alpha =
-  K \langle X_1 \cdots X_n H_1 \cdots H_m : X_i \in \mathfrak{g}_{\alpha_i},
-  H_i \in \mathfrak{h}, \alpha_i \in \Delta, \alpha = \alpha_1 + \cdots +
-  \alpha_n \rangle\). However, \(\dim \mathcal{U}(\mathfrak{g})_\alpha =
-  \infty\). For instance, \(\mathcal{U}(\mathfrak{g})_0\) is \emph{precisely}
-  the commutator of \(\mathfrak{h}\) in \(\mathcal{U}(\mathfrak{g})\), which
-  contains \(\mathcal{U}(\mathfrak{h})\) and is therefore infinite-dimensional.
-\end{example}
-
-\begin{note}
-  We should stress that the weight spaces \(M_\lambda \subset M\) of a given
-  weight \(\mathfrak{g}\)-module \(M\) are \emph{not}
-  \(\mathfrak{g}\)-submodules. Nevertheless, \(M_\lambda\) is a
-  \(\mathfrak{h}\)-submodule. More generally, \(M_\lambda\) is a
-  \(\mathcal{U}(\mathfrak{g})_0\)-submodule, where
-  \(\mathcal{U}(\mathfrak{g})_0\) is the centralizer of \(\mathfrak{h}\) in
-  \(\mathcal{U}(\mathfrak{g})\) -- which coincides with the weight space of \(0
-  \in \mathfrak{h}^*\) in the adjoint \(\mathfrak{g}\)-module
-  \(\mathcal{U}(\mathfrak{g})\), as seen in
-  Example~\ref{ex:adjoint-action-in-universal-enveloping-is-weight}.
-\end{note}
-
-A particularly well behaved class of examples are the so called
-\emph{bounded} modules.
-
-\begin{definition}\index{\(\mathfrak{g}\)-module!bounded modules}\index{\(\mathfrak{g}\)-module!(essential) support}
-  A weight \(\mathfrak{g}\)-module \(M\) is called \emph{bounded} if \(\dim
-  M_\lambda\) is bounded. The lowest upper bound \(\deg M\) for \(\dim
-  M_\lambda\) is called \emph{the degree of \(M\)}. The \emph{essential
-  support} of \(M\) is the set \(\operatorname{supp}_{\operatorname{ess}} M =
-  \{ \lambda \in \mathfrak{h}^* : \dim M_\lambda = \deg M \}\).
-\end{definition}
-
-\begin{example}\label{ex:supp-ess-of-tensor-is-product}
-  Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras with Cartan
-  subalgebras \(\mathfrak{h}_i \subset \mathfrak{g}_i\) and take \(\mathfrak{g}
-  = \mathfrak{g}_1 \oplus \mathfrak{g}_2\). Given bounded
-  \(\mathfrak{g}_i\)-modules \(M_i\), it follows from
-  Example~\ref{ex:tensor-prod-of-weight-is-weight} that \(M_1 \otimes M_2\) is
-  a bounded \(\mathfrak{g}\)-module with \(\deg M_1 \otimes M_2 = \deg M_1
-  \cdot \deg M_2\) and 
-  \[
-    \operatorname{supp}_{\operatorname{ess}} M_1 \otimes M_2
-    = \operatorname{supp}_{\operatorname{ess}} M_1 \oplus
-      \operatorname{supp}_{\operatorname{ess}} M_2
-    = \{
-        \lambda_1 + \lambda_2 : \lambda_i \in
-        \operatorname{supp}_{\operatorname{ess}} M_i \subset \mathfrak{h}_i^*
-      \}
-  \]
-\end{example}
-
-\begin{example}\label{ex:laurent-polynomial-mod}
-  There is a natural action of \(\mathfrak{sl}_2(K)\) on the space \(K[x,
-  x^{-1}]\) of Laurent polynomials, given by the formulas in
-  (\ref{eq:laurent-polynomials-cusp-mod}). One can quickly verify \(K[x,
-  x^{-1}]_{2 k} = K x^k\) and \(K[x, x^{-1}]_\lambda = 0\) for any \(\lambda
-  \notin 2 \mathbb{Z}\), so that \(K[x, x^{-1}] = \bigoplus_{k \in \mathbb{Z}}
-  K x^k\) is a degree \(1\) bounded weight \(\mathfrak{sl}_2(K)\)-module. It
-  follows from the remark at the end of Example~\ref{ex:submod-is-weight-mod}
-  that any nonzero submodule \(N \subset K[x, x^{-1}]\) must contain a
-  monomial \(x^k\). But since the operators \(-\frac{\mathrm{d}}{\mathrm{d}x} +
-  \frac{x^{-1}}{2}, x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \frac{x}{2} : K[x,
-  x^{-1}] \to K[x, x^{-1}]\) are both injective, this implies all other
-  monomials can be found in \(N\) by successively applying \(f\) and \(e\).
-  Hence \(N = K[x, x^{-1}]\) and \(K[x, x^{-1}]\) is a simple module.
-  \begin{align}\label{eq:laurent-polynomials-cusp-mod}
-    e \cdot p
-    & = \left( x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \frac{x}{2} \right) p &
-    f \cdot p
-    & = \left(- \frac{\mathrm{d}}{\mathrm{d}x} + \frac{x^{-1}}{2} \right) p &
-    h \cdot p
-    & = 2 x \frac{\mathrm{d}}{\mathrm{d}x} p
-  \end{align}
-\end{example}
-
-Notice that the support of \(K[x, x^{-1}]\) is the trivial \(2
-\mathbb{Z}\)-coset \(0 + 2 \mathbb{Z}\). This is representative of the general
-behavior in the following sense: if \(M\) is a simple weight
-\(\mathfrak{g}\)-module, since \(M[\lambda] = \bigoplus_{\alpha \in Q}
-M_{\lambda + \alpha}\) is stable under the action of \(\mathfrak{g}\) for all
-\(\lambda \in \mathfrak{h}^*\), \(\bigoplus_{\alpha \in Q} M_{\lambda +
-\alpha}\) is either \(0\) or all of \(M\). In other words, the support of a
-simple weight module is always contained in a single \(Q\)-coset.
-
-However, the behavior of \(K[x, x^{-1}]\) deviates from that of an arbitrary
-bounded \(\mathfrak{g}\)-module in the sense its essential support is
-precisely the entire \(Q\)-coset it inhabits -- i.e.
-\(\operatorname{supp}_{\operatorname{ess}} K[x, x^{-1}] = 2 \mathbb{Z}\). This
-isn't always the case. Nevertheless, in general we find\dots
-
-\begin{proposition}\label{thm:ess-supp-is-zariski-dense}
-  Let \(\mathfrak{g}\) be a finite-dimensional semisimple Lie algebra and \(M\)
-  be a simple infinite-dimensional bounded \(\mathfrak{g}\)-module. The
-  essential support \(\operatorname{supp}_{\operatorname{ess}} M\) is
-  Zariski-dense\footnote{Any choice of basis for $\mathfrak{h}^*$ induces a
-  $K$-linear isomorphism $\mathfrak{h}^* \isoto K^n$. In particular, a choice
-  of basis induces a unique topology in $\mathfrak{h}^*$ such that the map
-  $\mathfrak{h}^* \to K^n$ is a homeomorphism onto $K^n$ with the Zariski
-  topology. Any two basis induce the same topology in $\mathfrak{h}^*$, which
-  we call \emph{the Zariski topology of $\mathfrak{h}^*$}.} in
-  \(\mathfrak{h}^*\).
-\end{proposition}
-
-This proof was deemed too technical to be included in here, but see Proposition
-3.5 of \cite{mathieu} for the case where \(\mathfrak{g} = \mathfrak{s}\) is a
-simple Lie algebra. The general case then follows from
-Example~\ref{thm:simple-weight-mod-is-tensor-prod},
-Example~\ref{ex:supp-ess-of-tensor-is-product} and the asserting that the
-product of Zariski-dense subsets in \(K^n\) and \(K^m\) is Zariski-dense in
-\(K^{n + m} = K^n \times K^m\).
-
-We now begin a systematic investigation of the problem of classifying the
-infinite-dimensional simple weight modules of a given Lie algebra
-\(\mathfrak{g}\). As in the previous chapter, let \(\mathfrak{g}\) be a
-finite-dimensional semisimple Lie algebra. As a first approximation of a
-solution to our problem, we consider the induction functors
-\(\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}} :
-\mathfrak{p}\text{-}\mathbf{Mod} \to \mathfrak{g}\text{-}\mathbf{Mod}\), where
-\(\mathfrak{p} \subset \mathfrak{g}\) is some subalgebra. 
-
-% TODOO: Are you sure that these are indeed the weight spaces of the induced
-% module? Check this out?
-These functors have already proved themselves a powerful tool for constructing
-modules in the previous chapters. Our first observation is that if
-\(\mathfrak{p} \subset \mathfrak{g}\) contains the Borel subalgebra
-\(\mathfrak{b}\) then \(\mathfrak{h}\) is a Cartan subalgebra of
-\(\mathfrak{p}\) and \((\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}}
-M)_\lambda = \mathcal{U}(\mathfrak{g}) \otimes_{\mathcal{U}(\mathfrak{p})}
-M_\lambda\) for all \(\lambda \in \mathfrak{h}^*\). In particular,
-\(\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}}\) takes weight
-\(\mathfrak{p}\)-modules to weight \(\mathfrak{g}\)-modules. This leads us to
-the following definition.
-
-\begin{definition}\index{Lie subalgebra!parabolic subalgebra}
-  A subalgebra \(\mathfrak{p} \subset \mathfrak{g}\) is called \emph{parabolic}
-  if \(\mathfrak{b} \subset \mathfrak{p}\).
-\end{definition}
-
-% TODOO: Why is the fact that p is not reductive relevant?? Why do we need to
-% look at the quotient by nil(p)??
-Parabolic subalgebras thus give us a process for constructing weight
-\(\mathfrak{g}\)-modules from modules of smaller (parabolic) subalgebras. Our
-hope is that by iterating this process again and again we can get a large class
-of simple weight \(\mathfrak{g}\)-modules. However, there is a small catch: a
-parabolic subalgebra \(\mathfrak{p} \subset \mathfrak{g}\) needs not to be
-reductive. We can get around this limitation by modding out by
-\(\mathfrak{nil}(\mathfrak{p})\) and noticing that
-\(\mathfrak{nil}(\mathfrak{p})\) acts trivially in any weight
-\(\mathfrak{p}\)-module \(M\). By applying the universal property of quotients
-we can see that \(M\) has the natural structure of a
-\(\mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}\)-module, which is always
-a reductive algebra.
-\begin{center}
-  \begin{tikzcd}
-    \mathfrak{p}                                       \rar \dar          &
-    \mathfrak{gl}(M)                                                      \\
-    \mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})} \arrow[dotted]{ur} &
-  \end{tikzcd}
-\end{center}
-
-Let \(\mathfrak{p}\) be a parabolic subalgebra and \(M\) be a simple
-weight \(\mathfrak{p}\)-module. We should point out that while
-\(\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}} M\) is a weight
-\(\mathfrak{g}\)-module, it isn't necessarily simple. Nevertheless, we can
-use it to produce a simple weight \(\mathfrak{g}\)-module via a
-construction very similar to that of Verma modules.
-
-\begin{definition}\index{\(\mathfrak{g}\)-module!(generalized) Verma modules}
-  Given any \(\mathfrak{p}\)-module \(M\), the module \(M_{\mathfrak{p}}(M) =
-  \operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}} M\) is called \emph{a
-  generalized Verma module}.
-\end{definition}
-
-\begin{proposition}\label{thm:generalized-verma-has-simple-quotient}
-  Given a simple \(\mathfrak{p}\)-module \(M\), the generalized Verma module
-  \(M_{\mathfrak{p}}(M)\) has a unique maximal \(\mathfrak{p}\)-submodule
-  \(N_{\mathfrak{p}}(M)\) and a unique irreducible quotient
-  \(L_{\mathfrak{p}}(M) = \mfrac{M_{\mathfrak{p}}(M)}{N_{\mathfrak{p}}(M)}\).
-  The irreducible quotient \(L_{\mathfrak{p}}(M)\) is a weight module.
-\end{proposition}
-
-The proof of Proposition~\ref{thm:generalized-verma-has-simple-quotient} is
-entirely analogous to that of Proposition~\ref{thm:max-verma-submod-is-weight}.
-This leads us to the following definitions.
-
-\begin{definition}\index{\(\mathfrak{g}\)-module!parabolic induced modules}\index{\(\mathfrak{g}\)-module!cuspidal modules}
-  A \(\mathfrak{g}\)-module is called \emph{parabolic induced} if it is
-  isomorphic to \(L_{\mathfrak{p}}(M)\) for some proper parabolic subalgebra
-  \(\mathfrak{p} \subsetneq \mathfrak{g}\) and some \(\mathfrak{p}\)-module
-  \(M\). An \emph{simple cuspidal \(\mathfrak{g}\)-module} is a simple
-  \(\mathfrak{g}\)-module which is \emph{not} parabolic induced.
-\end{definition}
-
-Since every weight \(\mathfrak{p}\)-module \(M\) is an
-\(\mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}\)-module, it makes sense
-to call \(M\) \emph{cuspidal} if it is a cuspidal
-\(\mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}\)-module. The first
-breakthrough regarding our classification problem was given by Fernando in his
-now infamous paper \citetitle{fernando} \cite{fernando}, where he proved that
-every simple weight \(\mathfrak{g}\)-module is parabolic induced. In other
-words\dots
-
-\begin{theorem}[Fernando]
-  Any simple weight \(\mathfrak{g}\)-module is isomorphic to
-  \(L_{\mathfrak{p}}(M)\) for some parabolic subalgebra \(\mathfrak{p} \subset
-  \mathfrak{g}\) and some simple cuspidal \(\mathfrak{p}\)-module \(M\).
-\end{theorem}
-
-We should point out that the relationship between simple weight
-\(\mathfrak{g}\)-modules and pairs \((\mathfrak{p}, M)\) -- where
-\(\mathfrak{p}\) is some parabolic subalgebra and \(M\) is a simple cuspidal
-\(\mathfrak{p}\)-module -- is not one-to-one. Nevertheless, this relationship
-is well understood. Namely, Fernando himself established\dots
-
-\begin{proposition}[Fernando]
-  Given a parabolic subalgebra \(\mathfrak{p} \subset \mathfrak{g}\), there
-  exists a basis \(\Sigma\) for \(\Delta\) such that \(\Sigma \subset
-  \Delta_{\mathfrak{p}} \subset \Delta\), where \(\Delta_{\mathfrak{p}}\)
-  denotes the set of roots of \(\mathfrak{p}\). Furthermore, if \(\mathfrak{p}'
-  \subset \mathfrak{g}\) is another parabolic subalgebra, \(M\) is a simple
-  cuspidal \(\mathfrak{p}\)-module and \(N\) is a simple cuspidal
-  \(\mathfrak{p}'\)-module then \(L_{\mathfrak{p}}(M) \cong
-  L_{\mathfrak{p}'}(N)\) if, and only if \(\mathfrak{p}' =
-  \twisted{\mathfrak{p}}{\sigma}\) and \(M \cong \twisted{N}{\sigma}\) for
-  some\footnote{Here $\twisted{\mathfrak{p}}{\sigma}$ denotes the image of
-  $\mathfrak{p}$ under the automorphism of $\sigma : \mathfrak{g} \to
-  \mathfrak{g}$ given by the canonical action of $W$ on $\mathfrak{g}$ and
-  $\twisted{N}{\sigma}$ is the $\mathfrak{p}$-module given by composing the map
-  $\mathfrak{p}' \to \mathfrak{gl}(N)$ with the restriction
-  $\sigma\!\restriction_{\mathfrak{p}} : \mathfrak{p} \to \mathfrak{p}'$.}
-  \(\sigma \in W_M\), where
-  \[
-    W_M
-    = \langle
-      \sigma_\beta : \beta \in \Sigma, H_\beta + \mathfrak{nil}(\mathfrak{p})
-      \ \text{is central in}\ \mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}
-      \ \text{and}\ H_\beta\ \text{acts on \(M\) as a positive integer}
-      \rangle
-    \subset W
-  \]
-\end{proposition}
-
-\begin{note}
-  The definition of the subgroup \(W_M \subset W\) is independent of the choice
-  of basis \(\Sigma\).
-\end{note}
-
-As a first consequence of Fernando's Theorem, we provide two alternative
-characterizations of cuspidal modules.
-
-\begin{corollary}[Fernando]\label{thm:cuspidal-mod-equivs}
-  Let \(M\) be a simple weight \(\mathfrak{g}\)-module. The following
-  conditions are equivalent.
-  \begin{enumerate}
-    \item \(M\) is cuspidal.
-    \item \(F_\alpha\) acts injectively on \(M\) for all
-      \(\alpha \in \Delta\) -- this is what is usually referred
-      to as a \emph{dense} module in the literature.
-    \item The support of \(M\) is precisely one \(Q\)-coset -- this is
-      what is usually referred to as a \emph{torsion-free} module in the
-      literature.
-  \end{enumerate}
-\end{corollary}
-
-\begin{example}
-  As noted in Example~\ref{ex:laurent-polynomial-mod}, the element \(f \in
-  \mathfrak{sl}_2(K)\) acts injectively on the space of Laurent polynomials.
-  Hence \(K[x, x^{-1}]\) is a cuspidal \(\mathfrak{sl}_2(K)\)-module.
-\end{example}
-
-Having reduced our classification problem to that of classifying simple
-cuspidal modules, we are now faced the daunting task of actually classifying
-them. Historically, this was first achieved by Olivier Mathieu in the early
-2000's in his paper \citetitle{mathieu} \cite{mathieu}. To do so, Mathieu
-introduced new tools which have since proved themselves remarkably useful
-throughout the field, known as\dots
-
-\section{Coherent Families}
-
-We begin our analysis with a simple question: how to do we go about
-constructing cuspidal modules? Specifically, given a cuspidal
-\(\mathfrak{g}\)-module, how can we use it to produce new cuspidal modules? To
-answer this question, we look back at the single example of a cuspidal module
-we have encountered so far: the \(\mathfrak{sl}_2(K)\)-module \(K[x, x^{-1}]\)
-of Laurent polynomials -- i.e. Example~\ref{ex:laurent-polynomial-mod}.
-
-Our first observation is that \(\mathfrak{sl}_2(K)\) acts on \(K[x, x^{-1}]\)
-via differential operators. In other words, the action map
-\(\mathcal{U}(\mathfrak{sl}_2(K)) \to \operatorname{End}(K[x, x^{-1}])\)
-factors through the inclusion of the algebra \(\operatorname{Diff}(K[x,
-x^{-1}]) = K\left[x, x^{-1}, \frac{\mathrm{d}}{\mathrm{d}x}\right]\) of
-differential operators in \(K[x, x^{-1}]\).
-\begin{center}
-  \begin{tikzcd}
-    \mathcal{U}(\mathfrak{sl}_2(K))   \rar &
-    \operatorname{Diff}(K[x, x^{-1}]) \rar &
-    \operatorname{End}(K[x, x^{-1}])
-  \end{tikzcd}
-\end{center}
-
-The space \(K[x, x^{-1}]\) can be regarded as a \(\operatorname{Diff}(K[x,
-x^{-1}])\)-module in the natural way, and we can produce new
-\(\operatorname{Diff}(K[x, x^{-1}])\)-modules by twisting \(K[x, x^{-1}]\) by
-automorphisms of \(\operatorname{Diff}(K[x, x^{-1}])\). For example, given
-\(\lambda \in K\) we may take the automorphism
-\begin{align*}
-  \varphi_\lambda : \operatorname{Diff}(K[x, x^{-1}]) &
-  \to \operatorname{Diff}(K[x, x^{-1}]) \\
-  x & \mapsto x \\
-  x^{-1} & \mapsto x^{-1} \\
-  \frac{\mathrm{d}}{\mathrm{d} x} & \mapsto \frac{\mathrm{d}}{\mathrm{d} x} +
-  \frac{\lambda}{2} x^{-1}
-\end{align*}
-and consider the twisted module \(\twisted{K[x, x^{-1}]}{\varphi_\lambda} =
-K[x, x^{-1}]\), where some operator \(P \in \operatorname{Diff}(K[x, x^{-1}])\)
-acts as \(\varphi_\lambda(P)\).
-
-By composing the action map \(\operatorname{Diff}(K[x, x^{-1}]) \to
-\operatorname{End}(\twisted{K[x, x^{-1}]}{\varphi_\lambda})\) with the
-homomorphism of algebras \(\mathcal{U}(\mathfrak{sl}_2(K)) \to
-\operatorname{Diff}(K[x, x^{-1}])\) we can give \(\twisted{K[x,
-x^{-1}]}{\varphi_\lambda}\) the structure of an \(\mathfrak{sl}_2(K)\)-module.
-Diagrammatically, we have
-\begin{center}
-  \begin{tikzcd}
-    \mathcal{U}(\mathfrak{sl}_2(K))   \rar                  &
-    \operatorname{Diff}(K[x, x^{-1}]) \rar{\varphi_\lambda} &
-    \operatorname{Diff}(K[x, x^{-1}]) \rar                  &
-    \operatorname{End}(K[x, x^{-1}])
-  \end{tikzcd},
-\end{center}
-where the maps \(\mathcal{U}(\mathfrak{sl}_2(K)) \to \operatorname{Diff}(K[x,
-x^{-1}])\) and \(\operatorname{Diff}(K[x, x^{1}]) \to \operatorname{End}(K[x,
-x^{-1}])\) are the ones from the previous diagram.
-
-Explicitly, we find that the action of \(\mathfrak{sl}_2(K)\) on
-\(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is given by
-\begin{align*}
-  p & \overset{e}{\mapsto}
-  \left(
-  x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \frac{1 + \lambda}{2} x
-  \right) p &
-  p & \overset{f}{\mapsto}
-  \left(
-  - \frac{\mathrm{d}}{\mathrm{d}x} + \frac{1 - \lambda}{2} x^{-1}
-  \right) p &
-  p & \overset{h}{\mapsto}
-  \left( 2 x \frac{\mathrm{d}}{\mathrm{d}x} + \lambda \right) p,
-\end{align*}
-so we can see \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}_{2 k +
-\frac{\lambda}{2}} = K x^k\) for all \(k \in \mathbb{Z}\) and \(\twisted{K[x,
-x^{-1}]}{\varphi_\lambda}_\mu = 0\) for all other \(\mu \in \mathfrak{h}^*\).
-
-Hence \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is a degree \(1\) bounded
-\(\mathfrak{sl}_2(K)\)-module with \(\operatorname{supp} \twisted{K[x,
-x^{-1}]}{\varphi_\lambda} = \frac{\lambda}{2} + 2 \mathbb{Z}\). One can also
-quickly check that if \(\lambda \notin 1 + 2 \mathbb{Z}\) then \(e\) and \(f\)
-act injectively in \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\), so that
-\(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is simple. In particular, if
-\(\lambda, \mu \notin 1 + 2 \mathbb{Z}\) with \(\lambda \notin \mu + 2
-\mathbb{Z}\) then \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) and
-\(\twisted{K[x, x^{-1}]}{\varphi_\mu}\) are non-isomorphic simple cuspidal
-\(\mathfrak{sl}_2(K)\)-modules, since their supports differ. These cuspidal
-modules can be ``glued together'' in a \emph{monstrous concoction} by summing
-over \(\lambda \in K\), as in
-\[
-  \mathcal{M}
-  = \bigoplus_{\lambda + 2 \mathbb{Z} \in \mfrac{K}{2 \mathbb{Z}}}
-    \twisted{K[x, x^{-1}]}{\varphi_\lambda},
-\]
-
-To a distracted spectator, \(\mathcal{M}\) may look like just another,
-innocent, \(\mathfrak{sl}_2(K)\)-module. However, the attentive reader may have
-already noticed some of the its bizarre features, most noticeable of which is
-the fact that \(\mathcal{M}\) is very big. In fact, \(\mathcal{M}\) is as big a
-degree \(1\) bounded module gets: \(\operatorname{supp} \mathcal{M}
-= \operatorname{supp}_{\operatorname{ess}} \mathcal{M}\) is the entirety of
-\(\mathfrak{h}^*\). This may look very alien the reader familiarized with the
-finite-dimensional setting, where the configuration of weights is very rigid.
-For this reason, \(\mathcal{M}\) deserves to be called ``a monstrous
-concoction''.
-
-On a perhaps less derogatory note, \(\mathcal{M}\) also deserves to be called
-\emph{a family}. This is because \(\mathcal{M}\) consists of lots of smaller
-cuspidal modules which fit together inside of it in a \emph{coherent} fashion.
-Mathieu's ingenious breakthrough was the realization that \(\mathcal{M}\) is a
-particular example of a more general pattern, which he named \emph{coherent
-families}.
-
-\begin{definition}\index{coherent family}
-  A \emph{coherent family \(\mathcal{M}\) of degree \(d\)} is a weight
-  \(\mathfrak{g}\)-module \(\mathcal{M}\) such that
-  \begin{enumerate}
-    \item \(\dim \mathcal{M}_\lambda = d\) for \emph{all} \(\lambda \in
-      \mathfrak{h}^*\) -- i.e. \(\operatorname{supp}_{\operatorname{ess}}
-      \mathcal{M} = \mathfrak{h}^*\).
-    \item For any \(u \in \mathcal{U}(\mathfrak{g})\) in the centralizer
-      \(\mathcal{U}(\mathfrak{g})_0\) of \(\mathfrak{h}\) in
-      \(\mathcal{U}(\mathfrak{g})\), the map
-      \begin{align*}
-        \mathfrak{h}^* & \to K \\
-               \lambda & \mapsto
-               \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\lambda})
-      \end{align*}
-      is polynomial in \(\lambda\).
-  \end{enumerate}
-\end{definition}
-
-\begin{example}\label{ex:sl-laurent-family}
-  The module \(\mathcal{M} = \bigoplus_{\lambda + 2 \mathbb{Z} \in
-  \mfrac{K}{2 \mathbb{Z}}} \twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is a
-  degree \(1\) coherent \(\mathfrak{sl}_2(K)\)-family.
-\end{example}
-
-\begin{example}
-  Given \(\lambda \in K\), \(\mathcal{M}(\lambda) = \bigoplus_{\mu \in K} K
-  x^\mu\) with
-  \begin{align*}
-    p & \overset{e}{\mapsto}
-        \left(x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \lambda x\right) p &
-    p & \overset{f}{\mapsto}
-        \left(-\frac{\mathrm{d}}{\mathrm{d}x} + \lambda x^{-1}\right) p &
-    p & \overset{h}{\mapsto} 2 x \frac{\mathrm{d}}{\mathrm{d}x} p,
-  \end{align*}
-  is a degree \(1\) coherent \(\mathfrak{sl}_2(K)\)-family -- where \(x^{\pm
-  1}, \sfrac{\mathrm{d}}{\mathrm{d}x} : \mathcal{M}(\lambda) \to
-  \mathcal{M}(\lambda)\) are given by \(x^{\pm 1} x^\mu = x^{\mu \pm 1}\) and
-  \(\sfrac{\mathrm{d}}{\mathrm{d}x} x^\mu = \mu x^{\mu - 1}\). It is easy to
-  check \(\mathcal{M}\) from Example~\ref{ex:sl-laurent-family} is isomorphic
-  to \(\mathcal{M}(\sfrac{1}{2})\) and \((\mathcal{M}(\sfrac{1}{2}))[0] \cong
-  K[x, x^{-1}]\).
-\end{example}
-
-\begin{note}
-  We would like to stress that coherent families have proven themselves useful
-  for problems other than the classification of cuspidal
-  \(\mathfrak{g}\)-modules. For instance, Nilsson's classification of rank 1
-  \(\mathfrak{h}\)-free \(\mathfrak{sp}_{2 n}(K)\)-modules is based on the
-  notion of coherent families and the so called \emph{weighting functor}.
-\end{note}
-
-Our hope is that given a simple cuspidal module \(M\), we can somehow fit \(M\)
-inside of a coherent \(\mathfrak{g}\)-family, such as in the case of \(K[x,
-x^{-1}]\) and \(\mathcal{M}\) from Example~\ref{ex:sl-laurent-family}. In
-addition, we hope that such coherent families are somehow \emph{uniquely
-determined} by \(M\). This leads us to the following definition.
-
-\begin{definition}\index{coherent family!coherent extension}
-  Given a bounded \(\mathfrak{g}\)-module \(M\) of degree \(d\), a
-  \emph{coherent extension \(\mathcal{M}\) of \(M\)} is a coherent family
-  \(\mathcal{M}\) of degree \(d\) that contains \(M\) as a subquotient.
-\end{definition}
-
-Our goal is now showing that every simple bounded module has a coherent
-extension. The idea then is to classify coherent families, and classify which
-submodules of a given coherent family are actually simple cuspidal modules. If
-every simple bounded \(\mathfrak{g}\)-module fits inside a coherent extension,
-this would lead to classification of all simple cuspidal
-\(\mathfrak{g}\)-modules, which we now know is the key for the solution of our
-classification problem. However, there are some complications to this scheme.
-
-Leaving aside the question of existence for a second, we should point out that
-coherent families turn out to be rather complicated on their own. In fact they
-are too complicated to classify in general. Ideally, we would like to find
-\emph{nice} coherent extensions -- ones we can actually classify. For instance,
-we may search for \emph{irreducible} coherent extensions, which are defined as
-follows.
-
-\begin{definition}\index{coherent family!irreducible coherent family}
-  A coherent family \(\mathcal{M}\) is called \emph{irreducible} if it contains
-  no proper coherent subfamilies -- i.e. \(\mathcal{M}\) is a simple object in
-  the full subcategory of \(\mathfrak{g}\text{-}\mathbf{Mod}\) consisting of
-  coherent families. Equivalently, we call \(\mathcal{M}\) irreducible if
-  \(\mathcal{M}_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module
-  for some \(\lambda \in \mathfrak{h}^*\).
-\end{definition}
-
-Another natural candidate for the role of ``nice extensions'' are the
-semisimple coherent families -- i.e. families which are semisimple as
-\(\mathfrak{g}\)-modules. These turn out to be very easy to produce. Namely,
-there is a construction, known as \emph{the semisimplification of a coherent
-family}, which takes a coherent extension of \(M\) to a semisimple coherent
-extension of \(M\).
-
-% Mathieu's proof of this is somewhat profane, I don't think it's worth
-% including it in here
-% TODO: Move this somewhere else? This holds in general for weight modules
-% whose suppert is contained in a single Q-coset
-\begin{lemma}\label{thm:component-coh-family-has-finite-length}
-  Given a coherent family \(\mathcal{M}\) and \(\lambda \in \mathfrak{h}^*\),
-  \(\mathcal{M}[\lambda]\) has finite length as a \(\mathfrak{g}\)-module.
-\end{lemma}
-
-\begin{proposition}\index{coherent family!semisimplification}
-  Let \(\mathcal{M}\) be a coherent family of degree \(d\). There exists a
-  unique semisimple coherent family \(\mathcal{M}^{\operatorname{ss}}\) of
-  degree \(d\) such that the composition series of
-  \(\mathcal{M}^{\operatorname{ss}}[\lambda]\) is the same as that of
-  \(\mathcal{M}[\lambda]\) for all \(\lambda \in \mathfrak{h}^*\), called
-  \emph{the semisimplification of \(\mathcal{M}\)}.
-
-  Namely, if \(\lambda \in \mathfrak{h}^*\) and \(0 = \mathcal{M}_{\lambda 0}
-  \subset \mathcal{M}_{\lambda 1} \subset \cdots \subset \mathcal{M}_{\lambda
-  r_\lambda} = \mathcal{M}[\lambda]\) is a composition series\footnote{Notice
-  that $\mathcal{M}[\lambda] = \mathcal{M}[\mu]$ for any $\mu \in \lambda + Q$.
-  Hence the sum $\bigoplus_{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q}}
-  \bigoplus_i \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}$ is
-  independent of the choice of representative for $\lambda + Q$ -- provided we
-  choose $\mathcal{M}_{\mu i} = \mathcal{M}_{\lambda i}$ for all $\mu \in
-  \lambda + Q$ and $i$.},
-  \[
-    \mathcal{M}^{\operatorname{ss}}
-    \cong \bigoplus_{\substack{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q} \\ i}}
-          \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
-  \]
-\end{proposition}
-
-\begin{proof}
-  The uniqueness of \(\mathcal{M}^{\operatorname{ss}}\) should be clear:
-  since \(\mathcal{M}^{\operatorname{ss}}\) is semisimple, so is
-  \(\mathcal{M}^{\operatorname{ss}}[\lambda]\). Hence by the Jordan-Hölder
-  Theorem
-  \[
-    \mathcal{M}^{\operatorname{ss}}[\lambda]
-    \cong
-    \bigoplus_i \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
-  \]
-
-  As for the existence of the semisimplification, it suffices to show
-  \[
-    \mathcal{M}^{\operatorname{ss}}
-    = \bigoplus_{\substack{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q} \\ i}}
-    \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
-  \]
-  is indeed a semisimple coherent family of degree \(d\).
-
-  We know from Examples~\ref{ex:submod-is-weight-mod} and
-  \ref{ex:quotient-is-weight-mod} that each quotient
-  \(\mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}\) is a weight
-  module. Hence \(\mathcal{M}^{\operatorname{ss}}\) is a weight module.
-  Furthermore, given \(\mu \in \mathfrak{h}^*\)
-  \[
-    \mathcal{M}_\mu^{\operatorname{ss}}
-    = \bigoplus_{\substack{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q} \\ i}}
-      \left(
-      \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
-      \right)_\mu
-    = \bigoplus_i
-      \left(
-      \mfrac{\mathcal{M}_{\mu i + 1}}{\mathcal{M}_{\mu i}}
-      \right)_\mu
-    \cong \bigoplus_i
-      \mfrac{(\mathcal{M}_{\mu i + 1})_\mu}
-            {(\mathcal{M}_{\mu i})_\mu}
-  \]
-
-  In particular,
-  \[
-    \dim \mathcal{M}_\mu^{\operatorname{ss}}
-    = \sum_i
-      \dim (\mathcal{M}_{\mu i + 1})_\mu - \dim (\mathcal{M}_{\mu i})_\mu
-    = \dim \mathcal{M}[\mu]_\mu
-    = \dim \mathcal{M}_\mu
-    = d
-  \]
-
-  Likewise, given \(u \in \mathcal{U}(\mathfrak{g})_0\) the value
-  \[
-    \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\mu^{\operatorname{ss}}})
-    = \sum_i
-      \operatorname{Tr}(u\!\restriction_{(\mathcal{M}_{\mu i + 1})_\mu})
-    - \operatorname{Tr}(u\!\restriction_{(\mathcal{M}_{\mu i})_\mu})
-    = \operatorname{Tr}(u\!\restriction_{\mathcal{M}[\mu]_\mu})
-    = \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\mu})
-  \]
-  is polynomial in \(\mu \in \mathfrak{h}^*\).
-\end{proof}
-
-\begin{note}
-  Although we have provided an explicit construction of
-  \(\mathcal{M}^{\operatorname{ss}}\) in terms of \(\mathcal{M}\), we should
-  point out this construction is not functorial. First, given a
-  \(\mathfrak{g}\)-homomorphism \(f : \mathcal{M} \to \mathcal{N}\) between
-  coherent families, it is unclear what \(f^{\operatorname{ss}} :
-  \mathcal{M}^{\operatorname{ss}} \to \mathcal{N}^{\operatorname{ss}}\) is
-  supposed to be. Secondly, and this is more relevant, our construction depends
-  on the choice of composition series \(0 = \mathcal{M}_{\lambda 0} \subset
-  \cdots \subset \mathcal{M}_{\lambda r_\lambda} = \mathcal{M}[\lambda]\).
-  While different choices of composition series yield isomorphic results, there
-  is no canonical isomorphism. In addition, there is no canonical choice of
-  composition series.
-\end{note}
-
-The proof of Lemma~\ref{thm:component-coh-family-has-finite-length} is
-extremely technical and will not be included in here. It suffices to note that,
-as in Proposition~\ref{thm:ess-supp-is-zariski-dense}, the general case follows
-from the case where \(\mathfrak{g}\) is simple, which may be found in
-\cite{mathieu} -- see Lemma 3.3. As promised, if \(\mathcal{M}\) is a coherent
-extension of \(M\) then so is \(\mathcal{M}^{\operatorname{ss}}\).
-
-\begin{proposition}
-  Let \(M\) be a simple bounded \(\mathfrak{g}\)-module and \(\mathcal{M}\)
-  be a coherent extension of \(M\). Then \(\mathcal{M}^{\operatorname{ss}}\) is
-  a coherent extension of \(M\), and \(M\) is in fact a submodule of
-  \(\mathcal{M}^{\operatorname{ss}}\).
-\end{proposition}
-
-\begin{proof}
-  Since \(M\) is simple, its support is contained in a single \(Q\)-coset.
-  This implies that \(M\) is a subquotient of \(\mathcal{M}[\lambda]\) for any
-  \(\lambda \in \operatorname{supp} M\). If we fix some composition series \(0
-  = \mathcal{M}_0 \subset \mathcal{M}_1 \subset \cdots \subset \mathcal{M}_r =
-  \mathcal{M}[\lambda]\) of \(\mathcal{M}[\lambda]\) with \(M \cong
-  \mfrac{\mathcal{M}_{i + 1}}{\mathcal{M}_i}\), there is a natural inclusion
-  \[
-    M
-    \isoto \mfrac{\mathcal{M}_{i + 1}}{\mathcal{M}_i}
-    \to \bigoplus_j \mfrac{\mathcal{M}_{j + 1}}{\mathcal{M}_j}
-    \cong \mathcal{M}^{\operatorname{ss}}[\lambda]
-  \]
-\end{proof}
-
-Given the uniqueness of the semisimplification, the semisimplification of any
-semisimple coherent extension \(\mathcal{M}\) is \(\mathcal{M}\)
-itself and therefore\dots
-
-\begin{corollary}\label{thm:bounded-is-submod-of-extension}
-  Let \(M\) be a simple bounded \(\mathfrak{g}\)-module and \(\mathcal{M}\)
-  be a semisimple coherent extension of \(M\). Then \(M\) is
-  contained in \(\mathcal{M}\).
-\end{corollary}
-
-These last results provide a partial answer to the question of existence of
-well behaved coherent extensions. As for the uniqueness \(\mathcal{M}\) in
-Corollary~\ref{thm:bounded-is-submod-of-extension}, it suffices to show that
-the multiplicities of the simple weight \(\mathfrak{g}\)-modules in
-\(\mathcal{M}\) are uniquely determined by \(M\). These multiplicities may be
-computed via the following lemma.
-
-\begin{lemma}\label{thm:centralizer-multiplicity}
-  Let \(M\) be a semisimple weight \(\mathfrak{g}\)-module. Then \(M_\lambda\)
-  is a semisimple \(\mathcal{U}(\mathfrak{g})_0\)-module for any \(\lambda \in
-  \operatorname{supp} M\). Moreover, if \(L\) is a simple weight
-  \(\mathfrak{g}\)-module such that \(\lambda \in \operatorname{supp} L\) then
-  \(L_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module and the
-  multiplicity \(L\) in \(M\) coincides with the multiplicity of \(L_\lambda\)
-  in \(M_\lambda\) as a \(\mathcal{U}(\mathfrak{g})_0\)-module.
-\end{lemma}
-
-\begin{proof}
-  We begin by showing that \(L_\lambda\) is simple. Let \(N \subset L_\lambda\)
-  be a nontrivial \(\mathcal{U}(\mathfrak{g})_0\)-submodule. We want to
-  establish that \(N = L_\lambda\).
-
-  If \(\mathcal{U}(\mathfrak{g})_\alpha\) denotes the root space of \(\alpha\)
-  in \(\mathcal{U}(\mathfrak{g})\) under the adjoint action of \(\mathfrak{g}\)
-  as in Example~\ref{ex:adjoint-action-in-universal-enveloping-is-weight},
-  \(\alpha \in Q\), a simple calculation shows
-  \(\mathcal{U}(\mathfrak{g})_\alpha \cdot N \subset L_{\lambda + \alpha}\).
-  Since \(L\) is simple and \(N\) is nonzero, it follows from
-  Example~\ref{ex:adjoint-action-in-universal-enveloping-is-weight} that
-  \[
-    L
-    = \mathcal{U}(\mathfrak{g}) \cdot N
-    = \bigoplus_{\alpha \in Q} \mathcal{U}(\mathfrak{g})_\alpha \cdot N
-  \]
-  and thus \(L_{\lambda + \alpha} = \mathcal{U}(\mathfrak{g})_\alpha \cdot N\).
-  In particular, \(L_\lambda = \mathcal{U}(\mathfrak{g})_0 \cdot N \subset N\)
-  and \(N = L_\lambda\).
-
-  Now given a semisimple weight \(\mathfrak{g}\)-module \(M = \bigoplus_i M_i\)
-  with \(M_i\) simple, it is clear \(M_\lambda = \bigoplus_i (M_i)_\lambda\).
-  Each \((M_i)_\lambda\) is either \(0\) or a simple
-  \(\mathcal{U}(\mathfrak{g})_0\)-module, so that \(M_\lambda\) is a semisimple
-  \(\mathcal{U}(\mathfrak{g})_0\)-module. In addition, to see that the
-  multiplicity of \(L\) in \(M\) coincides with the multiplicity of
-  \(L_\lambda\) in \(M_\lambda\) it suffices to show that if \((M_i)_\lambda
-  \cong (M_j)_\lambda\) are both nonzero then \(M_i \cong M_j\).
-
-  If \(I(M_i) = \mathcal{U}(\mathfrak{g}) \otimes_{\mathcal{U}(\mathfrak{g})_0}
-  (M_i)_\lambda\), the inclusion of \(\mathcal{U}(\mathfrak{g})_0\)-modules
-  \((M_i)_\lambda \to M_i\) induces a \(\mathfrak{g}\)-homomorphism
-  \begin{align*}
-            I(M_i) & \to     M_i       \\
-    u \otimes m & \mapsto u \cdot m
-  \end{align*}
-
-  Since \(M_i\) is simple and \(\lambda \in \operatorname{supp} M_i\), \(M_i =
-  \mathcal{U}(\mathfrak{g}) \cdot (M_i)_\lambda\). The homomorphism \(I(M_i)
-  \to M_i\) is thus surjective. Similarly, if \(I(M_j) =
-  \mathcal{U}(\mathfrak{g}) \otimes_{\mathcal{U}(\mathfrak{g})_0}
-  (M_j)_\lambda\) then there is a natural surjective
-  \(\mathfrak{g}\)-homomorphism \(I(M_j) \to M_j\). Now suppose there is an
-  isomorphism of \(\mathcal{U}(\mathfrak{g})_0\)-modules \(f: (M_i)_\lambda
-  \isoto (M_j)_\lambda\). Such an isomorphism induces an isomorphism of
-  \(\mathfrak{g}\)-modules
-  \begin{align*}
-    \tilde f : I(M_i) & \isoto  I(M_j)            \\
-       u \otimes m & \mapsto u \otimes f(m)
-  \end{align*}
-
-  By composing \(\tilde f\) with the projection \(I(M_j) \to M_j\) we get a
-  surjective homomorphism \(I(M_i) \to M_j\). We claim \(\ker (I(M_i) \to M_i)
-  = \ker (I(M_i) \to M_j)\).  To see this, notice that \(\ker(I(M_i) \to M_i)\)
-  coincides with the largest submodule \(Z(M_i) \subset I(M_i)\) contained in
-  \(\bigoplus_{\alpha \ne 0} \mathcal{U}(\mathfrak{g})_\alpha
-  \otimes_{\mathcal{U}(\mathfrak{g})_0} (M_i)_\lambda\). Indeed, a simple
-  computation shows \(\ker (I(M_i) \to M_i) \cap (\mathcal{U}(\mathfrak{g})_0
-  \otimes_{\mathcal{U}(\mathfrak{g})_0} (M_i)_\lambda) = 0\), which implies
-  \(\ker(I(M_i) \to M_i) \subset Z(M_i)\). Since \(M_i\) is simple, \(\ker
-  (I(M_i) \to M_i)\) is maximal and thus \(\ker(I(M_i) \to M_i) = Z(M_i)\). By
-  the same token, \(\ker (I(M_j) \to M_j)\) is the largest submodule of
-  \(I(M_j)\) contained in \(\bigoplus_{\alpha \ne 0}
-  \mathcal{U}(\mathfrak{g})_\alpha \otimes_{\mathcal{U}(\mathfrak{g})_0}
-  (M_j)_\lambda\) and therefore \(\ker(I(M_i) \to M_i) =
-  \tilde{f}^{-1}(\ker(I(M_j) \to M_j)) = \ker(I(M_i) \to M_j)\).
-
-  Hence there is an isomorphism \(\mfrac{I(M_i)}{\ker(I(M_i) \to M_i)} \isoto
-  M_j\) satisfying
-  \begin{center}
-    \begin{tikzcd}
-      I(M_i) \rar{\tilde f} \dar & I(M_j) \dar \\
-      \mfrac{I(M_i)}{\ker(I(M_i) \to M_i)} \rar{\sim} & M_j
-    \end{tikzcd}
-  \end{center}
- and finally \(M_i \cong \mfrac{I(M_i)}{\ker(I(M_i) \to M_i)} \cong M_j\).
-\end{proof}
-
-A complementary question now is: which submodules of a \emph{nice} coherent
-family are cuspidal representations?
-
-\begin{proposition}[Mathieu]
-  Let \(\mathcal{M}\) be an irreducible coherent family of degree \(d\) and
-  \(\lambda \in \mathfrak{h}^*\). The following conditions are equivalent.
-  \begin{enumerate}
-    \item \(\mathcal{M}[\lambda]\) is simple.
-    \item \(F_\alpha\!\restriction_{\mathcal{M}[\lambda]}\) is injective for
-      all \(\alpha \in \Delta\).
-    \item \(\mathcal{M}[\lambda]\) is cuspidal.
-  \end{enumerate}
-\end{proposition}
-
-\begin{proof}
-  The fact that \strong{(i)} and \strong{(iii)} are equivalent follows directly
-  from Corollary~\ref{thm:cuspidal-mod-equivs}. Likewise, it is clear from the
-  corollary that \strong{(iii)} implies \strong{(ii)}. All it is left is to
-  show \strong{(ii)} implies \strong{(iii)}. This isn't already clear from
-  Corollary~\ref{thm:cuspidal-mod-equivs} because, at first glance,
-  $\mathcal{M}[\lambda]$ may not be simple for some $\lambda$ satisfying
-  \strong{(ii)}. We will show this is never the case.
-
-  Suppose \(F_\alpha\) acts injectively on the submodule
-  \(\mathcal{M}[\lambda]\), for all \(\alpha \in \Delta\). Since
-  \(\mathcal{M}[\lambda]\) has finite length, \(\mathcal{M}[\lambda]\) contains
-  an infinite-dimensional simple \(\mathfrak{g}\)-submodule \(M\). Moreover,
-  again by Corollary~\ref{thm:cuspidal-mod-equivs} we conclude \(M\) is a
-  cuspidal module, and its degree is bounded by \(d\). We want to show
-  \(\mathcal{M}[\lambda] = M\).
-
-  We claim the set \(U = \{\mu \in \mathfrak{h}^* : \mathcal{M}_\mu \ \text{is
-  a simple $\mathcal{U}(\mathfrak{g})_0$-module}\}\) is Zariski-open. If we
-  suppose this is the case for a moment or two, it follows from the fact that
-  \(M\) is simple and \(\operatorname{supp}_{\operatorname{ess}} M\) is
-  Zariski-dense that \(U \cap \operatorname{supp}_{\operatorname{ess}} M\) is
-  non-empty. In other words, there is some \(\mu \in \mathfrak{h}^*\) such that
-  \(\mathcal{M}_\mu\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module and
-  \(\dim M_\mu = \deg M\).
-
-  In particular, \(M_\mu \ne 0\), so \(M_\mu = \mathcal{M}_\mu\). Now given any
-  simple \(\mathfrak{g}\)-module \(L\), it follows from
-  Lemma~\ref{thm:centralizer-multiplicity} that the multiplicity of \(L\)
-  in \(\mathcal{M}[\lambda]\) is the same as the multiplicity \(L_\mu\) in
-  \(\mathcal{M}_\mu\) as a \(\mathcal{U}(\mathfrak{g})_0\)-module -- which is,
-  of course, \(1\) if \(L \cong M\) and \(0\) otherwise. Hence
-  \(\mathcal{M}[\lambda] = M\) and \(\mathcal{M}[\lambda]\) is cuspidal.
-\end{proof}
-
-To finish the proof, we now show\dots
-
-\begin{lemma}
-  Let \(\mathcal{M}\) be a coherent family. The set \(U = \{\lambda \in
-  \mathfrak{h}^* : \mathcal{M}_\lambda \ \text{is a simple
-  $\mathcal{U}(\mathfrak{g})_0$-module}\}\) is Zariski-open.
-\end{lemma}
-
-\begin{proof}
-  For each \(\lambda \in \mathfrak{h}^*\) we introduce the bilinear form
-  \begin{align*}
-    B_\lambda : \mathcal{U}(\mathfrak{g})_0 \times \mathcal{U}(\mathfrak{g})_0
-    & \to K \\
-    (u, v)
-    & \mapsto \operatorname{Tr}(u v \!\restriction_{\mathcal{M}_\lambda})
-  \end{align*}
-  and consider its rank -- i.e. the dimension of the image of the induced
-  operator
-  \begin{align*}
-    \mathcal{U}(\mathfrak{g})_0 & \to     \mathcal{U}(\mathfrak{g})_0^* \\
-                              u & \mapsto B_\lambda(u, \cdot)
-  \end{align*}
-
-  Our first observation is that \(\operatorname{rank} B_\lambda \le d^2\). This
-  follows from the commutativity of
-  \begin{center}
-    \begin{tikzcd}
-      \mathcal{U}(\mathfrak{g})_0               \rar \dar  &
-      \mathcal{U}(\mathfrak{g})_0^*                        \\
-      \operatorname{End}(\mathcal{M}_\lambda)   \rar{\sim} &
-      \operatorname{End}(\mathcal{M}_\lambda)^* \uar
-    \end{tikzcd},
-  \end{center}
-  where the map \(\mathcal{U}(\mathfrak{g})_0 \to
-  \operatorname{End}(\mathcal{M}_\lambda)\) is given by the action of
-  \(\mathcal{U}(\mathfrak{g})_0\), the map
-  \(\operatorname{End}(\mathcal{M}_\lambda)^* \to
-  \mathcal{U}(\mathfrak{g})_0^*\) is its dual, and the isomorphism
-  \(\operatorname{End}(\mathcal{M}_\lambda) \isoto
-  \operatorname{End}(\mathcal{M}_\lambda)^*\) is induced by the trace form
-  \begin{align*}
-    \operatorname{End}(\mathcal{M}_\lambda) \times
-    \operatorname{End}(\mathcal{M}_\lambda) & \to K \\
-    (T, S) & \mapsto \operatorname{Tr}(T S)
-  \end{align*}
-
-  Indeed, \(\operatorname{rank} B_\lambda \le
-  \operatorname{rank}(\mathcal{U}(\mathfrak{g})_0 \to
-  \operatorname{End}(\mathcal{M}_\lambda)) \le \dim
-  \operatorname{End}(\mathcal{M}_\lambda) = d^2\). Furthermore, if
-  \(\operatorname{rank} B_\lambda = d^2\) then we must have
-  \(\operatorname{rank}(\mathcal{U}(\mathfrak{g})_0 \to
-  \operatorname{End}(\mathcal{M}_\lambda)) = d^2\) -- i.e. the map
-  \(\mathcal{U}(\mathfrak{g})_0 \to \operatorname{End}(\mathcal{M}_\lambda)\)
-  is surjective. In particular, if \(\operatorname{rank} B_\lambda = d^2\) then
-  \(\mathcal{M}_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module,
-  for if \(M \subset \mathcal{M}_\lambda\) is invariant under the action of
-  \(\mathcal{U}(\mathfrak{g})_0\) then \(M\) is invariant under any
-  \(K\)-linear operator \(\mathcal{M}_\lambda \to \mathcal{M}_\lambda\), so
-  that \(M = 0\) or \(M = \mathcal{M}_\lambda\).
-
-  On the other hand, if \(\mathcal{M}_\lambda\) is simple then by Burnside's
-  Theorem on matrix algebras the map \(\mathcal{U}(\mathfrak{g})_0 \to
-  \operatorname{End}(\mathcal{M}_\lambda)\) is surjective. Hence the
-  commutativity of the previously drawn diagram, as well as the fact that
-  \(\operatorname{rank}(\mathcal{U}(\mathfrak{g})_0 \to
-  \operatorname{End}(\mathcal{M}_\lambda)) =
-  \operatorname{rank}(\operatorname{End}(\mathcal{M}_\lambda)^* \to
-  \mathcal{U}(\mathfrak{g})_0^*)\), imply that \(\operatorname{rank} B_\lambda
-  = d^2\). This goes to show that \(U\) is precisely the set of all \(\lambda\)
-  such that \(B_\lambda\) has maximal rank \(d^2\). We now show that \(U\) is
-  Zariski-open. First, notice that
-  \[
-    U =
-    \bigcup_{\substack{V \subset \mathcal{U}(\mathfrak{g})_0 \\ \dim V = d}}
-    U_V,
-  \]
-  where \(U_V = \{\lambda \in \mathfrak{h}^* : \operatorname{rank}
-  B_\lambda\!\restriction_V = d^2 \}\). Here \(V\) ranges over all
-  \(d\)-dimensional subspaces of \(\mathcal{U}(\mathfrak{g})_0\) -- \(V\) is
-  not necessarily a \(\mathcal{U}(\mathfrak{g})_0\)-submodule.
-
-  Indeed, if \(\operatorname{rank} B_\lambda = d^2\) it follows from the
-  subjectivity of the map \(\mathcal{U}(\mathfrak{g})_0 \to
-  \operatorname{End}(\mathcal{M}_\lambda)\) that there is some \(V \subset
-  \mathcal{U}(\mathfrak{g})_0\) with \(\dim V = d\) such that the restriction
-  \(V \to \operatorname{End}(\mathcal{M}_\lambda)\) is surjective. The
-  commutativity of
-  \begin{center}
-    \begin{tikzcd}
-      V                                         \rar \dar  & V^* \\
-      \operatorname{End}(\mathcal{M}_\lambda)   \rar{\sim} &
-      \operatorname{End}(\mathcal{M}_\lambda)^* \uar
-    \end{tikzcd}
-  \end{center}
-  then implies \(\operatorname{rank} B_\lambda\!\restriction_V = d^2\). In
-  other words, \(U \subset \bigcup_V U_V\).
-
-  Likewise, if \(\operatorname{rank} B_\lambda\!\restriction_V = d^2\) for some
-  \(V\), then the commutativity of
-  \begin{center}
-    \begin{tikzcd}
-      V                             \rar \dar & V^* \\
-      \mathcal{U}(\mathfrak{g})_0   \rar      &
-      \mathcal{U}(\mathfrak{g})_0^* \uar
-    \end{tikzcd}
-  \end{center}
-  implies \(\operatorname{rank} B_\lambda \ge d^2\), which goes to show
-  \(\bigcup_V U_V \subset U\).
-
-  Given \(\lambda \in U_V\), the surjectivity of \(V \to
-  \operatorname{End}(\mathcal{M}_\lambda)\) and the fact that \(\dim V <
-  \infty\) imply \(V \to V^*\) is invertible. Since \(\mathcal{M}\) is a
-  coherent family, \(B_\lambda\) depends polynomially in \(\lambda\). Hence so
-  does the induced maps \(V \to V^*\). In particular, there is some Zariski
-  neighborhood \(U'\) of \(\lambda\) such that the map \(V \to V^*\) induced by
-  \(B_\mu\!\restriction_V\) is invertible for all \(\mu \in U'\).
-
-  But the surjectivity of the map induced by \(B_\mu\!\restriction_V\) implies
-  \(\operatorname{rank} B_\mu = d^2\), so \(\mu \in U_V\) and therefore \(U'
-  \subset U_V\). This implies \(U_V\) is open for all \(V\). Finally, \(U\) is
-  the union of Zariski-open subsets and is therefore open. We are done.
-\end{proof}
-
-The major remaining question for us to tackle is that of the existence of
-coherent extensions, which will be the focus of our next section.
-
-\section{Localizations \& the Existence of Coherent Extensions}
-
-Let \(M\) be a simple bounded \(\mathfrak{g}\)-module of degree \(d\). Our
-goal is to prove that \(M\) has a (unique) irreducible semisimple coherent
-extension \(\mathcal{M}\). Since \(M\) is simple, we know \(M \subset
-\mathcal{M}[\lambda]\) for any \(\lambda \in \operatorname{supp} M\). Our first
-task is constructing \(\mathcal{M}[\lambda]\). The issue here is that
-\(\operatorname{supp}_{\operatorname{ess}} M\) may not be all of \(\lambda + Q
-= \operatorname{supp}_{\operatorname{ess}} \mathcal{M}[\lambda]\), so we may
-find \(M \subsetneq \mathcal{M}[\lambda]\). In fact, we may find
-\(\operatorname{supp} M \subsetneq \lambda + Q\).
-
-This wasn't an issue an Example~\ref{ex:laurent-polynomial-mod} because we
-verified that the action of \(f \in \mathfrak{sl}_2(K)\) on \(K[x, x^{-1}]\) is
-injective. Since all weight spaces of \(K[x, x^{-1}]\) are \(1\)-dimensional,
-this implies the action of \(f\) is actually bijective, so we can obtain a
-nonzero vector in \(K[x, x^{-1}]_{2 k} = K x^k\) for any \(k \in \mathbb{Z}\)
-by translating between weight spaced using \(f\) and \(f^{-1}\) -- here
-\(f^{-1}\) denotes the \(K\)-linear operator \((-
-\sfrac{\mathrm{d}}{\mathrm{d}x} + \sfrac{x^{-1}}{2})^{-1}\), which is the
-inverse of the action of \(f\) on \(K[x, x^{-1}]\).
-\begin{center}
-  \begin{tikzcd}
-    \cdots     \rar[bend left=60]{f^{-1}}
-    & K x^{-2} \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
-    & K x^{-1} \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
-    & K        \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
-    & K x      \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
-    & K x^2    \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
-    & \cdots   \lar[bend left=60]{f}
-  \end{tikzcd}
-\end{center}
-
-In the general case, the action of some \(F_\alpha \in \mathfrak{g}\) with
-\(\alpha \in \Delta\) in \(M\) may not be injective. In fact, we have seen that
-the action of \(F_\alpha\) is injective for all \(\alpha \in \Delta^+\) if, and
-only if \(M\) is cuspidal. Nevertheless, we could intuitively \emph{make it
-injective} by formally inverting the elements \(F_\alpha \in
-\mathcal{U}(\mathfrak{g})\). This would allow us to obtain nonzero vectors in
-\(M_\mu\) for all \(\mu \in \lambda + Q\) by successively applying elements of
-\(\{F_\alpha^{\pm 1}\}_{\alpha \in \Delta}\) to a nonzero weight vector \(m \in
-M_\lambda\). Moreover, if the actions of the \(F_\alpha\) were to be
-invertible, we would find that all \(M_\mu\) are \(d\)-dimensional for \(\mu
-\in \lambda + Q\).
-
-In a commutative domain, this can be achieved by tensoring our module by the
-field of fractions. However, \(\mathcal{U}(\mathfrak{g})\) is hardly ever
-commutative -- \(\mathcal{U}(\mathfrak{g})\) is commutative if, and only if
-\(\mathfrak{g}\) is Abelian -- and the situation is more delicate in the
-non-commutative case. For starters, a non-commutative \(K\)-algebra \(A\) may
-not even have a ``field of fractions'' -- i.e. an over-ring where all elements
-of \(A\) have inverses. Nevertheless, it is possible to formally invert
-elements of certain subsets of \(A\) via a process known as
-\emph{localization}, which we now describe.
-
-\begin{definition}\index{localization!multiplicative subsets}\index{localization!Ore's condition}
-  Let \(A\) be a \(K\)-algebra. A subset \(S \subset A\) is called
-  \emph{multiplicative} if \(s \cdot t \in S\) for all \(s, t \in S\) and \(0
-  \notin S\). A multiplicative subset \(S\) is said to satisfy \emph{Ore's
-  localization condition} if for each \(a \in A\) and \(s \in S\) there exists
-  \(b, c \in A\) and \(t, t' \in S\) such that \(s a = b t\) and \(a s = t'
-  c\).
-\end{definition}
-
-\begin{theorem}[Ore-Asano]\index{localization!Ore-Asano Theorem}
-  Let \(S \subset A\) be a multiplicative subset satisfying Ore's localization
-  condition. Then there exists a (unique) \(K\)-algebra \(S^{-1} A\), with a
-  canonical algebra homomorphism \(A \to S^{-1} A\), enjoying the universal
-  property that each algebra homomorphism \(f : A \to B\) such that \(f(s)\) is
-  invertible for all \(s \in S\) can be uniquely extended to an algebra
-  homomorphism \(S^{-1} A \to B\). \(S^{-1} A\) is called \emph{the
-  localization of \(A\) by \(S\)}, and the map \(A \to S^{-1} A\) is called
-  \emph{the localization map}.
-  \begin{center}
-    \begin{tikzcd}
-      A        \dar \rar{f}        & B \\
-      S^{-1} A \urar[swap, dotted] &
-    \end{tikzcd}
-  \end{center}
-\end{theorem}
-
-If we identify an element with its image under the localization map, it follows
-directly from Ore's construction that every element of \(S^{-1} A\) has the
-form \(s^{-1} a\) for some \(s \in S\) and \(a \in A\). Likewise, any element
-of \(S^{-1} A\) can also be written as \(b t^{-1}\) for some \(t \in S\), \(b
-\in A\).
-
-Ore's localization condition may seem a bit arbitrary at first, but a more
-thorough investigation reveals the intuition behind it. The issue in question
-here is that in the non-commutative case we can no longer take the existence of
-common denominators for granted. However, the existence of common denominators
-is fundamental to the proof of the fact the field of fractions is a ring -- it
-is used, for example, to define the sum of two elements in the field of
-fractions. We thus need to impose their existence for us to have any hope of
-defining consistent arithmetics in the localization of an algebra, and Ore's
-condition is actually equivalent to the existence of common denominators --
-see the discussion in the introduction of \cite[ch.~6]{goodearl-warfield} for
-further details.
-
-We should also point out that there are numerous other conditions -- which may
-be easier to check than Ore's -- known to imply Ore's condition. For
-instance\dots
-
-\begin{lemma}
-  Let \(S \subset A\) be a multiplicative subset generated by finitely many
-  locally \(\operatorname{ad}\)-nilpotent elements -- i.e. elements \(s \in S\)
-  such that for each \(a \in A\) there exists \(r > 0\) such that
-  \(\operatorname{ad}(s)^r a = [s, [s, \cdots [s, a]]\cdots] = 0\). Then \(S\)
-  satisfies Ore's localization condition.
-\end{lemma}
-
-In our case, we are more interested in formally inverting the action of
-\(F_\alpha\) on \(M\) than in inverting \(F_\alpha\) itself. To that end, we
-introduce one further construction, known as \emph{the localization of a
-module}.
-
-\begin{definition}\index{localization!localization of modules}
-  Let \(S \subset A\) be a multiplicative subset satisfying Ore's localization
-  condition and \(M\) be an \(A\)-module. The \(S^{-1} A\)-module \(S^{-1} M =
-  S^{-1} A \otimes_A M\) is called \emph{the localization of \(M\) by \(S\)},
-  and the homomorphism of \(A\)-modules
-  \begin{align*}
-    M & \to     S^{-1} M    \\
-    m & \mapsto 1 \otimes m
-  \end{align*}
-  is called \emph{the localization map of \(M\)}.
-\end{definition}
-
-Notice that the \(S^{-1} A\)-module \(S^{-1} M\) has the natural structure of
-an \(A\)-module, where the action of \(A\) is given by the localization map \(A
-\to S^{-1} A\).
-
-It is interesting to observe that, unlike in the case of the field of fractions
-of a commutative domain, in general the localization map \(A \to S^{-1} A\) --
-i.e. the map \(a \mapsto \frac{a}{1}\) -- may not be injective. For instance,
-if \(S\) contains a divisor of zero \(s\), its image under the localization map
-is invertible and therefore cannot be a divisor of zero in \(S^{-1} A\). In
-particular, if \(a \in A\) is nonzero and such that \(s a = 0\) or \(a s = 0\)
-then its image under the localization map has to be \(0\). However, the
-existence of divisors of zero in \(S\) turns out to be the only obstruction to
-the injectivity of the localization map, as shown in\dots
-
-\begin{lemma}
-  Let \(S \subset A\) be a multiplicative subset satisfying Ore's localization
-  condition and \(M\) be an \(A\)-module. If \(S\) acts injectively on \(M\)
-  then the localization map \(M \to S^{-1} M\) is injective. In particular, if
-  \(S\) has no zero divisors then \(A\) is a subalgebra of \(S^{-1} A\).
-\end{lemma}
-
-Again, in our case we are interested in inverting the actions of the
-\(F_\alpha\) on \(M\). However, for us to be able to translate between all
-weight spaces associated with elements of \(\lambda + Q\), \(\lambda \in
-\operatorname{supp} M\), we only need to invert the \(F_\alpha\)'s for
-\(\alpha\) in some subset of \(\Delta\) which spans all of \(Q = \mathbb{Z}
-\Delta\). In other words, it suffices to invert \(F_\beta\) for all \(\beta\)
-in some basis \(\Sigma\) for \(\Delta\). We can choose such a basis to be
-well-behaved. For example, we can show\dots
-
-\begin{lemma}\label{thm:nice-basis-for-inversion}
-  Let \(M\) be a simple infinite-dimensional bounded \(\mathfrak{g}\)-module.
-  There is a basis \(\Sigma = \{\beta_1, \ldots, \beta_r\}\) for \(\Delta\)
-  such that the elements \(F_{\beta_i}\) all act injectively on \(M\) and
-  satisfy \([F_{\beta_i}, F_{\beta_j}] = 0\).
-\end{lemma}
-
-\begin{note}
-  The basis \(\Sigma\) in Lemma~\ref{thm:nice-basis-for-inversion} may very
-  well depend on the representation \(M\)! This is another obstruction to the
-  functoriality of our constructions.
-\end{note}
-
-The proof of the previous Lemma is quite technical and was deemed too tedious
-to be included in here. See Lemma 4.4 of \cite{mathieu} for a full proof. Since
-\(F_\alpha\) is locally \(\operatorname{ad}\)-nilpotent for all \(\alpha \in
-\Delta\), we can see\dots
-
-\begin{corollary}
-  Let \(\Sigma\) be as in Lemma~\ref{thm:nice-basis-for-inversion} and
-  \((F_\beta)_{\beta \in \Sigma} \subset \mathcal{U}(\mathfrak{g})\) be the
-  multiplicative subset generated by the \(F_\beta\)'s. The \(K\)-algebra
-  \(\Sigma^{-1} \mathcal{U}(\mathfrak{g}) = (F_\beta)_{\beta \in \Sigma}^{-1}
-  \mathcal{U}(\mathfrak{g})\) is well defined. Moreover, if we denote by
-  \(\Sigma^{-1} M\) the localization of \(M\) by \((F_\beta)_{\beta \in
-  \Sigma}\), the localization map \(M \to \Sigma^{-1} M\) is injective.
-\end{corollary}
-
-From now on let \(\Sigma\) be some fixed basis for \(\Delta\) satisfying the
-hypothesis of Lemma~\ref{thm:nice-basis-for-inversion}. We now show that
-\(\Sigma^{-1} M\) is a weight \(\mathfrak{g}\)-module whose support is an
-entire \(Q\)-coset.
-
-\begin{proposition}\label{thm:irr-bounded-is-contained-in-nice-mod}
-  The restriction of the localization \(\Sigma^{-1} M\) is a bounded
-  \(\mathfrak{g}\)-module of degree \(d\) with \(\operatorname{supp}
-  \Sigma^{-1} M = Q + \operatorname{supp} M\) and \(\dim \Sigma^{-1} M_\lambda
-  = d\) for all \(\lambda \in \operatorname{supp} \Sigma^{-1} M\).
-\end{proposition}
-
-\begin{proof}
-  Fix some \(\beta \in \Sigma\). We begin by showing that \(F_\beta\) and
-  \(F_\beta^{-1}\) map the weight space \(\Sigma^{-1} M_\lambda\) to
-  \(\Sigma^{-1} M_{\lambda - \beta}\) and \(\Sigma^{-1} M_{\lambda + \beta}\),
-  respectively. Indeed, given \(m \in M_\lambda\) and \(H \in \mathfrak{h}\) we
-  have
-  \[
-    H \cdot (F_\beta \cdot m)
-    = ([H, F_\beta] + F_\beta H) \cdot m
-    = F_\beta (-\beta(H) + H) \cdot m
-    = (\lambda - \beta)(H) F_\beta \cdot m
-  \]
-
-  On the other hand,
-  \[
-    0
-    = [H, 1]
-    = [H, F_\beta F_\beta^{-1}]
-    = F_\beta [H, F_\beta^{-1}] + [H, F_\beta] F_\beta^{-1}
-    = F_\beta [H, F_\beta^{-1}] - \beta(H) F_\beta F_\beta^{-1},
-  \]
-  so that \([H, F_\beta^{-1}] = \beta(H) \cdot F_\beta^{-1}\) and therefore
-  \[
-    H \cdot (F_\beta^{-1} \cdot m)
-    = ([H, F_\beta^{-1}] + F_\beta^{-1} H) \cdot m
-    = F_\beta^{-1} (\beta(H) + H) \cdot m
-    = (\lambda + \beta)(H) F_\beta^{-1} \cdot m
-  \]
-
-  From the fact that \(F_\beta^{\pm 1}\) maps \(M_\lambda\) to \(\Sigma^{-1}
-  M_{\lambda \pm \beta}\) follows our first conclusion: since \(M\) is a weight
-  module and every element of \(\Sigma^{-1} M\) has the form \(s^{-1} \cdot m =
-  s^{-1} \otimes m\) for \(s \in (F_\beta)_{\beta \in \Sigma}\) and \(m \in
-  M\), we can see that \(\Sigma^{-1} M = \bigoplus_\lambda \Sigma^{-1}
-  M_\lambda\). Furthermore, since the action of each \(F_\beta\) on
-  \(\Sigma^{-1} M\) is bijective and \(\Sigma\) is a basis for \(Q\) we obtain
-  \(\operatorname{supp} \Sigma^{-1} M = Q + \operatorname{supp} M\).
-
-  Again, because of the bijectivity of the \(F_\beta\)'s, to see that \(\dim
-  \Sigma^{-1} M_\lambda = d\) for all \(\lambda \in \operatorname{supp}
-  \Sigma^{-1} M\) it suffices to show that \(\dim \Sigma^{-1} M_\lambda = d\)
-  for some \(\lambda \in \operatorname{supp} \Sigma^{-1} M\). We may take
-  \(\lambda \in \operatorname{supp} M\) with \(\dim M_\lambda = d\). For any
-  finite-dimensional subspace \(V \subset \Sigma^{-1} M_\lambda\) we can find
-  \(s \in (F_\beta)_{\beta \in \Sigma}\) such that \(s \cdot V \subset M\). If
-  \(s = F_{\beta_{i_1}} \cdots F_{\beta_{i_r}}\), it is clear \(s \cdot V
-  \subset M_{\lambda - \beta_{i_1} - \cdots - \beta_{i_r}}\), so \(\dim V =
-  \dim s \cdot V \le d\). This holds for all finite-dimensional \(V \subset
-  \Sigma^{-1} M_\lambda\), so \(\dim \Sigma^{-1} M_\lambda \le d\). It then
-  follows from the fact that \(M_\lambda \subset \Sigma^{-1} M_\lambda\) that
-  \(M_\lambda = \Sigma^{-1} M_\lambda\) and therefore \(\dim \Sigma^{-1}
-  M_\lambda = d\).
-\end{proof}
-
-We now have a good candidate for a coherent extension of \(M\), but
-\(\Sigma^{-1} M\) is still not a coherent extension since its support is
-contained in a single \(Q\)-coset. In particular, \(\operatorname{supp}
-\Sigma^{-1} M \ne \mathfrak{h}^*\) and \(\Sigma^{-1} M\) is not a coherent
-family. To obtain a coherent family we thus need somehow extend \(\Sigma^{-1}
-M\). To that end, we will attempt to replicate the construction of the coherent
-extension of the \(\mathfrak{sl}_2(K)\)-module \(K[x, x^{-1}]\). Specifically,
-the idea is that if twist \(\Sigma^{-1} M\) by an automorphism which shifts its
-support by some \(\lambda \in \mathfrak{h}^*\), we can construct a coherent
-family by summing these modules over \(\lambda\) as in
-Example~\ref{ex:sl-laurent-family}.
-
-For \(K[x, x^{-1}]\) this was achieved by twisting the
-\(\operatorname{Diff}(K[x, x^{-1}])\)-module \(K[x, x^{-1}]\) by the
-automorphisms \(\varphi_\lambda : \operatorname{Diff}(K[x, x^{-1}]) \to
-\operatorname{Diff}(K[x, x^{-1}])\) and restricting the results to
-\(\mathcal{U}(\mathfrak{sl}_2(K))\) via the map
-\(\mathcal{U}(\mathfrak{sl}_2(K)) \to \operatorname{Diff}(K[x, x^{-1}])\), but
-this approach is inflexible since not every \(\mathfrak{sl}_2(K)\)-module
-factors through \(\operatorname{Diff}(K[x, x^{-1}])\). Nevertheless, we could
-just as well twist \(K[x, x^{-1}]\) by automorphisms of
-\(\mathcal{U}(\mathfrak{sl}_2(K))_f\) directly -- where
-\(\mathcal{U}(\mathfrak{sl}_2(K))_f = (f)^{-1} \mathcal{U}(\mathfrak{g})\) is
-the localization of \(\mathcal{U}(\mathfrak{sl}_2(K))\) by the multiplicative
-subset generated by \(f\).
-
-In general, we may twist the \(\Sigma^{-1} \mathcal{U}(\mathfrak{g})\)-module
-\(\Sigma^{-1} M\) by automorphisms of \(\Sigma^{-1}
-\mathcal{U}(\mathfrak{g})\). For \(\lambda = \beta \in \Sigma\) the map
-\begin{align*}
-  \theta_\beta : \Sigma^{-1} \mathcal{U}(\mathfrak{g}) & \to
-                 \Sigma^{-1} \mathcal{U}(\mathfrak{g}) \\
-                 u & \mapsto F_\beta u F_\beta^{-1}
-\end{align*}
-is a natural candidate for such a twisting automorphism. Indeed, we will soon
-see that \(\twisted{(\Sigma^{-1} M)}{\theta_\beta}_\lambda = \Sigma^{-1}
-M_{\lambda + \beta}\). However, this is hardly useful to us, since \(\beta \in
-Q\) and therefore \(\beta + \operatorname{supp} \Sigma^{-1} M =
-\operatorname{supp} \Sigma^{-1} M\). If we want to expand the support of
-\(\Sigma^{-1} M\) we will have to twist by automorphisms that shift its support
-by \(\lambda \in \mathfrak{h}^*\) lying \emph{outside} of \(Q\).
-
-The situation is much less obvious in this case. Nevertheless, it turns out we
-can extend the family \(\{\theta_\beta\}_{\beta \in \Sigma}\) to a family of
-automorphisms \(\{\theta_\lambda\}_{\lambda \in \mathfrak{h}^*}\).
-Explicitly\dots
-
-\begin{proposition}\label{thm:nice-automorphisms-exist}
-  There is a family of automorphisms \(\{\theta_\lambda : \Sigma^{-1}
-  \mathcal{U}(\mathfrak{g}) \to \Sigma^{-1}
-  \mathcal{U}(\mathfrak{g})\}_{\lambda \in \mathfrak{h}^*}\) such that
-  \begin{enumerate}
-    \item \(\theta_{k_1 \beta_1 + \cdots + k_r \beta_r}(u) = F_{\beta_1}^{k_1}
-      \cdots F_{\beta_r}^{k_r} u F_{\beta_r}^{- k_r} \cdots F_{\beta_1}^{-
-      k_1}\) for all \(u \in \Sigma^{-1} \mathcal{U}(\mathfrak{g})\) and \(k_1,
-      \ldots, k_r \in \mathbb{Z}\).
-
-    \item For each \(u \in \Sigma^{-1} \mathcal{U}(\mathfrak{g})\) the map
-      \begin{align*}
-        \mathfrak{h}^* & \to     \Sigma^{-1} \mathcal{U}(\mathfrak{g}) \\
-               \lambda & \mapsto \theta_\lambda(u)
-      \end{align*}
-      is polynomial.
-
-    \item If \(\lambda, \mu \in \mathfrak{h}^*\), \(N\) is a \(\Sigma^{-1}
-      \mathcal{U}(\mathfrak{g})\)-module whose restriction to
-      \(\mathcal{U}(\mathfrak{g})\) is a weight \(\mathfrak{g}\)-module and
-      \(\twisted{N}{\theta_\lambda}\) is the \(\Sigma^{-1}
-      \mathcal{U}(\mathfrak{g})\)-module \(N\) twisted by the automorphism
-      \(\theta_\lambda\) then \(N_\mu = \twisted{N}{\theta_\lambda}_{\mu +
-      \lambda}\). In particular, \(\operatorname{supp}
-      \twisted{N}{\theta_\lambda} = \lambda + \operatorname{supp} N\).
-  \end{enumerate}
-\end{proposition}
-
-\begin{proof}
-  Since the elements \(F_\beta\), \(\beta \in \Sigma\) commute with one
-  another, the endomorphisms
-  \begin{align*}
-    \theta_{k_1 \beta_1 + \cdots + k_r \beta_r}
-    : \Sigma^{-1} \mathcal{U}(\mathfrak{g}) &
-    \to \Sigma^{-1} \mathcal{U}(\mathfrak{g}) \\
-    u & \mapsto
-    F_{\beta_1}^{k_1} \cdots F_{\beta_r}^{k_r}
-    u
-    F_{\beta_1}^{- k_r} \cdots F_{\beta_r}^{- k_1}
-  \end{align*}
-  are well defined for all \(k_1, \ldots, k_r \in \mathbb{Z}\).
-
-  Fix some \(u \in \Sigma^{-1} \mathcal{U}(\mathfrak{g})\). For any \(s \in
-  (F_\beta)_{\beta \in \Sigma}\) and \(k > 0\) we have \(s^k u = \binom{k}{0}
-  \operatorname{ad}(s)^0 u s^{k - 0} + \cdots + \binom{k}{k}
-  \operatorname{ad}(s)^k u s^{k - k}\). Now if we take \(\ell\) such
-  \(\operatorname{ad}(F_\beta)^{\ell + 1} u = 0\) for all \(\beta \in \Sigma\)
-  we find
-  \[
-    \theta_{k_1 \beta_1 + \cdots + k_r \beta_r}(u)
-    = \sum_{i_1, \ldots, i_r = 1, \ldots, \ell}
-    \binom{k_1}{i_1} \cdots \binom{k_r}{i_r}
-    \operatorname{ad}(F_{\beta_1})^{i_1} \cdots
-    \operatorname{ad}(F_{\beta_r})^{i_r}
-    u
-    F_{\beta_1}^{- i_1} \cdots F_{\beta_r}^{- i_r}
-  \]
-  for all \(k_1, \ldots, k_r \in \mathbb{N}\).
-
-  Since the binomial coefficients \(\binom{x}{k} = \frac{x (x-1) \cdots (x - k
-  + 1)}{k!}\) can be uniquely extended to polynomial functions in \(x \in K\),
-  we may in general define
-  \[
-    \theta_\lambda(u)
-    = \sum_{i_1, \ldots, i_r \ge 0}
-    \binom{\lambda_1}{i_1} \cdots \binom{\lambda_r}{i_r}
-    \operatorname{ad}(F_{\beta_1})^{i_1} \cdots
-    \operatorname{ad}(F_{\beta_r})^{i_r}
-    r
-    F_{\beta_1}^{- i_1} \cdots F_{\beta_r}^{- i_r}
-  \]
-  for \(\lambda_1, \ldots, \lambda_r \in K\), \(\lambda = \lambda_1 \beta_1 +
-  \cdots + \lambda_r \beta_r \in \mathfrak{h}^*\).
-
-  It is clear that the \(\theta_\lambda\) are endomorphisms. To see that the
-  \(\theta_\lambda\) are indeed automorphisms, notice \(\theta_{- k_1 \beta_1 -
-  \cdots - k_r \beta_r} = \theta_{k_1 \beta_1 + \cdots + k_r \beta_r}^{-1}\).
-  The uniqueness of the polynomial extensions then implies \(\theta_{- \lambda}
-  = \theta_\lambda^{-1}\) in general: given \(u \in \Sigma^{-1}
-  \mathcal{U}(\mathfrak{g})\), the map
-  \begin{align*}
-    \mathfrak{h}^* & \to \Sigma^{-1} \mathcal{U}(\mathfrak{g})        \\
-           \lambda & \mapsto \theta_\lambda(\theta_{-\lambda}(u)) - u
-  \end{align*}
-  is a polynomial extension of the zero map \(\mathbb{Z} \beta_1 \oplus \cdots
-  \oplus \mathbb{Z} \beta_r \to \Sigma^{-1} \mathcal{U}(\mathfrak{g})\) and is
-  therefore identically zero.
-
-  Finally, let \(N\) be a \(\Sigma^{-1} \mathcal{U}(\mathfrak{g})\)-module
-  whose restriction is a weight module. If \(n \in N\) then
-  \[
-    n \in \twisted{N}{\theta_\lambda}_{\mu + \lambda}
-    \iff \theta_\lambda(H) \cdot n = (\mu + \lambda)(H) n
-    \, \forall H \in \mathfrak{h}
-  \]
-
-  But
-  \[
-    \theta_\beta(H)
-    = F_\beta H F_\beta^{-1}
-    = ([F_\beta, H] + H F_\beta) F_\beta^{-1}
-    = (\beta(H) + H) F_\beta F_\beta^{-1}
-    = \beta(H) + H
-  \]
-  for all \(H \in \mathfrak{h}\) and \(\beta \in \Sigma\). In general,
-  \(\theta_\lambda(H) = \lambda(H) + H\) for all \(\lambda \in \mathfrak{h}^*\)
-  and hence
-  \[
-    \begin{split}
-      n \in \twisted{N}{\theta_\lambda}_{\mu + \lambda}
-      & \iff (\lambda(H) + H) \cdot n = (\mu + \lambda)(H) n
-        \; \forall H \in \mathfrak{h} \\
-      & \iff H \cdot n = \mu(H) n \; \forall H \in \mathfrak{h} \\
-      & \iff n \in N_\mu
-    \end{split},
-  \]
-  so that \(\twisted{N}{\theta_\lambda}_{\mu + \lambda} = N_\mu\).
-\end{proof}
-
-It should now be obvious\dots
-
-\begin{proposition}[Mathieu]
-  There exists a coherent extension \(\mathcal{M}\) of \(M\).
-\end{proposition}
-
-\begin{proof}
-  Take\footnote{Here we fix some $\lambda_\xi \in \xi$ for each $Q$-coset $\xi
-  \in \mfrac{\mathfrak{h}^*}{Q}$. While there is a natural isomorphism
-  $\twisted{(\Sigma^{-1} M)}{\theta_\lambda} \isoto \twisted{(\Sigma^{-1}
-  M)}{\theta_\mu}$ for each $\mu \in \lambda + Q$, they are not the same
-  \(\mathfrak{g}\)-modules strictly speaking. This is yet another obstruction
-  to the functoriality of our constructions.}
-  \[
-    \mathcal{M}
-    = \bigoplus_{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q}}
-      \twisted{(\Sigma^{-1} M)}{\theta_\lambda}
-  \]
-
-  It is clear \(M\) lies in \(\Sigma^{-1} M = \twisted{(\Sigma^{-1}
-  M)}{\theta_0}\) and therefore \(M \subset \mathcal{M}\). On the other hand,
-  \(\dim \mathcal{M}_\mu = \dim \twisted{(\Sigma^{-1} M)}{\theta_\lambda}_\mu =
-  \dim \Sigma^{-1} M_{\mu - \lambda} = d\) for all \(\mu \in \lambda + Q\) --
-  \(\lambda\) standing for some fixed representative of its \(Q\)-coset.
-  Furthermore, given \(u \in \mathcal{U}(\mathfrak{g})_0\) and \(\mu \in
-  \lambda + Q\),
-  \[
-    \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\mu})
-    = \operatorname{Tr}
-      (\theta_\lambda(u)\!\restriction_{\Sigma^{-1} M_{\mu - \lambda}})
-  \]
-  is polynomial in \(\mu\) because of the second item of
-  Proposition~\ref{thm:nice-automorphisms-exist}.
-\end{proof}
-
-Lo and behold\dots
-
-\begin{theorem}[Mathieu]\index{coherent family!Mathieu's \(\mExt\) coherent extension}
-  There exists a unique semisimple coherent extension \(\mExt(M)\) of \(M\).
-  More precisely, if \(\mathcal{M}\) is any coherent extension of \(M\), then
-  \(\mathcal{M}^{\operatorname{ss}} \cong \mExt(M)\). Furthermore, \(\mExt(M)\)
-  is a irreducible coherent family.
-\end{theorem}
-
-\begin{proof}
-  The existence part should be clear from the previous discussion: it suffices
-  to fix some coherent extension \(\mathcal{M}\) of \(M\) and take
-  \(\mExt(M) = \mathcal{M}^{\operatorname{ss}}\).
-
-  To see that \(\mExt(M)\) is irreducible, recall from
-  Corollary~\ref{thm:bounded-is-submod-of-extension} that \(M\) is a
-  \(\mathfrak{g}\)-submodule of \(\mExt(M)\). Since the degree of \(M\) is the
-  same as the degree of \(\mExt(M)\), some of its weight spaces have maximal
-  dimension inside of \(\mExt(M)\). In particular, it follows from
-  Lemma~\ref{thm:centralizer-multiplicity} that \(\mExt(M)_\lambda =
-  M_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module for some
-  \(\lambda \in \operatorname{supp} M\).
-
-  As for the uniqueness of \(\mExt(M)\), fix some other semisimple coherent
-  extension \(\mathcal{N}\) of \(M\). We claim that the multiplicity of a given
-  simple \(\mathfrak{g}\)-module \(L\) in \(\mathcal{N}\) is determined by its
-  \emph{trace function}
-  \begin{align*}
-    \mathfrak{h}^* \times \mathcal{U}(\mathfrak{g})_0 &
-    \to K \\
-    (\lambda, u) &
-    \mapsto \operatorname{Tr}(u\!\restriction_{\mathcal{N}_\lambda})
-  \end{align*}
-
-  It is a well known fact of the theory of modules that, given an associative
-  \(K\)-algebra \(A\), a finite-dimensional semisimple \(A\)-module \(L\) is
-  determined, up to isomorphism, by its \emph{character}
-  \begin{align*}
-    \chi_L : A & \to     K                                    \\
-             a & \mapsto \operatorname{Tr}(a\!\restriction_L)
-  \end{align*}
-
-  In particular, the multiplicity of \(L\) in \(\mathcal{N}\), which is the
-  same as the multiplicity of \(L_\lambda\) in \(\mathcal{N}_\lambda\), is
-  determined by the character \(\chi_{\mathcal{N}_\lambda} :
-  \mathcal{U}(\mathfrak{g})_0 \to K\). Since this holds for all simple weight
-  \(\mathfrak{g}\)-modules, it follows that \(\mathcal{N}\) is determined by
-  its trace function. Of course, the same holds for \(\mExt(M)\). We now claim
-  that the trace function of \(\mathcal{N}\) is the same as that of
-  \(\mExt(M)\). Clearly,
-  \(\operatorname{Tr}(u\!\restriction_{\mExt(M)_\lambda}) =
-  \operatorname{Tr}(u\!\restriction_{M_\lambda}) =
-  \operatorname{Tr}(u\!\restriction_{\mathcal{N}_\lambda})\) for all \(\lambda
-  \in \operatorname{supp}_{\operatorname{ess}} M\), \(u \in
-  \mathcal{U}(\mathfrak{g})_0\). Since the essential support of \(M\) is
-  Zariski-dense and the maps \(\lambda \mapsto
-  \operatorname{Tr}(u\!\restriction_{\mExt(M)_\lambda})\) and \(\lambda \mapsto
-  \operatorname{Tr}(u\!\restriction_{\mathcal{N}_\lambda})\) are polynomial in
-  \(\lambda \in \mathfrak{h}^*\), it follows that these maps coincide for all
-  \(u\).
-
-  In conclusion, \(\mathcal{N} \cong \mExt(M)\) and \(\mExt(M)\) is unique.
-\end{proof}
-
-% This is a very important theorem, but since we won't classify the coherent
-% extensions in here we don't need it, and there is no other motivation behind
-% it. Including this would also require me to explain what central characters
-% are, which is a bit of a pain
-%\begin{proposition}[Mathieu]
-%  The central characters of the irreducible submodules of
-%  \(\operatorname{Ext}(M)\) are all the same.
-%\end{proposition}
-
-We have thus concluded our classification of cuspidal modules in terms of
-coherent families. Of course, to get an explicit construction of all simple
-\(\mathfrak{g}\)-modules we would have to classify the irreducible semisimple
-coherent \(\mathfrak{g}\)-families themselves, which is the subject of sections
-7, 8 and 9 of \cite{mathieu}. In addition, in sections 11 and 12 of
-\cite{mathieu} Mathieu provides an explicit construction of coherent families.
-We unfortunately do not have the necessary space to discuss these results in
-detail, but we will now provide a brief overview.
-
-First and foremost, notice that because of
-Example~\ref{thm:simple-weight-mod-is-tensor-prod} the problem of classifying
-the simple weight \(\mathfrak{g}\)-modules can be reduced to that of
-classifying the simple weight modules of its simple components. In addition, it
-turns out that very few simple Lie algebras admit cuspidal modules at all.
-Specifically\dots
-
-\begin{proposition}[Fernando]\label{thm:only-sl-n-sp-have-cuspidal}
-  Let \(\mathfrak{s}\) be a finite-dimensional simple Lie algebra. Suppose
-  there exists a simple cuspidal \(\mathfrak{s}\)-module. Then \(\mathfrak{s}
-  \cong \mathfrak{sl}_n(K)\) or \(\mathfrak{s} \cong \mathfrak{sp}_{2 n}(K)\).
-\end{proposition}
-
-Hence it suffices to classify the irreducible semisimple coherent families of
-\(\mathfrak{sl}_n(K)\) and \(\mathfrak{sp}_{2 n}(K)\). These can be described
-either algebraically, using combinatorial invariants -- which Mathieu does in
-sections 7, 8 and 9 of his paper -- or geometrically, using algebraic varieties
-and differential forms -- which is done in sections 11 and 12. While rather
-complicated on its own, the geometric construction is more concrete than its
-combinatorial counterpart.
-
-This construction also brings us full circle to the beginning of these notes,
-where we saw in Proposition~\ref{thm:geometric-realization-of-uni-env} that
-\(\mathfrak{g}\)-modules may be understood as geometric objects. In fact,
-throughout the previous four chapters we have seen a tremendous number of
-geometrically motivated examples, which further emphasizes the connection
-between representation theory and geometry. I would personally go as far as
-saying that the beautiful interplay between the algebraic and the geometric is
-precisely what makes representation theory such a fascinating and charming
-subject.
-
-Alas, our journey has come to an end. All it is left is to wonder at the beauty
-of Lie algebras and their representations.
-
-\label{end-47}
diff --git a/sections/semisimple-algebras.tex /dev/null
@@ -1,1044 +0,0 @@
-\chapter{Finite-Dimensional Simple Modules}
-
-In this chapter we classify the finite-dimensional simple
-\(\mathfrak{g}\)-modules for a finite-dimensional semisimple Lie algebra
-\(\mathfrak{g}\) over \(K\). At the heart of our analysis of
-\(\mathfrak{sl}_2(K)\) and \(\mathfrak{sl}_3(K)\) was the decision to consider
-the eigenspace decomposition
-\begin{equation}\label{sym-diag}
-  M = \bigoplus_\lambda M_\lambda
-\end{equation}
-
-This was simple enough to do in the case of \(\mathfrak{sl}_2(K)\), but the
-rational behind it and the reason why equation (\ref{sym-diag}) holds are
-harder to explain in the case of \(\mathfrak{sl}_3(K)\). The eigenspace
-decomposition associated with an operator \(M \to M\) is a very well-known
-tool, and readers familiarized with basic concepts of linear algebra should be
-used to this type of argument. On the other hand, the eigenspace decomposition
-of \(M\) with respect to the action of an arbitrary subalgebra \(\mathfrak{h}
-\subset \mathfrak{gl}(M)\) is neither well-known nor does it hold in general:
-as indicated in the previous chapter, it may very well be that
-\[
-  \bigoplus_{\lambda \in \mathfrak{h}^*} M_\lambda \subsetneq M
-\]
-
-We should note, however, that these two cases are not as different as they may
-sound at first glance. Specifically, we can regard the eigenspace decomposition
-of a \(\mathfrak{sl}_2(K)\)-module \(M\) with respect to the eigenvalues of the
-action of \(h\) as the eigenvalue decomposition of \(M\) with respect to the
-action of the subalgebra \(\mathfrak{h} = K h \subset \mathfrak{sl}_2(K)\).
-Furthermore, in both cases \(\mathfrak{h} \subset \mathfrak{sl}_n(K)\) is the
-subalgebra of diagonal matrices, which is Abelian. The fundamental difference
-between these two cases is thus the fact that \(\dim \mathfrak{h} = 1\) for
-\(\mathfrak{h} \subset \mathfrak{sl}_2(K)\) while \(\dim \mathfrak{h} > 1\) for
-\(\mathfrak{h} \subset \mathfrak{sl}_3(K)\). The question then is: why did we
-choose \(\mathfrak{h}\) with \(\dim \mathfrak{h} > 1\) for
-\(\mathfrak{sl}_3(K)\)?
-
-The rational behind fixing an Abelian subalgebra \(\mathfrak{h}\) is a simple
-one: we have seen in the previous chapter that representations of Abelian
-algebras are generally much simpler to understand than the general case. Thus
-it make sense to decompose a given \(\mathfrak{g}\)-module \(M\) of into
-subspaces invariant under the action of \(\mathfrak{h}\), and then analyze how
-the remaining elements of \(\mathfrak{g}\) act on these subspaces. The bigger
-\(\mathfrak{h}\) is, the simpler our problem gets, because there are fewer
-elements outside of \(\mathfrak{h}\) left to analyze.
-
-Hence we are generally interested in maximal Abelian subalgebras \(\mathfrak{h}
-\subset \mathfrak{g}\), which leads us to the following definition.
-
-\begin{definition}\index{Cartan subalgebra}
-  A subalgebra \(\mathfrak{h} \subset \mathfrak{g}\) is called \emph{a Cartan
-  subalgebra of \(\mathfrak{g}\)} if is self-normalizing -- i.e. \([X, H] \in
-  \mathfrak{h}\) for all \(H \in \mathfrak{h}\) if, and only if \(X \in
-  \mathfrak{h}\) -- and nilpotent. Equivalently for reductive \(\mathfrak{g}\),
-  \(\mathfrak{h}\) is called \emph{a Cartan subalgebra of \(\mathfrak{g}\)} if
-  it is Abelian, \(\operatorname{ad}(H)\) is diagonalizable for each \(H \in
-  \mathfrak{h}\) and if \(\mathfrak{h}\) is maximal with respect to the former
-  two properties.
-\end{definition}
-
-\begin{proposition}
-  There exists a Cartan subalgebra \(\mathfrak{h} \subset \mathfrak{g}\).
-\end{proposition}
-
-\begin{proof}
-  Notice that \(0 \subset \mathfrak{g}\) is an Abelian subalgebra whose
-  elements act as diagonal operators via the adjoint \(\mathfrak{g}\)-module.
-  Indeed, \(0\), the only element of \(0 \subset \mathfrak{g}\), is such that
-  \(\operatorname{ad}(0) = 0\) is a diagonalizable operator. Furthermore, given
-  a chain of Abelian subalgebras
-  \[
-    0 \subset \mathfrak{h}_1 \subset \mathfrak{h}_2 \subset \cdots
-  \]
-  such that \(\operatorname{ad}(H)\) is a diagonal operator for each \(H \in
-  \mathfrak{h}_i\), the subalgebra \(\bigcup_i \mathfrak{h}_i \subset
-  \mathfrak{g}\) is Abelian, and its elements also act diagonally in
-  \(\mathfrak{g}\). It then follows from Zorn's Lemma that there exists a
-  subalgebra \(\mathfrak{h}\) which is maximal with respect to both these
-  properties, also known as a Cartan subalgebra.
-\end{proof}
-
-We have already seen some concrete examples. Namely\dots
-
-\begin{example}\label{ex:cartan-of-gl}
-  The Lie subalgebra
-  \[
-    \mathfrak{h} =
-    \begin{pmatrix}
-           K &      0 & \cdots &      0 \\
-           0 &      K & \cdots &      0 \\
-      \vdots & \vdots & \ddots & \vdots \\
-           0 &      0 & \cdots &      K
-    \end{pmatrix}
-    \subset \mathfrak{gl}_n(K)
-  \]
-  of diagonal matrices is a Cartan subalgebra.
-  Indeed, every pair of diagonal matrices commutes, so that \(\mathfrak{h}\)
-  is an Abelian -- and hence nilpotent -- subalgebra. A
-  simple calculation also shows that if \(i \ne j\) then the coefficient of
-  \(E_{i j}\) in \([E_{i i}, X]\) is the same as the coefficient of \(E_{i j}\)
-  in \(X\), for all \(X \in \mathfrak{gl}_n(K)\). In particular, if \([E_{i i},
-  X]\) is diagonal for all \(i\), then so is \(X\) -- i.e. \(\mathfrak{h}\) is
-  self-normalizing.
-\end{example}
-
-\begin{example}
-  Let \(\mathfrak{h}\) be as in Example~\ref{ex:cartan-of-gl}. Then the
-  subalgebra \(\mathfrak{h} \cap \mathfrak{sl}_n(K)\) of traceless diagonal
-  matrices is a Cartan subalgebra of \(\mathfrak{sl}_n(K)\).
-\end{example}
-
-\begin{example}\label{ex:cartan-direct-sum}
-  Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras and
-  \(\mathfrak{h}_i \subset \mathfrak{g}_i\) be Cartan subalgebras. Then
-  \(\mathfrak{h}_1 \oplus \mathfrak{h}_2\) is a Cartan subalgebra of
-  \(\mathfrak{g}_1 \oplus \mathfrak{g}_2\).
-\end{example}
-
-\index{Cartan subalgebra!simultaneous diagonalization}
-The intersection of such subalgebra with \(\mathfrak{sl}_n(K)\) -- i.e. the
-subalgebra of traceless diagonal matrices -- is a Cartan subalgebra of
-\(\mathfrak{sl}_n(K)\). In particular, if \(n = 2\) or \(n = 3\) we get to the
-subalgebras described the previous chapter. The remaining question then is: if
-\(\mathfrak{h} \subset \mathfrak{g}\) is a Cartan subalgebra and \(M\) is a
-\(\mathfrak{g}\)-module, does the eigenspace decomposition
-\[
-  M = \bigoplus_\lambda M_\lambda
-\]
-of \(M\) hold? The answer to this question turns out to be yes. This is a
-consequence of something known as \emph{simultaneous diagonalization}, which is
-the primary tool we will use to generalize the results of the previous section.
-What is simultaneous diagonalization all about then?
-
-\begin{definition}\label{def:sim-diag}
-  Given a \(K\)-vector space \(V\), a set of operators \(\{T_j : V \to V\}_j\)
-  is called \emph{simultaneously diagonalizable} if there is a basis \(\{v_1,
-  \ldots, v_n\}\) for \(V\) such that \(T_j v_i\) is a scalar multiple of
-  \(v_i\), for all \(i, j\).
-\end{definition}
-
-\begin{proposition}
-  Given a \emph{finite-dimensional} vector space \(V\), a set of diagonalizable
-  operators \(V \to V\) is simultaneously diagonalizable if, and only if all of
-  its elements commute with one another.
-\end{proposition}
-
-We should point out that simultaneous diagonalization \emph{only works in the
-finite-dimensional setting}. In fact, simultaneous diagonalization is usually
-framed as an equivalent statement about diagonalizable \(n \times n\) matrices.
-Simultaneous diagonalization implies that to show \(M = \bigoplus_\lambda
-M_\lambda\) it suffices to show that \(H\!\restriction_M : M \to M\) is a
-diagonalizable operator for each \(H \in \mathfrak{h}\). To that end, we
-introduce \emph{the Jordan decomposition of an operator} and \emph{the abstract
-Jordan decomposition of a semisimple Lie algebra}.
-
-\begin{proposition}[Jordan]
-  Given a finite-dimensional vector space \(V\) and an operator \(T : V \to
-  V\), there are unique commuting operators \(T_{\operatorname{ss}},
-  T_{\operatorname{nil}} : V \to V\), with \(T_{\operatorname{ss}}\)
-  diagonalizable and \(T_{\operatorname{nil}}\) nilpotent, such that \(T =
-  T_{\operatorname{ss}} + T_{\operatorname{nil}}\). The pair
-  \((T_{\operatorname{ss}}, T_{\operatorname{nil}})\) is known as \emph{the Jordan
-  decomposition of \(T\)}.
-\end{proposition}
-
-\begin{proposition}\index{abstract Jordan decomposition}
-  Given \(\mathfrak{g}\) semisimple and \(X \in \mathfrak{g}\), there are
-  \(X_{\operatorname{ss}}, X_{\operatorname{nil}} \in \mathfrak{g}\) such that \(X
-  = X_{\operatorname{ss}} + X_{\operatorname{nil}}\), \([X_{\operatorname{ss}},
-  X_{\operatorname{nil}}] = 0\), \(\operatorname{ad}(X_{\operatorname{ss}})\) is a
-  diagonalizable operator and \(\operatorname{ad}(X_{\operatorname{nil}})\) is a
-  nilpotent operator. The pair \((X_{\operatorname{ss}}, X_{\operatorname{nil}})\)
-  is known as \emph{the Jordan decomposition of \(X\)}.
-\end{proposition}
-
-It should be clear from the uniqueness of
-\(\operatorname{ad}(X)_{\operatorname{ss}}\) and
-\(\operatorname{ad}(X)_{\operatorname{nil}}\) that the Jordan decomposition of
-\(\operatorname{ad}(X)\) is \(\operatorname{ad}(X) =
-\operatorname{ad}(X_{\operatorname{ss}}) +
-\operatorname{ad}(X_{\operatorname{nil}})\). What is perhaps more remarkable is
-the fact this holds for \emph{any} finite-dimensional \(\mathfrak{g}\)-module.
-In other words\dots
-
-\begin{proposition}\label{thm:preservation-jordan-form}
-  Let \(M\) be a finite-dimensional \(\mathfrak{g}\)-module and \(X
-  \in \mathfrak{g}\). Denote by \(X\!\restriction_M\) the action of \(X\) on
-  \(M\). Then \(X_{\operatorname{ss}}\!\restriction_M =
-  (X\!\restriction_M)_{\operatorname{ss}}\) and
-  \(X_{\operatorname{nil}}\!\restriction_M =
-  (X\!\restriction_M)_{\operatorname{nil}}\).
-\end{proposition}
-
-This last result is known as \emph{the preservation of the Jordan form}, and a
-proof can be found in appendix C of \cite{fulton-harris}. As promised this
-implies\dots
-
-\begin{corollary}\label{thm:finite-dim-is-weight-mod}
-  Let \(\mathfrak{g}\) be a semisimple Lie algebra, \(\mathfrak{h} \subset
-  \mathfrak{g}\) be a Cartan subalgebra and \(M\) be any finite-dimensional
-  \(\mathfrak{g}\)-module. Then there is a basis \(\{m_1, \ldots,
-  m_r\}\) of \(M\) so that each \(m_i\) is simultaneously an eigenvector of all
-  elements of \(\mathfrak{h}\) -- i.e. each element of \(\mathfrak{h}\) acts as
-  a diagonal matrix in this basis. In other words, there are linear functionals
-  \(\lambda_i \in \mathfrak{h}^*\) so that
-  \(
-    H \cdot m_i = \lambda_i(H) m_i
-  \)
-  for all \(H \in \mathfrak{h}\). In particular,
-  \[
-    M = \bigoplus_{\lambda \in \mathfrak{h}^*} M_\lambda
-  \]
-\end{corollary}
-
-\begin{proof}
-  Fix some \(H \in \mathfrak{h}\). It suffices to show that \(H\!\restriction_M
-  : M \to M\) is a diagonalizable operator.
-
-  If we write \(H = H_{\operatorname{ss}} + H_{\operatorname{nil}}\) for the
-  abstract Jordan decomposition of \(H\), we know
-  \(\operatorname{ad}(H_{\operatorname{ss}}) =
-  \operatorname{ad}(H)_{\operatorname{ss}}\). But \(\operatorname{ad}(H)\) is a
-  diagonalizable operator, so that \(\operatorname{ad}(H)_{\operatorname{ss}} =
-  \operatorname{ad}(H)\). This implies
-  \(\operatorname{ad}(H_{\operatorname{nil}}) =
-  \operatorname{ad}(H)_{\operatorname{nil}} = 0\), so that
-  \(H_{\operatorname{nil}}\) is a central element of \(\mathfrak{g}\). Since
-  \(\mathfrak{g}\) is semisimple, \(H_{\operatorname{nil}} = 0\).
-  Proposition~\ref{thm:preservation-jordan-form} then implies
-  \((H\!\restriction_M)_{\operatorname{nil}} =
-  H_{\operatorname{nil}}\!\restriction_M = 0\), so \(H\!\restriction_M =
-  (H\!\restriction_M)_{\operatorname{ss}}\) is a diagonalizable operator.
-\end{proof}
-
-We should point out that this last proof only works for semisimple Lie
-algebras. This is because we rely heavily on
-Proposition~\ref{thm:preservation-jordan-form}, as well in the fact that
-semisimple Lie algebras are centerless. In fact,
-Corollary~\ref{thm:finite-dim-is-weight-mod} fails even for reductive Lie
-algebras. For a counterexample, consider the algebra \(\mathfrak{g} = K\): the
-Cartan subalgebra of \(\mathfrak{g}\) is \(\mathfrak{g}\) itself, and a
-\(\mathfrak{g}\)-module is simply a vector space \(M\) endowed with an operator
-\(M \to M\) -- which corresponds to the action of \(1 \in \mathfrak{g}\) on
-\(M\). In particular, if we choose an operator \(M \to M\) which is \emph{not}
-diagonalizable we find \(M \ne \bigoplus_{\lambda \in \mathfrak{h}^*}
-M_\lambda\).
-
-However, Corollary~\ref{thm:finite-dim-is-weight-mod} does work for reductive
-\(\mathfrak{g}\) if we assume that the \(\mathfrak{g}\)-module \(M\) in
-question is simple, since central elements of \(\mathfrak{g}\) act on simple
-\(\mathfrak{g}\)-modules as scalar operators. The hypothesis of
-finite-dimensionality is also of huge importance. For instance, consider\dots
-
-\begin{example}\label{ex:regular-mod-is-not-weight-mod}
-  Let \(\mathcal{U}(\mathfrak{g})\) denote the regular \(\mathfrak{g}\)-module.
-  Notice that \(\mathcal{U}(\mathfrak{g})_\lambda = 0\) for all \(\lambda \in
-  \mathfrak{h}^*\). Indeed, since \(\mathcal{U}(\mathfrak{g})\) is a domain, if
-  \((H - \lambda(H)) u = 0\) for some nonzero \(H \in \mathfrak{h}\) then \(u =
-  0\). In particular,
-  \[
-    \bigoplus_{\lambda \in \mathfrak{h}^*} \mathcal{U}(\mathfrak{g})_\lambda
-    = 0 \neq \mathcal{U}(\mathfrak{g})
-  \]
-\end{example}
-
-As a first consequence of Corollary~\ref{thm:finite-dim-is-weight-mod} we
-show\dots
-
-\begin{corollary}
-  The restriction of the Killing form \(\kappa\) to \(\mathfrak{h}\) is
-  non-degenerate.
-\end{corollary}
-
-\begin{proof}
-  Consider the root space decomposition \(\mathfrak{g} = \mathfrak{g}_0 \oplus
-  \bigoplus_\alpha \mathfrak{g}_\alpha\) of the adjoint
-  \(\mathfrak{g}\)-module, where \(\alpha\) ranges over all nonzero eigenvalues
-  of the adjoint action of \(\mathfrak{h}\). We claim \(\mathfrak{g}_0 =
-  \mathfrak{h}\).
-
-  Indeed, since \(\mathfrak{h}\) is Abelian, \(\operatorname{ad}(\mathfrak{h})
-  \mathfrak{h} = 0\) -- i.e. \(\mathfrak{h} \subset \mathfrak{g}_0\). On the
-  other hand, since \(\mathfrak{h}\) is self-normalizing, if \([X, H] = 0 \in
-  \mathfrak{h}\) for all \(H \in \mathfrak{h}\) then \(X \in \mathfrak{h}\) --
-  i.e. \(\mathfrak{g}_0 \subset \mathfrak{h}\). So the eigenspace decomposition
-  becomes
-  \[
-    \mathfrak{g} = \mathfrak{h} \oplus \bigoplus_\alpha \mathfrak{g}_\alpha
-  \]
-
-  We furthermore claim that \(\mathfrak{h} = \mathfrak{g}_0\) is orthogonal to
-  \(\mathfrak{g}_\alpha\) with respect to \(\kappa\) for any \(\alpha \ne 0\).
-  Indeed, given \(X \in \mathfrak{g}_\alpha\) and \(H_1, H_2 \in \mathfrak{h}\)
-  with \(\alpha(H_1) \ne 0\) we have
-  \[
-    \alpha(H_1) \cdot \kappa(X, H_2)
-    = \kappa([H_1, X], H_2)
-    = - \kappa([X, H_1], H_2)
-    = - \kappa(X, [H_1, H_2])
-    = 0
-  \]
-
-  Hence the non-degeneracy of \(\kappa\) implies the non-degeneracy of its
-  restriction.
-\end{proof}
-
-We should point out that the restriction of \(\kappa\) to \(\mathfrak{h}\) is
-\emph{not} the Killing form of \(\mathfrak{h}\). In fact, since
-\(\mathfrak{h}\) is Abelian, its Killing form is identically zero -- which is
-hardly ever a non-degenerate form.
-
-\begin{note}
-  Since \(\kappa\) induces an isomorphism \(\mathfrak{h} \isoto
-  \mathfrak{h}^*\), it induces a bilinear form \((\kappa(X, \cdot), \kappa(Y,
-  \cdot)) \mapsto \kappa(X, Y)\) in \(\mathfrak{h}^*\). As in
-  section~\ref{sec:sl3-reps}, we denote this form by \(\kappa\) as well.
-\end{note}
-
-We now have most of the necessary tools to reproduce the results of the
-previous chapter in a general setting. Let \(\mathfrak{g}\) be a
-finite-dimensional semisimple algebra with a Cartan subalgebra \(\mathfrak{h}\)
-and let \(M\) be a finite-dimensional simple \(\mathfrak{g}\)-module. We will
-proceed, as we did before, by generalizing the results of the previous two
-sections in order. By now the pattern should be starting to become clear, so we
-will mostly omit technical details and proofs analogous to the ones on the
-previous sections. Further details can be found in appendix D of
-\cite{fulton-harris} and in \cite{humphreys}.
-
-\section{The Geometry of Roots and Weights}
-
-We begin our analysis, as we did for \(\mathfrak{sl}_2(K)\) and
-\(\mathfrak{sl}_3(K)\), by investigating the locus of roots of and weights of
-\(\mathfrak{g}\). Throughout chapter~\ref{ch:sl3} we have seen that the weights
-of any given finite-dimensional module of \(\mathfrak{sl}_2(K)\) or
-\(\mathfrak{sl}_3(K)\) can only assume very rigid configurations. For instance,
-we have seen that the roots of \(\mathfrak{sl}_2(K)\) and
-\(\mathfrak{sl}_3(K)\) are symmetric with respect to the origin. In this
-chapter we will generalize most results from chapter~\ref{ch:sl3} regarding the
-rigidity of the geometry of the set of weights of a given module.
-
-As for the aforementioned result on the symmetry of roots, this turns out to be
-a general fact, which is a consequence of the non-degeneracy of the restriction
-of the Killing form to the Cartan subalgebra.
-
-\begin{proposition}\label{thm:weights-symmetric-span}
-  The roots \(\alpha\) of \(\mathfrak{g}\) are symmetrical about the origin --
-  i.e. \(- \alpha\) is also a root -- and they span all of \(\mathfrak{h}^*\).
-\end{proposition}
-
-\begin{proof}
-  We will start with the first claim. Let \(\alpha\) and \(\beta\) be two
-  roots. Notice \([\mathfrak{g}_\alpha, \mathfrak{g}_\beta] \subset
-  \mathfrak{g}_{\alpha + \beta}\). Indeed, if \(X \in \mathfrak{g}_\alpha\) and
-  \(Y \in \mathfrak{g}_\beta\) then
-  \[
-    [H [X, Y]]
-    = [X, [H, Y]] - [Y, [H, X]]
-    = (\alpha + \beta)(H) \cdot [X, Y]
-  \]
-  for all \(H \in \mathfrak{h}\).
-
-  This implies that if \(\alpha + \beta \ne 0\) then \(\operatorname{ad}(X)
-  \operatorname{ad}(Y)\) is nilpotent: if \(Z \in \mathfrak{g}_\gamma\) then
-  \[
-    (\operatorname{ad}(X) \operatorname{ad}(Y))^r Z
-    = [X, [Y, [ \ldots, [X, [Y, Z]]] \ldots ]
-    \in \mathfrak{g}_{r \alpha + r \beta + \gamma}
-    = 0
-  \]
-  for \(r\) large enough. In particular, \(\kappa(X, Y) =
-  \operatorname{Tr}(\operatorname{ad}(X) \operatorname{ad}(Y)) = 0\). Now if
-  \(- \alpha\) is not an eigenvalue we find \(\kappa(X, \mathfrak{g}_\beta) =
-  0\) for all roots \(\beta\), which contradicts the non-degeneracy of
-  \(\kappa\). Hence \(- \alpha\) must be an eigenvalue of the adjoint action of
-  \(\mathfrak{h}\).
-
-  For the second statement, note that if the roots of \(\mathfrak{g}\) do not
-  span all of \(\mathfrak{h}^*\) then there is some nonzero \(H \in
-  \mathfrak{h}\) such that \(\alpha(H) = 0\) for all roots \(\alpha\), which is
-  to say, \(\operatorname{ad}(H) X = [H, X] = 0\) for all \(X \in
-  \mathfrak{g}\). Another way of putting it is to say \(H\) is an element of
-  the center \(\mathfrak{z} = 0\) of \(\mathfrak{g}\), a contradiction.
-\end{proof}
-
-Furthermore, as in the case of \(\mathfrak{sl}_2(K)\) and
-\(\mathfrak{sl}_3(K)\) one can show\dots
-
-\begin{proposition}\label{thm:root-space-dim-1}
-  The root spaces \(\mathfrak{g}_\alpha\) are all \(1\)-dimensional.
-\end{proposition}
-
-The proof of the first statement of
-Proposition~\ref{thm:weights-symmetric-span} highlights something interesting:
-if we fix some eigenvalue \(\alpha\) of the adjoint action of \(\mathfrak{h}\)
-on \(\mathfrak{g}\) and a eigenvector \(X \in \mathfrak{g}_\alpha\), then for
-each \(H \in \mathfrak{h}\) and \(m \in M_\lambda\) we find
-\[
-  H \cdot (X \cdot m)
-  = X H \cdot m + [H, X] \cdot m
-  = (\lambda + \alpha)(H) X \cdot m
-\]
-
-Thus \(X\) sends \(m\) to \(M_{\lambda + \alpha}\). We have encountered this
-formula twice in these notes: again, we find \(\mathfrak{g}_\alpha\) \emph{acts
-on \(M\) by translating vectors between eigenspaces}. In particular, if we
-denote by \(\Delta\) the set of all roots of \(\mathfrak{g}\) then\dots
-
-\begin{theorem}\label{thm:weights-congruent-mod-root}\index{weights!root lattice}
-  The weights of a finite-dimensional simple \(\mathfrak{g}\)-module \(M\) are
-  all congruent modulo the root lattice \(Q = \mathbb{Z} \Delta\) of
-  \(\mathfrak{g}\). In other words, all weights of \(M\) lie in the same
-  \(Q\)-coset \(\xi \in \mfrac{\mathfrak{h}^*}{Q}\).
-\end{theorem}
-
-Again, we may leverage our knowledge of \(\mathfrak{sl}_2(K)\) to obtain
-further restrictions on the geometry of the locus of weights of \(M\). Namely,
-as in the case of \(\mathfrak{sl}_3(K)\) we show\dots
-
-\begin{proposition}\label{thm:distinguished-subalgebra}
-  Given a root \(\alpha\) of \(\mathfrak{g}\) the subspace
-  \(\mathfrak{s}_\alpha = \mathfrak{g}_\alpha \oplus \mathfrak{g}_{- \alpha}
-  \oplus [\mathfrak{g}_\alpha, \mathfrak{g}_{- \alpha}]\) is a subalgebra
-  isomorphic to \(\mathfrak{sl}_2(K)\).
-\end{proposition}
-
-\begin{corollary}\label{thm:distinguished-subalg-rep}
-  For all weights \(\mu\), the subspace
-  \[
-    \bigoplus_k M_{\mu - k \alpha}
-  \]
-  is invariant under the action of the subalgebra \(\mathfrak{s}_\alpha\)
-  and the weight spaces in this string match the eigenspaces of \(h\).
-\end{corollary}
-
-The proof of Proposition~\ref{thm:distinguished-subalgebra} is very technical
-in nature and we won't include it here, but the idea behind it is simple:
-recall that \(\mathfrak{g}_\alpha\) and \(\mathfrak{g}_{- \alpha}\) are both
-\(1\)-dimensional, so that \(\dim [\mathfrak{g}_\alpha, \mathfrak{g}_{-
-\alpha}]\) is at most 1. We check that \([\mathfrak{g}_\alpha, \mathfrak{g}_{-
-\alpha}] \ne 0\) and that no generator of \([\mathfrak{g}_\alpha,
-\mathfrak{g}_{- \alpha}]\) is annihilated by \(\alpha\), so that by adjusting
-scalars we can find \(E_\alpha \in \mathfrak{g}_\alpha\) and \(F_\alpha \in
-\mathfrak{g}_{- \alpha}\) such that \(H_\alpha = [E_\alpha, F_\alpha]\)
-satisfies
-\begin{align*}
-  [H_\alpha, F_\alpha] & = -2 F_\alpha &
-  [H_\alpha, E_\alpha] & =  2 E_\alpha
-\end{align*}
-
-The elements \(E_\alpha, F_\alpha \in \mathfrak{g}\) are not uniquely
-determined by this condition, but \(H_\alpha\) is. As promised, the second
-statement of Corollary~\ref{thm:distinguished-subalg-rep} imposes strong
-restrictions on the weights of \(M\). Namely, if \(\lambda\) is a weight,
-\(\lambda(H_\alpha)\) is an eigenvalue of \(h\) on some
-\(\mathfrak{sl}_2(K)\)-module, so it must be an integer. In other words\dots
-
-\begin{definition}\label{def:weight-lattice}\index{weights!weight lattice}
-  The lattice \(P = \{ \lambda \in \mathfrak{h}^* : \lambda(H_\alpha) \in
-  \mathbb{Z} \, \forall \alpha \in \Delta \} \subset \mathfrak{h}^*\) is called
-  \emph{the weight lattice of \(\mathfrak{g}\)}. We call the elements of \(P\)
-  \emph{integral}.
-\end{definition}
-
-\begin{proposition}\label{thm:weights-fit-in-weight-lattice}
-  The weights of a finite-dimensional simple \(\mathfrak{g}\)-module \(M\) of
-  all lie in the weight lattice \(P\).
-\end{proposition}
-
-Proposition~\ref{thm:weights-fit-in-weight-lattice} is clearly analogous to
-Corollary~\ref{thm:sl3-weights-fit-in-weight-lattice}. In fact, the weight
-lattice of \(\mathfrak{sl}_3(K)\) -- as in Definition~\ref{def:weight-lattice}
--- is precisely \(\mathbb{Z} \langle \alpha_1, \alpha_2, \alpha_3 \rangle\). To
-proceed further, we would like to take \emph{the highest weight of \(M\)} as in
-section~\ref{sec:sl3-reps}, but the meaning of \emph{highest} is again unclear
-in this situation. We could simply fix a linear function \(\mathbb{Q} P \to
-\mathbb{Q}\) -- as we did in section~\ref{sec:sl3-reps} -- and choose a weight
-\(\lambda\) of \(M\) that maximizes this functional, but at this point it is
-convenient to introduce some additional tools to our arsenal. These tools are
-called \emph{basis}.
-
-\begin{definition}\label{def:basis-of-root}\index{weights!basis}
-  A subset \(\Sigma = \{\beta_1, \ldots, \beta_r\} \subset \Delta\) of linearly
-  independent roots is called \emph{a basis for \(\Delta\)} if, given \(\alpha
-  \in \Delta\), there are \(k_1, \ldots, k_r \in \mathbb{N}\) such that
-  \(\alpha = \pm(k_1 \beta_1 + \cdots + k_r \beta_r)\).
-\end{definition}
-
-The interesting thing about basis for \(\Delta\) is that they allow us to
-compare weights of a given \(\mathfrak{g}\)-module. At this point the reader
-should be asking himself: how? Definition~\ref{def:basis-of-root} isn't exactly
-all that intuitive. Well, the thing is that any choice of basis induces a
-partial order in \(Q\), where elements are ordered by their \emph{heights}.
-
-\begin{definition}\index{weights!orderings of roots}
-  Let \(\Sigma = \{\beta_1, \ldots, \beta_r\}\) be a basis for \(\Delta\).
-  Given \(\alpha = k_1 \beta_1 + \cdots + k_r \beta_r \in Q\) with \(k_1,
-  \ldots, k_r \in \mathbb{Z}\), we call the number \(\operatorname{ht}(\alpha)
-  = k_1 + \cdots + k_r \in \mathbb{Z}\) \emph{the height of \(\alpha\)}. We say
-  that \(\alpha \preceq \beta\) if \(\operatorname{ht}(\alpha) \le
-  \operatorname{ht}(\beta)\).
-\end{definition}
-
-\begin{definition}
-  Given a basis \(\Sigma\) for \(\Delta\), there is a canonical
-  partition\footnote{Notice that $\operatorname{ht}(\alpha) = 0$ if, and only
-  if $\alpha = 0$. Since $0$ is, by definition, not a root, the sets $\Delta^+$
-  and $\Delta^-$ account for all roots.} \(\Delta^+ \cup \Delta^- = \Delta\),
-  where \(\Delta^+ = \{ \alpha \in \Delta : \alpha \succ 0 \}\) and \(\Delta^-
-  = \{ \alpha \in \Delta : \alpha \prec 0 \}\). The elements of \(\Delta^+\)
-  and \(\Delta^-\) are called \emph{positive} and \emph{negative roots},
-  respectively.
-\end{definition}
-
-\begin{definition}
-  Let \(\Sigma\) be a basis for \(\Delta\). The subalgebra \(\mathfrak{b} =
-  \mathfrak{h} \oplus \bigoplus_{\alpha \in \Delta^+} \mathfrak{g}_\alpha\) is
-  called \emph{the Borel subalgebra associated with \(\mathfrak{h}\) and
-  \(\Sigma\)}.
-\end{definition}
-
-It should be obvious that the binary relation \(\preceq\) in \(Q\) is a partial
-order. In addition, we may compare the elements of a given \(Q\)-coset
-\(\lambda + Q\) by comparing their difference with \(0 \in Q\). In other words,
-given \(\lambda \in \mu + Q\), we say \(\lambda \preceq \mu\) if \(\lambda -
-\mu \preceq 0\). In particular, since the weights of \(M\) all lie in a single
-\(Q\)-coset, we may compare them in this way. Given a basis \(\Sigma\) for
-\(\Delta\) we may take ``the highest weight of \(M\)'' as a maximal weight
-\(\lambda\) of \(M\). The obvious question then is: can we always find a basis
-for \(\Delta\)?
-
-\begin{proposition}
-  There is a basis \(\Sigma\) for \(\Delta\).
-\end{proposition}
-
-The intuition behind the proof of this proposition is similar to our original
-idea of fixing a direction in \(\mathfrak{h}^*\) in the case of
-\(\mathfrak{sl}_3(K)\). Namely, one can show that \(\kappa(\alpha, \beta) \in
-\mathbb{Z}\) for all \(\alpha, \beta \in \Delta\), so that the Killing form
-\(\kappa\) restricts to a nondegenerate \(\mathbb{Q}\)-bilinear form
-\(\mathbb{Q} \Delta \times \mathbb{Q} \Delta \to \mathbb{Q}\). We can then fix
-a nonzero vector \(\gamma \in \mathbb{Q} \Delta\) and consider the orthogonal
-projection \(f : \mathbb{Q} \Delta \to \mathbb{Q} \gamma \cong \mathbb{Q}\). We
-say a root \(\alpha \in \Delta\) is \emph{positive} if \(f(\alpha) > 0\), and
-we call a positive root \(\alpha\) \emph{simple} if it cannot be written as the
-sum two other positive roots. The subset \(\Sigma \subset \Delta\) of all
-simple roots is a basis for \(\Delta\), and all other basis can be shown to
-arise in this way.
-
-Fix some basis \(\Sigma\) for \(\Delta\), with corresponding decomposition
-\(\Delta^+ \cup \Delta^- = \Delta\). Let \(\lambda\) be a maximal weight of
-\(M\). We call \(\lambda\) \emph{the highest weight of \(M\)}, and we call any
-nonzero \(m \in M_\lambda\) \emph{a highest weight vector}. The strategy then
-is to describe all weight spaces of \(M\) in terms of \(\lambda\) and \(m\), as
-in Theorem~\ref{thm:sl3-irr-weights-class}. Unsurprisingly we do so by
-reproducing the proof of the case of \(\mathfrak{sl}_3(K)\).
-
-First, we note that any highest weight vector \(m \in M_\lambda\) is
-annihilated by all positive root spaces, for if \(\alpha \in \Delta^+\) then
-\(E_\alpha \cdot m \in M_{\lambda + \alpha}\) must be zero -- or otherwise we
-would have that \(\lambda + \alpha\) is a weight with \(\lambda \prec \lambda +
-\alpha\). In particular,
-\[
-  \bigoplus_{k \in \mathbb{Z}}   M_{\lambda - k \alpha}
-  = \bigoplus_{k \in \mathbb{N}} M_{\lambda - k \alpha}
-\]
-and \(\lambda(H_\alpha)\) is the right-most eigenvalue of the action of \(h\)
-on the \(\mathfrak{sl}_2(K)\)-module \(\bigoplus_k M_{\lambda - k \alpha}\).
-
-This has a number of important consequences. For instance\dots
-
-\begin{corollary}
-  If \(\alpha \in \Delta^+\) and \(\sigma_\alpha : \mathfrak{h}^* \to
-  \mathfrak{h}^*\) is the reflection in the hyperplane perpendicular to
-  \(\alpha\) with respect to the Killing form, the weights of \(M\) occurring
-  in the line joining \(\lambda\) and \(\sigma_\alpha\) are precisely the \(\mu
-  \in P\) lying between \(\lambda\) and \(\sigma_\alpha(\lambda)\).
-\end{corollary}
-
-\begin{proof}
-  Notice that any \(\mu \in P\) in the line joining \(\lambda\) and
-  \(\sigma_\alpha(\lambda)\) has the form \(\mu = \lambda - k \alpha\) for some
-  \(k\), so that \(M_\mu\) corresponds the eigenspace associated with the
-  eigenvalue \(\lambda(H_\alpha) - 2k\) of the action of \(h\) on \(\bigoplus_k
-  M_{\lambda - k \alpha}\). If \(\mu\) lies between \(\lambda\) and
-  \(\sigma_\alpha(\lambda)\) then \(k\) lies between \(0\) and
-  \(\lambda(H_\alpha)\), in which case \(M_\mu \neq 0\) and therefore \(\mu\)
-  is a weight.
-
-  On the other hand, if \(\mu\) does not lie between \(\lambda\) and
-  \(\sigma_\alpha(\lambda)\) then either \(k < 0\) or \(k >
-  \lambda(H_\alpha)\). Suppose \(\mu\) is a weight. In the first case \(\mu
-  \succ \lambda\), a contradiction. On the second case the fact that \(M_\mu
-  \ne 0\) implies \(M_{\lambda  + (k - \lambda(H_\alpha)) \alpha} =
-  M_{\sigma_\alpha(\mu)} \ne 0\), which contradicts the fact that \(M_{\lambda
-  + \ell \alpha} = 0\) for all \(\ell \ge 0\).
-\end{proof}
-
-This is entirely analogous to the situation of \(\mathfrak{sl}_3(K)\), where we
-found that the weights of the simple \(\mathfrak{sl}_3(K)\)-modules formed
-continuous strings symmetric with respect to the lines \(K \alpha\) with
-\(\kappa(\alpha_i - \alpha_j, \alpha) = 0\). As in the case of
-\(\mathfrak{sl}_3(K)\), the same class of arguments leads us to the
-conclusion\dots
-
-\begin{definition}\index{Weyl group}
-  We refer to the group \(W = \langle \sigma_\alpha : \alpha \in
-  \Delta^+ \rangle \subset \operatorname{O}(\mathfrak{h}^*)\) as \emph{the Weyl
-  group of \(\mathfrak{g}\)}.
-\end{definition}
-
-\begin{theorem}\label{thm:irr-weight-class}
-  The weights of a simple \(\mathfrak{g}\)-module \(M\) with highest weight
-  \(\lambda\) are precisely the elements of the weight lattice \(P\) congruent
-  to \(\lambda\) modulo the root lattice \(Q\) lying inside the convex hull of
-  the orbit of \(\lambda\) under the action of the Weyl group \(W\).
-\end{theorem}
-
-\index{Weyl group!actions}
-Aside from showing up in the previous theorem, the Weyl group will also play an
-important role in chapter~\ref{ch:mathieu} by virtue of the existence of a
-canonical action of \(W\) on \(\mathfrak{h}\). By its very nature,
-\(W\) acts in \(\mathfrak{h}^*\). If we conjugate the action
-\(\sigma\!\restriction_{\mathfrak{h}^*} : \mathfrak{h}^* \isoto
-\mathfrak{h}^*\) of some \(\sigma \in W\) by the isomorphism
-\(\mathfrak{h}^* \isoto \mathfrak{h}\) afforded by the restriction of the
-Killing for to \(\mathfrak{h}\) we get a linear automorphism
-\(\sigma\!\restriction_{\mathfrak{h}} : \mathfrak{h} \isoto \mathfrak{h}\). As
-it turns out, the \(\sigma\!\restriction_{\mathfrak{h}}\) can be extended to an
-automorphism of Lie algebras \(\mathfrak{g} \isoto \mathfrak{g}\). This
-translates into the following results, which we do not prove -- but see
-\cite[sec.~14.3]{humphreys}.
-
-\begin{proposition}\label{thm:weyl-group-action}
-  Given \(\alpha \in \Delta^+\), let\footnote{Notice that since $\mathfrak{g}$
-  is finite-dimensional, $\operatorname{ad}(X)$ is nilpotent for each root
-  vector $X \in \mathfrak{g}$, so that the linear automorphism
-  $e^{\operatorname{ad}(X)} = \operatorname{Id} + \operatorname{ad}(X) +
-  \frac{\operatorname{ad}(X)^2}{2!} + \cdots : \mathfrak{g} \isoto
-  \mathfrak{g}$ is well defined.} \(\tilde{\sigma}_\alpha =
-  e^{\operatorname{ad}(E_\alpha)} e^{- \operatorname{ad}(F_\alpha)}
-  e^{\operatorname{ad}(E_\alpha)} : \mathfrak{g} \isoto \mathfrak{g}\). Then
-  \(\tilde\sigma_\alpha\) is an automorphism of Lie algebras, and this defines
-  an action of \(W\) on \(\mathfrak{g}\) which is compatible with the
-  canonical action of \(W\) on \(\mathfrak{h}\) -- i.e.
-  \(\tilde\sigma\!\restriction_{\mathfrak{h}} =
-  \sigma\!\restriction_{\mathfrak{h}}\) for all \(\sigma \in W\).
-\end{proposition}
-
-\begin{note}
-  Notice that the action of \(W\) on \(\mathfrak{g}\) from
-  Proposition~\ref{thm:weyl-group-action} is not canonical, since it depends on
-  the choice of \(E_\alpha\) and \(F_\alpha\). Nevertheless, \(\mathfrak{h}\)
-  is stable under the action of \(W\) -- i.e. \(W \cdot
-  \mathfrak{h} \subset \mathfrak{h}\) -- and the restriction of this action to
-  \(\mathfrak{h}\) is independent of any choices.
-\end{note}
-
-We should point out that the results in this section regarding the geometry
-roots and weights are only the beginning of a well develop axiomatic theory of
-the so called \emph{root systems}, which was used by Cartan in the early 20th
-century to classify all finite-dimensional simple complex Lie algebras in terms
-of Dynking diagrams. This and much more can be found in \cite[III]{humphreys}
-and \cite[ch.~21]{fulton-harris}. Having found all of the weights of \(M\), the
-only thing we are missing for a complete classification is an existence and
-uniqueness theorem analogous to Theorem~\ref{thm:sl2-exist-unique} and
-Theorem~\ref{thm:sl3-existence-uniqueness}. This will be the focus of the next
-section.
-
-\section{Verma Modules \& the Highest Weight Theorem}
-
-It is already clear from the previous discussion that if \(\lambda\) is the
-highest weight of \(M\) then \(\lambda(H_\alpha) \ge 0\) for all positive roots
-\(\alpha\). In other words, having \(\lambda(H_\alpha) \ge 0\), for all
-\(\alpha \in \Delta^+\), is a necessary condition for the existence of a simple
-\(\mathfrak{g}\)-module with highest weight given by \(\lambda\). Surprisingly,
-this condition is also sufficient. In other words\dots
-
-\begin{definition}\index{weights!dominant weight}\index{weights!integral weight}
-  An element \(\lambda\) of \(P\) such that \(\lambda(H_\alpha) \ge 0\) for all
-  \(\alpha \in \Delta^+\) is referred to as an \emph{dominant integral weight
-  of \(\mathfrak{g}\)}.
-\end{definition}
-
-\begin{theorem}\label{thm:dominant-weight-theo}
-  For each dominant integral \(\lambda \in P\) there exists precisely one
-  finite-dimensional simple \(\mathfrak{g}\)-module \(M\) whose highest weight
-  is \(\lambda\).
-\end{theorem}
-
-\index{weights!Highest Weight Theorem} This is known as \emph{the Highest
-Weight Theorem}, and its proof is the focus of this section. The ``uniqueness''
-part of the theorem follows at once from the arguments used for
-\(\mathfrak{sl}_3(K)\). However, the ``existence'' part of the theorem is more
-nuanced.
-
-Our first instinct is, of course, to try to generalize the proof used for
-\(\mathfrak{sl}_3(K)\). The issue is that our proof relied heavily on our
-knowledge of the roots of \(\mathfrak{sl}_3(K)\). It is thus clear that we need
-a more systematic approach for the general setting. We begin by asking a
-simpler question: how can we construct \emph{any} -- potentially
-infinite-dimensional -- \(\mathfrak{g}\)-module \(M\) of highest weight
-\(\lambda\)? In the process of answering this question we will come across a
-surprisingly elegant solution to our problem.
-
-If \(M\) is a module with highest weight vector \(m^+ \in M_\lambda\), we
-already know \(H \cdot m^+ = \lambda(H) m^+\) for all \(\mathfrak{h}\) and \(X
-\cdot m^+ = 0\) for \(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\). If
-\(M\) is simple we find \(M = \mathcal{U}(\mathfrak{g}) \cdot m^+\), which
-implies the restriction of \(M\) to the Borel subalgebra \(\mathfrak{b} \subset
-\mathfrak{g}\) has a prescribed action. On the other hand, we have essentially
-no information about the action of the rest of \(\mathfrak{g}\) on \(M\).
-Nevertheless, given a \(\mathfrak{b}\)-module we may obtain a
-\(\mathfrak{g}\)-module by formally extending the action of \(\mathfrak{b}\)
-via induction. This leads us to the following definition.
-
-\begin{definition}\label{def:verma}\index{\(\mathfrak{g}\)-module!(generalized) Verma modules}
-  The \(\mathfrak{g}\)-module \(M(\lambda) =
-  \operatorname{Ind}_{\mathfrak{b}}^{\mathfrak{g}} K m^+\), where the action of
-  \(\mathfrak{b}\) on \(K m^+\) is given by \(H \cdot m^+ = \lambda(H) m^+\)
-  for all \(H \in \mathfrak{h}\) and \(X \cdot m^+ = 0\) for \(X \in
-  \mathfrak{g}_{\alpha}\), \(\alpha \in \Delta^+\), is called \emph{the Verma
-  module of weight \(\lambda\)}.
-\end{definition}
-
-It turns out that \(M(\lambda)\) enjoys many of the features we've grown used
-to in the past chapters. Explicitly\dots
-
-\begin{proposition}\label{thm:verma-is-weight-mod}
-  The Verma module \(M(\lambda)\) is generated \(m^+ = 1 \otimes m^+ \in
-  M(\lambda)\) as in Definition~\ref{def:verma}. The weight spaces
-  decomposition
-  \[
-    M(\lambda) = \bigoplus_{\mu \in \mathfrak{h}^*} M(\lambda)_\mu
-  \]
-  holds. Furthermore, \(\dim M(\lambda)_\mu < \infty\) for all \(\mu \in
-  \mathfrak{h}^*\) and \(\dim M(\lambda)_\lambda = 1\). Finally, \(\lambda\) is
-  the highest weight of \(M(\lambda)\), with highest weight vector given by
-  \(m^+ \in M(\lambda)\).
-\end{proposition}
-
-\begin{proof}
-  The PBW Theorem implies that \(M(\lambda)\) is spanned by the vectors
-  \(F_{\alpha_{i_1}} F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+\) for
-  \(\Delta^+ = \{\alpha_1, \ldots, \alpha_r\}\) and \(F_{\alpha_i} \in
-  \mathfrak{g}_{- \alpha_i}\) as in the proof of
-  Proposition~\ref{thm:distinguished-subalgebra}. But
-  \[
-    \begin{split}
-      H \cdot (F_{\alpha_{i_1}} F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+)
-      & = ([H, F_{\alpha_{i_1}}] + F_{\alpha_{i_1}} H)
-          F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+ \\
-      & = - \alpha_{i_1}(H) F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+
-        + F_{\alpha_{i_1}} ([H, F_{\alpha_{i_2}}] + F_{\alpha_{i_2}} H)
-          F_{\alpha_{i_2}} \cdots F_{\alpha_{i_s}} \cdot m^+ \\
-      & \;\; \vdots \\
-      & = (- \alpha_{i_1} - \cdots - \alpha_{i_s})(H)
-          F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+
-        + F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} H \cdot m^+ \\
-      & = (\lambda - \alpha_{i_1} - \cdots - \alpha_{i_s})(H)
-          F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+ \\
-      & \therefore F_{\alpha_{i_1}} \cdots F_{\alpha_{i_s}} \cdot m^+
-        \in M(\lambda)_{\lambda - \alpha_{i_1} - \cdots - \alpha_{i_s}}
-    \end{split}
-  \]
-
-  Hence \(M(\lambda) \subset \bigoplus_{\mu \in \mathfrak{h}^*}
-  M(\lambda)_\mu\), as desired. In fact we have established
-  \[
-    M(\lambda)
-    \subset
-    \bigoplus_{k_i \in \mathbb{N}}
-    M(\lambda)_{\lambda - k_1 \cdot \alpha_1 - \cdots - k_r \cdot \alpha_r}
-  \]
-  where \(\{\alpha_1, \ldots, \alpha_r\} = \Delta^+\), so that all weights of
-  \(M(\lambda)\) have the form \(\mu = \lambda - k_1 \cdot \alpha_1 - \cdots -
-  k_r \cdot \alpha_r\).
-
-  This already gives us that the weights of \(M(\lambda)\) are bounded by
-  \(\lambda\). To see that \(\lambda\) is indeed a weight, we show that
-  \(m^+\) is nonzero weight vector. Clearly \(m^+ \in M_\lambda\). The
-  Poincaré-Birkhoff-Witt Theorem implies \(\mathcal{U}(\mathfrak{g})\) is a
-  free \(\mathfrak{b}\)-module, so that
-  \[
-    M(\lambda)
-    \cong \left(\bigoplus_i \mathcal{U}(\mathfrak{b}) \right)
-    \otimes_{\mathcal{U}(\mathfrak{b})} K m^+
-    \cong \bigoplus_i \mathcal{U}(\mathfrak{b})
-    \otimes_{\mathcal{U}(\mathfrak{b})} K m^+
-    \cong \bigoplus_i K m^+
-    \ne 0
-  \]
-  as \(\mathfrak{b}\)-modules. We then conclude \(m^+ \ne 0\) in
-  \(M(\lambda)\), for if this was not the case we would find \(M(\lambda) =
-  \mathcal{U}(\mathfrak{g}) \cdot m^+ = 0\). Hence \(M_\lambda \ne 0\) and
-  therefore \(\lambda\) is the highest weight of \(M(\lambda)\), with highest
-  weight vector \(m^+\).
-
-  To see that \(\dim M(\lambda)_\mu < \infty\), simply note that there are only
-  finitely many monomials \(F_{\alpha_1}^{k_1} F_{\alpha_2}^{k_2} \cdots
-  F_{\alpha_s}^{k_s}\) such that \(\mu = \lambda + k_1 \cdot \alpha_1 + \cdots
-  + k_s \cdot \alpha_s\). Since \(M(\lambda)_\mu\) is spanned by the images of
-  \(m^+\) under such monomials, we conclude \(\dim M(\lambda) < \infty\). In
-  particular, there is a single monomials \(F_{\alpha_1}^{k_1}
-  F_{\alpha_2}^{k_2} \cdots F_{\alpha_s}^{k_s}\) such that \(\lambda = \lambda
-  + k_1 \cdot \alpha_1 + \cdots + k_s \cdot \alpha_s\) -- which is, of course,
-  the monomial where \(k_1 = \cdots = k_n = 0\). Hence \(\dim M_\lambda = 1\).
-\end{proof}
-
-\begin{example}\label{ex:sl2-verma}
-  If \(\mathfrak{g} = \mathfrak{sl}_2(K)\), then we can take \(\mathfrak{h} = K
-  h\) and \(\mathfrak{b} = K e \oplus K h\). If \(\lambda \in
-  \mathfrak{h}^*\) is the map \(h \mapsto 2\) then \(M(\lambda) =
-  \bigoplus_{k \ge 0} K f^k \cdot m^+\), and the action of
-  \(\mathfrak{sl}_2(K)\) on \(M(\lambda)\) is given by the formulas in
-  (\ref{eq:sl2-verma-formulas}). Visually,
-  \begin{center}
-    \begin{tikzcd}
-      \cdots            \rar[bend left=60]{-10}
-      & M(\lambda)_{-6} \rar[bend left=60]{-4} \lar[bend left=60]{1}
-      & M(\lambda)_{-4} \rar[bend left=60]{0}  \lar[bend left=60]{1}
-      & M(\lambda)_{-2} \rar[bend left=60]{2}  \lar[bend left=60]{1}
-      & M(\lambda)_0    \rar[bend left=60]{2}  \lar[bend left=60]{1}
-      & M(\lambda)_2                           \lar[bend left=60]{1}
-    \end{tikzcd}
-  \end{center}
-  where \(M(\lambda)_{2 - 2 k} = K f^k \cdot m^+\). Here the top arrows
-  represent the action of \(e\) and the bottom arrows represent the action of
-  \(f\). The scalars labeling each arrow indicate to which multiple of \(f^{k
-  \pm 1} \cdot m^+\) the elements \(e\) and \(f\) send \(f^k \cdot m^+\). The
-  string of weight spaces to the left of the diagram is infinite.
-  \begin{equation}\label{eq:sl2-verma-formulas}
-    \begin{aligned}
-      f^k \cdot m^+ & \overset{e}{\mapsto} (2 - k(k + 1)) f^{k - 1} \cdot m^+ &
-      f^k \cdot m^+ & \overset{f}{\mapsto} f^{k + 1} \cdot m^+                &
-      f^k \cdot m^+ & \overset{h}{\mapsto} (2 - 2k) f^k \cdot m^+             &
-    \end{aligned}
-  \end{equation}
-\end{example}
-
-The Verma module \(M(\lambda)\) should really be though-of as ``the freest
-highest weight \(\mathfrak{g}\)-module of weight \(\lambda\)''. Unfortunately
-for us, this is not a proof of Theorem~\ref{thm:dominant-weight-theo}, since in
-general \(M(\lambda)\) is neither simple nor finite-dimensional. Indeed, the
-dimension of \(M(\lambda)\) is the same as the codimension of
-\(\mathcal{U}(\mathfrak{b})\) in \(\mathcal{U}(\mathfrak{g})\), which is always
-infinite. Nevertheless, we may use \(M(\lambda)\) to prove
-Theorem~\ref{thm:dominant-weight-theo} as follows.
-
-Given a \(\mathfrak{g}\)-module \(M\), any \(\mathfrak{g}\)-homomorphism \(f :
-M(\lambda) \to M\) is determined by the image of \(m^+\). Indeed, \(f(u \cdot
-m^+) = u \cdot f(m^+)\) for all \(u \in \mathcal{U}(\mathfrak{g})\). In
-addition, it is clear that
-\[
-  H \cdot f(m^+) = f(H \cdot m^+) = f(\lambda(H) m^+) = \lambda(H) f(m^+)
-\]
-for all \(H \in \mathfrak{h}\) and, similarly, \(X \cdot f(m^+) = 0\) for all
-\(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\). This leads us to the
-universal property of \(M(\lambda)\).
-
-\begin{definition}
-  Let \(M\) be a \(\mathfrak{g}\)-module and \(m \in M\). If \(X \cdot m = 0\)
-  for all \(X \in \mathfrak{g}_\alpha\), \(\alpha \in \Delta^+\), then \(m\) is
-  called \emph{a singular vector of \(M\)}.
-\end{definition}
-
-\begin{proposition}
-  Let \(M\) be a \(\mathfrak{g}\)-module and \(m \in M_\lambda\) be a singular
-  vector. Then there exists a unique \(\mathfrak{g}\)-homomorphism \(f :
-  M(\lambda) \to M\) such that \(f(m^+) = m\). Furthermore, all homomorphisms
-  \(M(\lambda) \to M\) are given in this fashion.
-  \[
-    \operatorname{Hom}_{\mathfrak{g}}(M(\lambda), M)
-    \cong \{ m \in M_\lambda : m \ \text{is singular}\}
-  \]
-\end{proposition}
-
-\begin{proof}
-  The result follows directly from Proposition~\ref{thm:frobenius-reciprocity}.
-  Indeed, by the Frobenius Reciprocity Theorem, a \(\mathfrak{g}\)-homomorphism
-  \(f : M(\lambda) \to M\) is the same as a \(\mathfrak{b}\)-homomorphism \(g :
-  K m^+ \to M = \operatorname{Res}_{\mathfrak{b}}^{\mathfrak{g}} M\). More
-  specifically, given a \(\mathfrak{b}\)-homomorphism \(g : K m^+ \to M\),
-  there exists a unique \(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\)
-  such that \(f(u \otimes m) = u \cdot g(m)\) for all \(m \in K m^+\), and all
-  \(\mathfrak{g}\)-homomorphism \(M(\lambda) \to M\) arise in this fashion.
-
-  Any \(K\)-linear map \(g : K m^+ \to M\) is determined by the image
-  \(g(m^+)\) of \(m^+\) and such an image is a singular vector if, and only if
-  \(g\) is a \(\mathfrak{b}\)-homomorphism.
-\end{proof}
-
-Notice that any highest weight vector is a singular vector. Now suppose \(M\)
-is a simple finite-dimensional \(\mathfrak{g}\)-module of highest weight vector
-\(m \in M_\lambda\). By the last proposition, there is a
-\(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\) such that \(f(m^+) = m\).
-Since \(M\) is simple, \(M = \mathcal{U}(\mathfrak{g}) \cdot m\) and therefore
-\(M \cong \mfrac{M(\lambda)}{\ker f}\). It then follows from the simplicity of
-\(M\) that \(\ker f \subset M(\lambda)\) is a maximal
-\(\mathfrak{g}\)-submodule. Maximal submodules of Verma modules are thus of
-primary interest to us. As it turns out, these can be easily classified.
-
-\begin{proposition}\label{thm:max-verma-submod-is-weight}
-  Every submodule \(N \subset M(\lambda)\) is the direct sum of its weight
-  spaces. In particular, \(M(\lambda)\) has a unique maximal submodule
-  \(N(\lambda)\) and a unique simple quotient \(L(\lambda) =
-  \sfrac{M(\lambda)}{N(\lambda)}\).
-\end{proposition}
-
-\begin{proof}
-  Let \(N \subset M(\lambda)\) be a submodule and take any nonzero \(n \in N\).
-  Because of Proposition~\ref{thm:verma-is-weight-mod}, we know there are
-  \(\mu_1, \ldots, \mu_r \in \mathfrak{h}^*\) and nonzero \(m_i \in
-  M(\lambda)_{\mu_i}\) such that \(n = m_1 + \cdots + m_r\). We want to show
-  \(m_i \in N\) for all \(i\).
-
-  Fix some \(H_2 \in \mathfrak{h}\) such that \(\mu_1(H_2) \ne \mu_2(H_2)\).
-  Then
-  \[
-    m_1
-    - \frac{(\mu_3 - \mu_1)(H_2)}{(\mu_2 - \mu_1)(H_2)} \cdot m_3
-    - \cdots
-    - \frac{(\mu_r - \mu_1)(H_2)}{(\mu_2 - \mu_1)(H_2)} \cdot m_r
-    = \left( 1 - \frac{H_2 - \mu_1(H_2)}{(\mu_2 - \mu_1)(H_2)} \right) \cdot n
-    \in N
-  \]
-
-  Now take \(H_3 \in \mathfrak{h}\) such that \(\mu_1(H_3) \ne \mu_3(H_3)\). By
-  applying the same procedure again we get
-  \begin{multline*}
-    m_1
-    -
-    \frac{(\mu_4 - \mu_3)(H_3) \cdot (\mu_4 - \mu_1)(H_2)}
-         {(\mu_3 - \mu_1)(H_3) \cdot (\mu_2 - \mu_1)(H_2)} \cdot m_4
-    - \cdots -
-    \frac{(\mu_r - \mu_3)(H_3) \cdot (\mu_r - \mu_1)(H_2)}
-         {(\mu_3 - \mu_1)(H_3) \cdot (\mu_2 - \mu_1)(H_2)} \cdot m_r \\
-    =
-    \left(1 - \frac{H_3 - \mu_1(H_3)}{(\mu_3 - \mu_1)(H_3)} \right)
-    \left(1 - \frac{H_2 - \mu_1(H_2)}{(\mu_2 - \mu_1)(H_2)} \right) \cdot n
-    \in N
-  \end{multline*}
-
-  By applying the same procedure over and over again we can see that \(m_1 = u
-  \cdot n \in N\) for some \(u \in \mathcal{U}(\mathfrak{g})\). Furthermore, if
-  we reproduce all this for \(m_2 + \cdots + m_r = n - m_1 \in N\) we get that
-  \(m_2 \in N\). All in all we find \(m_1, \ldots, m_r \in N\). Hence
-  \[
-    N = \bigoplus_\mu N_\mu = \bigoplus_\mu M(\lambda)_\mu \cap N
-  \]
-
-  Since \(M(\lambda) = \mathcal{U}(\mathfrak{g}) \cdot m^+\), if \(N\) is a
-  proper submodule then \(m^+ \notin N\). Hence any proper submodule lies in
-  the sum of weight spaces other than \(M(\lambda)_\lambda\), so the sum
-  \(N(\lambda)\) of all such submodules is still proper. This implies
-  \(N(\lambda)\) is the unique maximal submodule of \(M(\lambda)\) and
-  \(L(\lambda) = \sfrac{M(\lambda)}{N(\lambda)}\) is its unique simple
-  quotient.
-\end{proof}
-
-\begin{example}\label{ex:sl2-verma-quotient}
-  If \(\mathfrak{g} = \mathfrak{sl}_2(K)\) and \(\lambda : h \mapsto 2\), we
-  can see from Example~\ref{ex:sl2-verma} that \(N(\lambda) = \bigoplus_{k \ge
-  3} K f^k \cdot m^+\), so that \(L(\lambda)\) is the \(3\)-dimensional simple
-  \(\mathfrak{sl}_2(K)\)-module -- i.e. the finite-dimensional simple module
-  with highest weight \(\lambda\) constructed in chapter~\ref{ch:sl3}.
-\end{example}
-
-All its left to prove the Highest Weight Theorem is verifying that the
-situation encountered in Example~\ref{ex:sl2-verma-quotient} holds for any
-\(\lambda \in P\). In other words, we need to show\dots
-
-\begin{proposition}\label{thm:verma-is-finite-dim}
-  If \(\mathfrak{g}\) is semisimple and \(\lambda\) is dominant integral then
-  the unique simple quotient \(L(\lambda)\) of \(M(\lambda)\) is
-  finite-dimensional.
-\end{proposition}
-
-The proof of Proposition~\ref{thm:verma-is-finite-dim} is very technical and we
-won't include it here, but the idea behind it is to show that the set of
-weights of \(L(\lambda)\) is stable under the natural action of the Weyl group
-\(W\) on \(\mathfrak{h}^*\). One can then show that the every weight
-of \(L(\lambda)\) is conjugate to a single dominant integral weight of
-\(L(\lambda)\), and that the set of dominant integral weights of \(L(\lambda)\)
-is finite. Since \(W\) is finitely generated, this implies the set of
-weights of the unique simple quotient of \(M(\lambda)\) is finite. But
-each weight space is finite-dimensional. Hence so is the simple quotient
-\(L(\lambda)\).
-
-We refer the reader to \cite[ch. 21]{humphreys} for further details. We are now
-ready to prove the Highest Weight Theorem.
-
-\begin{proof}[Proof of Theorem~\ref{thm:dominant-weight-theo}]
-  We begin with the ``existence'' part of the theorem by showing that
-  \(L(\lambda)\) is indeed a finite-dimensional simple module whose
-  highest-weight is \(\lambda\). It suffices to show that the highest weight of
-  \(L(\lambda)\) is \(\lambda\). We have already seen that \(m^+ \in
-  M(\lambda)_\lambda\) is a highest weight vector. Now since \(m^+\) lies
-  outside of the maximal submodule of \(M(\lambda)\), the projection \(m^+ +
-  N(\lambda) \in L(\lambda)\) is nonzero.
-
-  We now claim that \(m^+ + N(\lambda) \in L(\lambda)_\lambda\). Indeed,
-  \[
-    H \cdot (m^+ + N(\lambda))
-    = H \cdot m^+ + N(\lambda)
-    = \lambda(H) (m^+ + N(\lambda))
-  \]
-  for all \(H \in \mathfrak{h}\). Hence \(\lambda\) is a weight of
-  \(L(\lambda)\), with weight vector \(m^+ + N(\lambda)\). Finally, we remark
-  that \(\lambda\) is the highest weight of \(L(\lambda)\), for if this was not
-  the case we could find a weight \(\mu\) of \(M(\lambda)\) with \(\mu \succ
-  \lambda\).
-
-  Now suppose \(M\) is some other finite-dimensional simple
-  \(\mathfrak{g}\)-module with highest weight vector \(m \in M_\lambda\). By
-  the universal property of the Verma module, there is a
-  \(\mathfrak{g}\)-homomorphism \(f : M(\lambda) \to M\) such that \(f(m^+) =
-  m\). As indicated before, since \(M\) is simple, \(M =
-  \mathcal{U}(\mathfrak{g}) \cdot m\) and therefore \(f\) is surjective. It
-  then follows \(M \cong \mfrac{M(\lambda)}{\ker f}\).
-
-  Since \(M\) is simple, \(\ker f \subset M(\lambda)\) is maximal and therefore
-  \(\ker f = N(\lambda)\). In other words, \(M \cong \mfrac{M(\lambda)}{\ker f}
-  = L(\lambda)\). We are done.
-\end{proof}
-
-We should point out that Proposition~\ref{thm:verma-is-finite-dim} fails for
-non-dominant \(\lambda \in P\). While \(\lambda\) is always a maximal weight of
-\(M(\lambda)\), one can show that if \(\lambda \in P\) is not dominant then
-\(N(\lambda) = 0\) and \(M(\lambda)\) is simple. For instance, if
-\(\mathfrak{g} = \mathfrak{sl}_2(K)\) and \(\lambda : h \mapsto -2\) then the
-action of \(\mathfrak{g}\) on \(M(\lambda)\) is given by
-\begin{center}
-  \begin{tikzcd}
-    \cdots             \rar[bend left=60]{-20}
-    & M(\lambda)_{-8}  \rar[bend left=60]{-12} \lar[bend left=60]{1}
-    & M(\lambda)_{-6}  \rar[bend left=60]{-6}  \lar[bend left=60]{1}
-    & M(\lambda)_{-4}  \rar[bend left=60]{-2}  \lar[bend left=60]{1}
-    & M(\lambda)_{-2}                          \lar[bend left=60]{1}
-  \end{tikzcd},
-\end{center}
-so we can see that \(M(\lambda)\) has no proper submodules. Verma modules can
-thus serve as examples of infinite-dimensional simple modules. Our next
-question is: what are \emph{all} the infinite-dimensional simple
-\(\mathfrak{g}\)-modules?
diff --git /dev/null b/sections/simple-weight.tex
@@ -0,0 +1,1611 @@
+\chapter{Simple Weight Modules}\label{ch:mathieu}
+
+In this chapter we will expand our results on finite-dimensional simple modules
+of semisimple Lie algebras by considering \emph{infinite-dimensional}
+\(\mathfrak{g}\)-modules, which introduces numerous complications to our
+analysis. 
+
+For instance, in the infinite-dimensional setting we can no longer take
+complete-reducibility for granted. Indeed, we have seen that even if
+\(\mathfrak{g}\) is a semisimple Lie algebra, there are infinite-dimensional
+\(\mathfrak{g}\)-modules which are not semisimple. For a counterexample look no
+further than Example~\ref{ex:regular-mod-is-not-semisimple}: the regular
+\(\mathfrak{g}\)-module \(\mathcal{U}(\mathfrak{g})\) is never semisimple.
+Nevertheless, for simplicity -- or shall we say \emph{semisimplicity} -- we
+will focus exclusively on \emph{semisimple} \(\mathfrak{g}\)-modules. Our
+strategy is, once again, that of classifying simple modules. The regular
+\(\mathfrak{g}\)-module hides further unpleasant surprises, however: recall
+from Example~\ref{ex:regular-mod-is-not-weight-mod} that
+\[
+  \bigoplus_\lambda \mathcal{U}(\mathfrak{g})_\lambda
+  = 0
+  \subsetneq \mathcal{U}(\mathfrak{g})
+\]
+and the weight space decomposition fails for \(\mathcal{U}(\mathfrak{g})\).
+
+Indeed, our proof of the weight space decomposition in the finite-dimensional
+case relied heavily in the simultaneous diagonalization of commuting operators
+in a finite-dimensional space. Even if we restrict ourselves to simple modules,
+there is still a diverse spectrum of counterexamples to
+Corollary~\ref{thm:finite-dim-is-weight-mod} in the infinite-dimensional
+setting. For instance, any \(\mathfrak{g}\)-module \(M\) whose restriction to
+\(\mathfrak{h}\) is a free module satisfies \(M_\lambda = 0\) for all
+\(\lambda\) as in Example~\ref{ex:regular-mod-is-not-weight-mod}. These are
+called \emph{\(\mathfrak{h}\)-free \(\mathfrak{g}\)-modules}, and rank \(1\)
+simple \(\mathfrak{h}\)-free \(\mathfrak{sp}_{2 n}(K)\)-modules where first
+classified by Nilsson in \cite{nilsson}. Dimitar's construction of the so
+called \emph{exponential tensor \(\mathfrak{sl}_n(K)\)-modules} in
+\cite{dimitar-exp} is also an interesting source of counterexamples.
+
+Since the weight space decomposition was perhaps the single most instrumental
+ingredient of our previous analysis, it is only natural to restrict ourselves
+to the case it holds. This brings us to the following definition.
+
+\begin{definition}\label{def:weight-mod}\index{\(\mathfrak{g}\)-module!weight modules}\index{weights!weight modules}\index{\(\mathfrak{g}\)-module!(essential) support}
+  A \(\mathfrak{g}\)-module \(M\) is called a \emph{weight
+  \(\mathfrak{g}\)-module} if \(M = \bigoplus_{\lambda \in \mathfrak{h}^*}
+  M_\lambda\) and \(\dim M_\lambda < \infty\) for all \(\lambda \in
+  \mathfrak{h}^*\). The \emph{support of \(M\)} is the set
+  \(\operatorname{supp} M = \{\lambda \in \mathfrak{h}^* : M_\lambda \ne 0\}\).
+\end{definition}
+
+\begin{example}
+  Corollary~\ref{thm:finite-dim-is-weight-mod} is equivalent to the fact that
+  every finite-dimensional module of a semisimple Lie algebra is a weight
+  module. More generally, every finite-dimensional simple module of a reductive
+  Lie algebra is a weight module.
+\end{example}
+
+\begin{example}\label{ex:reductive-alg-equivalence}
+  We have seen that every finite-dimensional \(\mathfrak{g}\)-module is a
+  weight module for semisimple \(\mathfrak{g}\). In particular, if
+  \(\mathfrak{g}\) is finite-dimensional then the adjoint
+  \(\mathfrak{g}\)-module \(\mathfrak{g}\) is a weight module. More generally,
+  a finite-dimensional Lie algebra \(\mathfrak{g}\) is reductive if, and only
+  if the adjoint \(\mathfrak{g}\)-module \(\mathfrak{g}\) is a weight module,
+  in which case its weight spaces are given by the root spaces of
+  \(\mathfrak{g}\)
+\end{example}
+
+\begin{example}\label{ex:submod-is-weight-mod}
+  Proposition~\ref{thm:verma-is-weight-mod} and
+  Proposition~\ref{thm:max-verma-submod-is-weight} imply that the Verma module
+  \(M(\lambda)\) and its maximal submodule are both weight modules. In
+  fact, the proof of Proposition~\ref{thm:max-verma-submod-is-weight} is
+  actually a proof of the fact that every submodule \(N \subset M\) of
+  a weight module \(M\) is a weight module, and \(N_\lambda = M_\lambda \cap
+  N\) for all \(\lambda \in \mathfrak{h}^*\).
+\end{example}
+
+\begin{example}\label{ex:quotient-is-weight-mod}
+  Given a weight module \(M\), a submodule \(N \subset M\) and \(\lambda \in
+  \mathfrak{h}^*\), it is clear that \(\mfrac{M_\lambda}{N} \subset
+  \left(\mfrac{M}{N}\right)_\lambda\). In addition, \(\mfrac{M}{N} =
+  \bigoplus_{\lambda \in \mathfrak{h}^*} \mfrac{M_\lambda}{N}\). Hence
+  \(\mfrac{M}{N}\) is weight \(\mathfrak{g}\)-module with
+  \(\left(\mfrac{M}{N}\right)_\lambda = \mfrac{M_\lambda}{N} \cong
+  \mfrac{M_\lambda}{N_\lambda}\).
+\end{example}
+
+\begin{example}\label{ex:tensor-prod-of-weight-is-weight}
+  Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras, \(M_1\) be a
+  weight \(\mathfrak{g}_1\)-module and \(M_2\) a weight
+  \(\mathfrak{g}_2\)-module. Recall from Example~\ref{ex:cartan-direct-sum}
+  that if \(\mathfrak{h}_i \subset \mathfrak{g}_i\) are Cartan subalgebras then
+  \(\mathfrak{h} = \mathfrak{h}_1 \oplus \mathfrak{h}_2\) is a Cartan
+  subalgebra of \(\mathfrak{g} = \mathfrak{g}_1 \oplus \mathfrak{g}_2\) with
+  \(\mathfrak{h}^* = \mathfrak{h}_1^* \oplus \mathfrak{h}_2^*\). In this
+  setting, one can readily check that \(M_1 \otimes M_2\) is a weight
+  \(\mathfrak{g}\)-module with
+  \[
+    (M_1 \otimes M_2)_{\lambda_1 + \lambda_2}
+    = (M_1)_{\lambda_1} \otimes (M_2)_{\lambda_2}
+  \]
+  for all \(\lambda_i \in \mathfrak{h}_i^*\) and \(\operatorname{supp} M_1
+  \otimes M_2 = \operatorname{supp} M_1 \oplus \operatorname{supp} M_2 = \{
+  \lambda_1 + \lambda_2 : \lambda_i \in \operatorname{supp} M_i \subset
+  \mathfrak{h}_i^*\}\).
+\end{example}
+
+\begin{example}\label{thm:simple-weight-mod-is-tensor-prod}
+  Let \(\mathfrak{g} = \mathfrak{z} \oplus \mathfrak{s}_1 \oplus \cdots \oplus
+  \mathfrak{s}_r\) be a reductive Lie algebra, where \(\mathfrak{z}\) is the
+  center of \(\mathfrak{g}\) and \(\mathfrak{s}_1, \ldots, \mathfrak{s}_r\) are
+  its simple components. As in
+  Example~\ref{ex:all-simple-reps-are-tensor-prod}, any simple weight
+  \(\mathfrak{g}\)-module \(M\) can be decomposed as
+  \[
+    M \cong Z \otimes M_1 \otimes \cdots \otimes M_r
+  \]
+  where \(Z\) is a \(1\)-dimensional representation of \(\mathfrak{z}\) and
+  \(M_i\) is a simple weight \(\mathfrak{s}_i\)-module. The modules \(Z\) and
+  \(M_i\) are uniquely determined up to isomorphism.
+\end{example}
+
+\begin{example}\label{ex:adjoint-action-in-universal-enveloping-is-weight}
+  We would like to show that the requirement of finite-dimensionality in
+  Definition~\ref{def:weight-mod} is not redundant. Let \(\mathfrak{g}\) be a
+  finite-dimensional reductive Lie algebra and consider the adjoint
+  \(\mathfrak{g}\)-module \(\mathcal{U}(\mathfrak{g})\) -- where \(X \in
+  \mathfrak{g}\) acts by taking commutators. Given \(\alpha \in Q\), a simple
+  computation shows \(K \langle X_1 \cdots X_n H_1 \cdots H_m : X_i \in
+  \mathfrak{g}_{\alpha_i}, H_i \in \mathfrak{h}, \alpha_i \in \Delta, \alpha =
+  \alpha_1 + \cdots + \alpha_n \rangle \subset
+  \mathcal{U}(\mathfrak{g})_\alpha\). The PBW Theorem and
+  Example~\ref{ex:reductive-alg-equivalence} thus imply that
+  \(\mathcal{U}(\mathfrak{g}) = \bigoplus_{\alpha \in Q}
+  \mathcal{U}(\mathfrak{g})_\alpha\) where \(\mathcal{U}(\mathfrak{g})_\alpha =
+  K \langle X_1 \cdots X_n H_1 \cdots H_m : X_i \in \mathfrak{g}_{\alpha_i},
+  H_i \in \mathfrak{h}, \alpha_i \in \Delta, \alpha = \alpha_1 + \cdots +
+  \alpha_n \rangle\). However, \(\dim \mathcal{U}(\mathfrak{g})_\alpha =
+  \infty\). For instance, \(\mathcal{U}(\mathfrak{g})_0\) is \emph{precisely}
+  the commutator of \(\mathfrak{h}\) in \(\mathcal{U}(\mathfrak{g})\), which
+  contains \(\mathcal{U}(\mathfrak{h})\) and is therefore infinite-dimensional.
+\end{example}
+
+\begin{note}
+  We should stress that the weight spaces \(M_\lambda \subset M\) of a given
+  weight \(\mathfrak{g}\)-module \(M\) are \emph{not}
+  \(\mathfrak{g}\)-submodules. Nevertheless, \(M_\lambda\) is a
+  \(\mathfrak{h}\)-submodule. More generally, \(M_\lambda\) is a
+  \(\mathcal{U}(\mathfrak{g})_0\)-submodule, where
+  \(\mathcal{U}(\mathfrak{g})_0\) is the centralizer of \(\mathfrak{h}\) in
+  \(\mathcal{U}(\mathfrak{g})\) -- which coincides with the weight space of \(0
+  \in \mathfrak{h}^*\) in the adjoint \(\mathfrak{g}\)-module
+  \(\mathcal{U}(\mathfrak{g})\), as seen in
+  Example~\ref{ex:adjoint-action-in-universal-enveloping-is-weight}.
+\end{note}
+
+A particularly well behaved class of examples are the so called
+\emph{bounded} modules.
+
+\begin{definition}\index{\(\mathfrak{g}\)-module!bounded modules}\index{\(\mathfrak{g}\)-module!(essential) support}
+  A weight \(\mathfrak{g}\)-module \(M\) is called \emph{bounded} if \(\dim
+  M_\lambda\) is bounded. The lowest upper bound \(\deg M\) for \(\dim
+  M_\lambda\) is called \emph{the degree of \(M\)}. The \emph{essential
+  support} of \(M\) is the set \(\operatorname{supp}_{\operatorname{ess}} M =
+  \{ \lambda \in \mathfrak{h}^* : \dim M_\lambda = \deg M \}\).
+\end{definition}
+
+\begin{example}\label{ex:supp-ess-of-tensor-is-product}
+  Let \(\mathfrak{g}_1\) and \(\mathfrak{g}_2\) be Lie algebras with Cartan
+  subalgebras \(\mathfrak{h}_i \subset \mathfrak{g}_i\) and take \(\mathfrak{g}
+  = \mathfrak{g}_1 \oplus \mathfrak{g}_2\). Given bounded
+  \(\mathfrak{g}_i\)-modules \(M_i\), it follows from
+  Example~\ref{ex:tensor-prod-of-weight-is-weight} that \(M_1 \otimes M_2\) is
+  a bounded \(\mathfrak{g}\)-module with \(\deg M_1 \otimes M_2 = \deg M_1
+  \cdot \deg M_2\) and 
+  \[
+    \operatorname{supp}_{\operatorname{ess}} M_1 \otimes M_2
+    = \operatorname{supp}_{\operatorname{ess}} M_1 \oplus
+      \operatorname{supp}_{\operatorname{ess}} M_2
+    = \{
+        \lambda_1 + \lambda_2 : \lambda_i \in
+        \operatorname{supp}_{\operatorname{ess}} M_i \subset \mathfrak{h}_i^*
+      \}
+  \]
+\end{example}
+
+\begin{example}\label{ex:laurent-polynomial-mod}
+  There is a natural action of \(\mathfrak{sl}_2(K)\) on the space \(K[x,
+  x^{-1}]\) of Laurent polynomials, given by the formulas in
+  (\ref{eq:laurent-polynomials-cusp-mod}). One can quickly verify \(K[x,
+  x^{-1}]_{2 k} = K x^k\) and \(K[x, x^{-1}]_\lambda = 0\) for any \(\lambda
+  \notin 2 \mathbb{Z}\), so that \(K[x, x^{-1}] = \bigoplus_{k \in \mathbb{Z}}
+  K x^k\) is a degree \(1\) bounded weight \(\mathfrak{sl}_2(K)\)-module. It
+  follows from the remark at the end of Example~\ref{ex:submod-is-weight-mod}
+  that any nonzero submodule \(N \subset K[x, x^{-1}]\) must contain a
+  monomial \(x^k\). But since the operators \(-\frac{\mathrm{d}}{\mathrm{d}x} +
+  \frac{x^{-1}}{2}, x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \frac{x}{2} : K[x,
+  x^{-1}] \to K[x, x^{-1}]\) are both injective, this implies all other
+  monomials can be found in \(N\) by successively applying \(f\) and \(e\).
+  Hence \(N = K[x, x^{-1}]\) and \(K[x, x^{-1}]\) is a simple module.
+  \begin{align}\label{eq:laurent-polynomials-cusp-mod}
+    e \cdot p
+    & = \left( x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \frac{x}{2} \right) p &
+    f \cdot p
+    & = \left(- \frac{\mathrm{d}}{\mathrm{d}x} + \frac{x^{-1}}{2} \right) p &
+    h \cdot p
+    & = 2 x \frac{\mathrm{d}}{\mathrm{d}x} p
+  \end{align}
+\end{example}
+
+Notice that the support of \(K[x, x^{-1}]\) is the trivial \(2
+\mathbb{Z}\)-coset \(0 + 2 \mathbb{Z}\). This is representative of the general
+behavior in the following sense: if \(M\) is a simple weight
+\(\mathfrak{g}\)-module, since \(M[\lambda] = \bigoplus_{\alpha \in Q}
+M_{\lambda + \alpha}\) is stable under the action of \(\mathfrak{g}\) for all
+\(\lambda \in \mathfrak{h}^*\), \(\bigoplus_{\alpha \in Q} M_{\lambda +
+\alpha}\) is either \(0\) or all of \(M\). In other words, the support of a
+simple weight module is always contained in a single \(Q\)-coset.
+
+However, the behavior of \(K[x, x^{-1}]\) deviates from that of an arbitrary
+bounded \(\mathfrak{g}\)-module in the sense its essential support is
+precisely the entire \(Q\)-coset it inhabits -- i.e.
+\(\operatorname{supp}_{\operatorname{ess}} K[x, x^{-1}] = 2 \mathbb{Z}\). This
+isn't always the case. Nevertheless, in general we find\dots
+
+\begin{proposition}\label{thm:ess-supp-is-zariski-dense}
+  Let \(\mathfrak{g}\) be a finite-dimensional semisimple Lie algebra and \(M\)
+  be a simple infinite-dimensional bounded \(\mathfrak{g}\)-module. The
+  essential support \(\operatorname{supp}_{\operatorname{ess}} M\) is
+  Zariski-dense\footnote{Any choice of basis for $\mathfrak{h}^*$ induces a
+  $K$-linear isomorphism $\mathfrak{h}^* \isoto K^n$. In particular, a choice
+  of basis induces a unique topology in $\mathfrak{h}^*$ such that the map
+  $\mathfrak{h}^* \to K^n$ is a homeomorphism onto $K^n$ with the Zariski
+  topology. Any two basis induce the same topology in $\mathfrak{h}^*$, which
+  we call \emph{the Zariski topology of $\mathfrak{h}^*$}.} in
+  \(\mathfrak{h}^*\).
+\end{proposition}
+
+This proof was deemed too technical to be included in here, but see Proposition
+3.5 of \cite{mathieu} for the case where \(\mathfrak{g} = \mathfrak{s}\) is a
+simple Lie algebra. The general case then follows from
+Example~\ref{thm:simple-weight-mod-is-tensor-prod},
+Example~\ref{ex:supp-ess-of-tensor-is-product} and the asserting that the
+product of Zariski-dense subsets in \(K^n\) and \(K^m\) is Zariski-dense in
+\(K^{n + m} = K^n \times K^m\).
+
+We now begin a systematic investigation of the problem of classifying the
+infinite-dimensional simple weight modules of a given Lie algebra
+\(\mathfrak{g}\). As in the previous chapter, let \(\mathfrak{g}\) be a
+finite-dimensional semisimple Lie algebra. As a first approximation of a
+solution to our problem, we consider the induction functors
+\(\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}} :
+\mathfrak{p}\text{-}\mathbf{Mod} \to \mathfrak{g}\text{-}\mathbf{Mod}\), where
+\(\mathfrak{p} \subset \mathfrak{g}\) is some subalgebra. 
+
+% TODOO: Are you sure that these are indeed the weight spaces of the induced
+% module? Check this out?
+These functors have already proved themselves a powerful tool for constructing
+modules in the previous chapters. Our first observation is that if
+\(\mathfrak{p} \subset \mathfrak{g}\) contains the Borel subalgebra
+\(\mathfrak{b}\) then \(\mathfrak{h}\) is a Cartan subalgebra of
+\(\mathfrak{p}\) and \((\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}}
+M)_\lambda = \mathcal{U}(\mathfrak{g}) \otimes_{\mathcal{U}(\mathfrak{p})}
+M_\lambda\) for all \(\lambda \in \mathfrak{h}^*\). In particular,
+\(\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}}\) takes weight
+\(\mathfrak{p}\)-modules to weight \(\mathfrak{g}\)-modules. This leads us to
+the following definition.
+
+\begin{definition}\index{Lie subalgebra!parabolic subalgebra}
+  A subalgebra \(\mathfrak{p} \subset \mathfrak{g}\) is called \emph{parabolic}
+  if \(\mathfrak{b} \subset \mathfrak{p}\).
+\end{definition}
+
+% TODOO: Why is the fact that p is not reductive relevant?? Why do we need to
+% look at the quotient by nil(p)??
+Parabolic subalgebras thus give us a process for constructing weight
+\(\mathfrak{g}\)-modules from modules of smaller (parabolic) subalgebras. Our
+hope is that by iterating this process again and again we can get a large class
+of simple weight \(\mathfrak{g}\)-modules. However, there is a small catch: a
+parabolic subalgebra \(\mathfrak{p} \subset \mathfrak{g}\) needs not to be
+reductive. We can get around this limitation by modding out by
+\(\mathfrak{nil}(\mathfrak{p})\) and noticing that
+\(\mathfrak{nil}(\mathfrak{p})\) acts trivially in any weight
+\(\mathfrak{p}\)-module \(M\). By applying the universal property of quotients
+we can see that \(M\) has the natural structure of a
+\(\mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}\)-module, which is always
+a reductive algebra.
+\begin{center}
+  \begin{tikzcd}
+    \mathfrak{p}                                       \rar \dar          &
+    \mathfrak{gl}(M)                                                      \\
+    \mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})} \arrow[dotted]{ur} &
+  \end{tikzcd}
+\end{center}
+
+Let \(\mathfrak{p}\) be a parabolic subalgebra and \(M\) be a simple
+weight \(\mathfrak{p}\)-module. We should point out that while
+\(\operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}} M\) is a weight
+\(\mathfrak{g}\)-module, it isn't necessarily simple. Nevertheless, we can
+use it to produce a simple weight \(\mathfrak{g}\)-module via a
+construction very similar to that of Verma modules.
+
+\begin{definition}\index{\(\mathfrak{g}\)-module!(generalized) Verma modules}
+  Given any \(\mathfrak{p}\)-module \(M\), the module \(M_{\mathfrak{p}}(M) =
+  \operatorname{Ind}_{\mathfrak{p}}^{\mathfrak{g}} M\) is called \emph{a
+  generalized Verma module}.
+\end{definition}
+
+\begin{proposition}\label{thm:generalized-verma-has-simple-quotient}
+  Given a simple \(\mathfrak{p}\)-module \(M\), the generalized Verma module
+  \(M_{\mathfrak{p}}(M)\) has a unique maximal \(\mathfrak{p}\)-submodule
+  \(N_{\mathfrak{p}}(M)\) and a unique irreducible quotient
+  \(L_{\mathfrak{p}}(M) = \mfrac{M_{\mathfrak{p}}(M)}{N_{\mathfrak{p}}(M)}\).
+  The irreducible quotient \(L_{\mathfrak{p}}(M)\) is a weight module.
+\end{proposition}
+
+The proof of Proposition~\ref{thm:generalized-verma-has-simple-quotient} is
+entirely analogous to that of Proposition~\ref{thm:max-verma-submod-is-weight}.
+This leads us to the following definitions.
+
+\begin{definition}\index{\(\mathfrak{g}\)-module!parabolic induced modules}\index{\(\mathfrak{g}\)-module!cuspidal modules}
+  A \(\mathfrak{g}\)-module is called \emph{parabolic induced} if it is
+  isomorphic to \(L_{\mathfrak{p}}(M)\) for some proper parabolic subalgebra
+  \(\mathfrak{p} \subsetneq \mathfrak{g}\) and some \(\mathfrak{p}\)-module
+  \(M\). An \emph{simple cuspidal \(\mathfrak{g}\)-module} is a simple
+  \(\mathfrak{g}\)-module which is \emph{not} parabolic induced.
+\end{definition}
+
+Since every weight \(\mathfrak{p}\)-module \(M\) is an
+\(\mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}\)-module, it makes sense
+to call \(M\) \emph{cuspidal} if it is a cuspidal
+\(\mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}\)-module. The first
+breakthrough regarding our classification problem was given by Fernando in his
+now infamous paper \citetitle{fernando} \cite{fernando}, where he proved that
+every simple weight \(\mathfrak{g}\)-module is parabolic induced. In other
+words\dots
+
+\begin{theorem}[Fernando]
+  Any simple weight \(\mathfrak{g}\)-module is isomorphic to
+  \(L_{\mathfrak{p}}(M)\) for some parabolic subalgebra \(\mathfrak{p} \subset
+  \mathfrak{g}\) and some simple cuspidal \(\mathfrak{p}\)-module \(M\).
+\end{theorem}
+
+We should point out that the relationship between simple weight
+\(\mathfrak{g}\)-modules and pairs \((\mathfrak{p}, M)\) -- where
+\(\mathfrak{p}\) is some parabolic subalgebra and \(M\) is a simple cuspidal
+\(\mathfrak{p}\)-module -- is not one-to-one. Nevertheless, this relationship
+is well understood. Namely, Fernando himself established\dots
+
+\begin{proposition}[Fernando]
+  Given a parabolic subalgebra \(\mathfrak{p} \subset \mathfrak{g}\), there
+  exists a basis \(\Sigma\) for \(\Delta\) such that \(\Sigma \subset
+  \Delta_{\mathfrak{p}} \subset \Delta\), where \(\Delta_{\mathfrak{p}}\)
+  denotes the set of roots of \(\mathfrak{p}\). Furthermore, if \(\mathfrak{p}'
+  \subset \mathfrak{g}\) is another parabolic subalgebra, \(M\) is a simple
+  cuspidal \(\mathfrak{p}\)-module and \(N\) is a simple cuspidal
+  \(\mathfrak{p}'\)-module then \(L_{\mathfrak{p}}(M) \cong
+  L_{\mathfrak{p}'}(N)\) if, and only if \(\mathfrak{p}' =
+  \twisted{\mathfrak{p}}{\sigma}\) and \(M \cong \twisted{N}{\sigma}\) for
+  some\footnote{Here $\twisted{\mathfrak{p}}{\sigma}$ denotes the image of
+  $\mathfrak{p}$ under the automorphism of $\sigma : \mathfrak{g} \to
+  \mathfrak{g}$ given by the canonical action of $W$ on $\mathfrak{g}$ and
+  $\twisted{N}{\sigma}$ is the $\mathfrak{p}$-module given by composing the map
+  $\mathfrak{p}' \to \mathfrak{gl}(N)$ with the restriction
+  $\sigma\!\restriction_{\mathfrak{p}} : \mathfrak{p} \to \mathfrak{p}'$.}
+  \(\sigma \in W_M\), where
+  \[
+    W_M
+    = \langle
+      \sigma_\beta : \beta \in \Sigma, H_\beta + \mathfrak{nil}(\mathfrak{p})
+      \ \text{is central in}\ \mfrac{\mathfrak{p}}{\mathfrak{nil}(\mathfrak{p})}
+      \ \text{and}\ H_\beta\ \text{acts on \(M\) as a positive integer}
+      \rangle
+    \subset W
+  \]
+\end{proposition}
+
+\begin{note}
+  The definition of the subgroup \(W_M \subset W\) is independent of the choice
+  of basis \(\Sigma\).
+\end{note}
+
+As a first consequence of Fernando's Theorem, we provide two alternative
+characterizations of cuspidal modules.
+
+\begin{corollary}[Fernando]\label{thm:cuspidal-mod-equivs}
+  Let \(M\) be a simple weight \(\mathfrak{g}\)-module. The following
+  conditions are equivalent.
+  \begin{enumerate}
+    \item \(M\) is cuspidal.
+    \item \(F_\alpha\) acts injectively on \(M\) for all
+      \(\alpha \in \Delta\) -- this is what is usually referred
+      to as a \emph{dense} module in the literature.
+    \item The support of \(M\) is precisely one \(Q\)-coset -- this is
+      what is usually referred to as a \emph{torsion-free} module in the
+      literature.
+  \end{enumerate}
+\end{corollary}
+
+\begin{example}
+  As noted in Example~\ref{ex:laurent-polynomial-mod}, the element \(f \in
+  \mathfrak{sl}_2(K)\) acts injectively on the space of Laurent polynomials.
+  Hence \(K[x, x^{-1}]\) is a cuspidal \(\mathfrak{sl}_2(K)\)-module.
+\end{example}
+
+Having reduced our classification problem to that of classifying simple
+cuspidal modules, we are now faced the daunting task of actually classifying
+them. Historically, this was first achieved by Olivier Mathieu in the early
+2000's in his paper \citetitle{mathieu} \cite{mathieu}. To do so, Mathieu
+introduced new tools which have since proved themselves remarkably useful
+throughout the field, known as\dots
+
+\section{Coherent Families}
+
+We begin our analysis with a simple question: how to do we go about
+constructing cuspidal modules? Specifically, given a cuspidal
+\(\mathfrak{g}\)-module, how can we use it to produce new cuspidal modules? To
+answer this question, we look back at the single example of a cuspidal module
+we have encountered so far: the \(\mathfrak{sl}_2(K)\)-module \(K[x, x^{-1}]\)
+of Laurent polynomials -- i.e. Example~\ref{ex:laurent-polynomial-mod}.
+
+Our first observation is that \(\mathfrak{sl}_2(K)\) acts on \(K[x, x^{-1}]\)
+via differential operators. In other words, the action map
+\(\mathcal{U}(\mathfrak{sl}_2(K)) \to \operatorname{End}(K[x, x^{-1}])\)
+factors through the inclusion of the algebra \(\operatorname{Diff}(K[x,
+x^{-1}]) = K\left[x, x^{-1}, \frac{\mathrm{d}}{\mathrm{d}x}\right]\) of
+differential operators in \(K[x, x^{-1}]\).
+\begin{center}
+  \begin{tikzcd}
+    \mathcal{U}(\mathfrak{sl}_2(K))   \rar &
+    \operatorname{Diff}(K[x, x^{-1}]) \rar &
+    \operatorname{End}(K[x, x^{-1}])
+  \end{tikzcd}
+\end{center}
+
+The space \(K[x, x^{-1}]\) can be regarded as a \(\operatorname{Diff}(K[x,
+x^{-1}])\)-module in the natural way, and we can produce new
+\(\operatorname{Diff}(K[x, x^{-1}])\)-modules by twisting \(K[x, x^{-1}]\) by
+automorphisms of \(\operatorname{Diff}(K[x, x^{-1}])\). For example, given
+\(\lambda \in K\) we may take the automorphism
+\begin{align*}
+  \varphi_\lambda : \operatorname{Diff}(K[x, x^{-1}]) &
+  \to \operatorname{Diff}(K[x, x^{-1}]) \\
+  x & \mapsto x \\
+  x^{-1} & \mapsto x^{-1} \\
+  \frac{\mathrm{d}}{\mathrm{d} x} & \mapsto \frac{\mathrm{d}}{\mathrm{d} x} +
+  \frac{\lambda}{2} x^{-1}
+\end{align*}
+and consider the twisted module \(\twisted{K[x, x^{-1}]}{\varphi_\lambda} =
+K[x, x^{-1}]\), where some operator \(P \in \operatorname{Diff}(K[x, x^{-1}])\)
+acts as \(\varphi_\lambda(P)\).
+
+By composing the action map \(\operatorname{Diff}(K[x, x^{-1}]) \to
+\operatorname{End}(\twisted{K[x, x^{-1}]}{\varphi_\lambda})\) with the
+homomorphism of algebras \(\mathcal{U}(\mathfrak{sl}_2(K)) \to
+\operatorname{Diff}(K[x, x^{-1}])\) we can give \(\twisted{K[x,
+x^{-1}]}{\varphi_\lambda}\) the structure of an \(\mathfrak{sl}_2(K)\)-module.
+Diagrammatically, we have
+\begin{center}
+  \begin{tikzcd}
+    \mathcal{U}(\mathfrak{sl}_2(K))   \rar                  &
+    \operatorname{Diff}(K[x, x^{-1}]) \rar{\varphi_\lambda} &
+    \operatorname{Diff}(K[x, x^{-1}]) \rar                  &
+    \operatorname{End}(K[x, x^{-1}])
+  \end{tikzcd},
+\end{center}
+where the maps \(\mathcal{U}(\mathfrak{sl}_2(K)) \to \operatorname{Diff}(K[x,
+x^{-1}])\) and \(\operatorname{Diff}(K[x, x^{1}]) \to \operatorname{End}(K[x,
+x^{-1}])\) are the ones from the previous diagram.
+
+Explicitly, we find that the action of \(\mathfrak{sl}_2(K)\) on
+\(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is given by
+\begin{align*}
+  p & \overset{e}{\mapsto}
+  \left(
+  x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \frac{1 + \lambda}{2} x
+  \right) p &
+  p & \overset{f}{\mapsto}
+  \left(
+  - \frac{\mathrm{d}}{\mathrm{d}x} + \frac{1 - \lambda}{2} x^{-1}
+  \right) p &
+  p & \overset{h}{\mapsto}
+  \left( 2 x \frac{\mathrm{d}}{\mathrm{d}x} + \lambda \right) p,
+\end{align*}
+so we can see \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}_{2 k +
+\frac{\lambda}{2}} = K x^k\) for all \(k \in \mathbb{Z}\) and \(\twisted{K[x,
+x^{-1}]}{\varphi_\lambda}_\mu = 0\) for all other \(\mu \in \mathfrak{h}^*\).
+
+Hence \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is a degree \(1\) bounded
+\(\mathfrak{sl}_2(K)\)-module with \(\operatorname{supp} \twisted{K[x,
+x^{-1}]}{\varphi_\lambda} = \frac{\lambda}{2} + 2 \mathbb{Z}\). One can also
+quickly check that if \(\lambda \notin 1 + 2 \mathbb{Z}\) then \(e\) and \(f\)
+act injectively in \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\), so that
+\(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is simple. In particular, if
+\(\lambda, \mu \notin 1 + 2 \mathbb{Z}\) with \(\lambda \notin \mu + 2
+\mathbb{Z}\) then \(\twisted{K[x, x^{-1}]}{\varphi_\lambda}\) and
+\(\twisted{K[x, x^{-1}]}{\varphi_\mu}\) are non-isomorphic simple cuspidal
+\(\mathfrak{sl}_2(K)\)-modules, since their supports differ. These cuspidal
+modules can be ``glued together'' in a \emph{monstrous concoction} by summing
+over \(\lambda \in K\), as in
+\[
+  \mathcal{M}
+  = \bigoplus_{\lambda + 2 \mathbb{Z} \in \mfrac{K}{2 \mathbb{Z}}}
+    \twisted{K[x, x^{-1}]}{\varphi_\lambda},
+\]
+
+To a distracted spectator, \(\mathcal{M}\) may look like just another,
+innocent, \(\mathfrak{sl}_2(K)\)-module. However, the attentive reader may have
+already noticed some of the its bizarre features, most noticeable of which is
+the fact that \(\mathcal{M}\) is very big. In fact, \(\mathcal{M}\) is as big a
+degree \(1\) bounded module gets: \(\operatorname{supp} \mathcal{M}
+= \operatorname{supp}_{\operatorname{ess}} \mathcal{M}\) is the entirety of
+\(\mathfrak{h}^*\). This may look very alien the reader familiarized with the
+finite-dimensional setting, where the configuration of weights is very rigid.
+For this reason, \(\mathcal{M}\) deserves to be called ``a monstrous
+concoction''.
+
+On a perhaps less derogatory note, \(\mathcal{M}\) also deserves to be called
+\emph{a family}. This is because \(\mathcal{M}\) consists of lots of smaller
+cuspidal modules which fit together inside of it in a \emph{coherent} fashion.
+Mathieu's ingenious breakthrough was the realization that \(\mathcal{M}\) is a
+particular example of a more general pattern, which he named \emph{coherent
+families}.
+
+\begin{definition}\index{coherent family}
+  A \emph{coherent family \(\mathcal{M}\) of degree \(d\)} is a weight
+  \(\mathfrak{g}\)-module \(\mathcal{M}\) such that
+  \begin{enumerate}
+    \item \(\dim \mathcal{M}_\lambda = d\) for \emph{all} \(\lambda \in
+      \mathfrak{h}^*\) -- i.e. \(\operatorname{supp}_{\operatorname{ess}}
+      \mathcal{M} = \mathfrak{h}^*\).
+    \item For any \(u \in \mathcal{U}(\mathfrak{g})\) in the centralizer
+      \(\mathcal{U}(\mathfrak{g})_0\) of \(\mathfrak{h}\) in
+      \(\mathcal{U}(\mathfrak{g})\), the map
+      \begin{align*}
+        \mathfrak{h}^* & \to K \\
+               \lambda & \mapsto
+               \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\lambda})
+      \end{align*}
+      is polynomial in \(\lambda\).
+  \end{enumerate}
+\end{definition}
+
+\begin{example}\label{ex:sl-laurent-family}
+  The module \(\mathcal{M} = \bigoplus_{\lambda + 2 \mathbb{Z} \in
+  \mfrac{K}{2 \mathbb{Z}}} \twisted{K[x, x^{-1}]}{\varphi_\lambda}\) is a
+  degree \(1\) coherent \(\mathfrak{sl}_2(K)\)-family.
+\end{example}
+
+\begin{example}
+  Given \(\lambda \in K\), \(\mathcal{M}(\lambda) = \bigoplus_{\mu \in K} K
+  x^\mu\) with
+  \begin{align*}
+    p & \overset{e}{\mapsto}
+        \left(x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \lambda x\right) p &
+    p & \overset{f}{\mapsto}
+        \left(-\frac{\mathrm{d}}{\mathrm{d}x} + \lambda x^{-1}\right) p &
+    p & \overset{h}{\mapsto} 2 x \frac{\mathrm{d}}{\mathrm{d}x} p,
+  \end{align*}
+  is a degree \(1\) coherent \(\mathfrak{sl}_2(K)\)-family -- where \(x^{\pm
+  1}, \sfrac{\mathrm{d}}{\mathrm{d}x} : \mathcal{M}(\lambda) \to
+  \mathcal{M}(\lambda)\) are given by \(x^{\pm 1} x^\mu = x^{\mu \pm 1}\) and
+  \(\sfrac{\mathrm{d}}{\mathrm{d}x} x^\mu = \mu x^{\mu - 1}\). It is easy to
+  check \(\mathcal{M}\) from Example~\ref{ex:sl-laurent-family} is isomorphic
+  to \(\mathcal{M}(\sfrac{1}{2})\) and \((\mathcal{M}(\sfrac{1}{2}))[0] \cong
+  K[x, x^{-1}]\).
+\end{example}
+
+\begin{note}
+  We would like to stress that coherent families have proven themselves useful
+  for problems other than the classification of cuspidal
+  \(\mathfrak{g}\)-modules. For instance, Nilsson's classification of rank 1
+  \(\mathfrak{h}\)-free \(\mathfrak{sp}_{2 n}(K)\)-modules is based on the
+  notion of coherent families and the so called \emph{weighting functor}.
+\end{note}
+
+Our hope is that given a simple cuspidal module \(M\), we can somehow fit \(M\)
+inside of a coherent \(\mathfrak{g}\)-family, such as in the case of \(K[x,
+x^{-1}]\) and \(\mathcal{M}\) from Example~\ref{ex:sl-laurent-family}. In
+addition, we hope that such coherent families are somehow \emph{uniquely
+determined} by \(M\). This leads us to the following definition.
+
+\begin{definition}\index{coherent family!coherent extension}
+  Given a bounded \(\mathfrak{g}\)-module \(M\) of degree \(d\), a
+  \emph{coherent extension \(\mathcal{M}\) of \(M\)} is a coherent family
+  \(\mathcal{M}\) of degree \(d\) that contains \(M\) as a subquotient.
+\end{definition}
+
+Our goal is now showing that every simple bounded module has a coherent
+extension. The idea then is to classify coherent families, and classify which
+submodules of a given coherent family are actually simple cuspidal modules. If
+every simple bounded \(\mathfrak{g}\)-module fits inside a coherent extension,
+this would lead to classification of all simple cuspidal
+\(\mathfrak{g}\)-modules, which we now know is the key for the solution of our
+classification problem. However, there are some complications to this scheme.
+
+Leaving aside the question of existence for a second, we should point out that
+coherent families turn out to be rather complicated on their own. In fact they
+are too complicated to classify in general. Ideally, we would like to find
+\emph{nice} coherent extensions -- ones we can actually classify. For instance,
+we may search for \emph{irreducible} coherent extensions, which are defined as
+follows.
+
+\begin{definition}\index{coherent family!irreducible coherent family}
+  A coherent family \(\mathcal{M}\) is called \emph{irreducible} if it contains
+  no proper coherent subfamilies -- i.e. \(\mathcal{M}\) is a simple object in
+  the full subcategory of \(\mathfrak{g}\text{-}\mathbf{Mod}\) consisting of
+  coherent families. Equivalently, we call \(\mathcal{M}\) irreducible if
+  \(\mathcal{M}_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module
+  for some \(\lambda \in \mathfrak{h}^*\).
+\end{definition}
+
+Another natural candidate for the role of ``nice extensions'' are the
+semisimple coherent families -- i.e. families which are semisimple as
+\(\mathfrak{g}\)-modules. These turn out to be very easy to produce. Namely,
+there is a construction, known as \emph{the semisimplification of a coherent
+family}, which takes a coherent extension of \(M\) to a semisimple coherent
+extension of \(M\).
+
+% Mathieu's proof of this is somewhat profane, I don't think it's worth
+% including it in here
+% TODO: Move this somewhere else? This holds in general for weight modules
+% whose suppert is contained in a single Q-coset
+\begin{lemma}\label{thm:component-coh-family-has-finite-length}
+  Given a coherent family \(\mathcal{M}\) and \(\lambda \in \mathfrak{h}^*\),
+  \(\mathcal{M}[\lambda]\) has finite length as a \(\mathfrak{g}\)-module.
+\end{lemma}
+
+\begin{proposition}\index{coherent family!semisimplification}
+  Let \(\mathcal{M}\) be a coherent family of degree \(d\). There exists a
+  unique semisimple coherent family \(\mathcal{M}^{\operatorname{ss}}\) of
+  degree \(d\) such that the composition series of
+  \(\mathcal{M}^{\operatorname{ss}}[\lambda]\) is the same as that of
+  \(\mathcal{M}[\lambda]\) for all \(\lambda \in \mathfrak{h}^*\), called
+  \emph{the semisimplification of \(\mathcal{M}\)}.
+
+  Namely, if \(\lambda \in \mathfrak{h}^*\) and \(0 = \mathcal{M}_{\lambda 0}
+  \subset \mathcal{M}_{\lambda 1} \subset \cdots \subset \mathcal{M}_{\lambda
+  r_\lambda} = \mathcal{M}[\lambda]\) is a composition series\footnote{Notice
+  that $\mathcal{M}[\lambda] = \mathcal{M}[\mu]$ for any $\mu \in \lambda + Q$.
+  Hence the sum $\bigoplus_{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q}}
+  \bigoplus_i \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}$ is
+  independent of the choice of representative for $\lambda + Q$ -- provided we
+  choose $\mathcal{M}_{\mu i} = \mathcal{M}_{\lambda i}$ for all $\mu \in
+  \lambda + Q$ and $i$.},
+  \[
+    \mathcal{M}^{\operatorname{ss}}
+    \cong \bigoplus_{\substack{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q} \\ i}}
+          \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
+  \]
+\end{proposition}
+
+\begin{proof}
+  The uniqueness of \(\mathcal{M}^{\operatorname{ss}}\) should be clear:
+  since \(\mathcal{M}^{\operatorname{ss}}\) is semisimple, so is
+  \(\mathcal{M}^{\operatorname{ss}}[\lambda]\). Hence by the Jordan-Hölder
+  Theorem
+  \[
+    \mathcal{M}^{\operatorname{ss}}[\lambda]
+    \cong
+    \bigoplus_i \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
+  \]
+
+  As for the existence of the semisimplification, it suffices to show
+  \[
+    \mathcal{M}^{\operatorname{ss}}
+    = \bigoplus_{\substack{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q} \\ i}}
+    \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
+  \]
+  is indeed a semisimple coherent family of degree \(d\).
+
+  We know from Examples~\ref{ex:submod-is-weight-mod} and
+  \ref{ex:quotient-is-weight-mod} that each quotient
+  \(\mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}\) is a weight
+  module. Hence \(\mathcal{M}^{\operatorname{ss}}\) is a weight module.
+  Furthermore, given \(\mu \in \mathfrak{h}^*\)
+  \[
+    \mathcal{M}_\mu^{\operatorname{ss}}
+    = \bigoplus_{\substack{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q} \\ i}}
+      \left(
+      \mfrac{\mathcal{M}_{\lambda i + 1}}{\mathcal{M}_{\lambda i}}
+      \right)_\mu
+    = \bigoplus_i
+      \left(
+      \mfrac{\mathcal{M}_{\mu i + 1}}{\mathcal{M}_{\mu i}}
+      \right)_\mu
+    \cong \bigoplus_i
+      \mfrac{(\mathcal{M}_{\mu i + 1})_\mu}
+            {(\mathcal{M}_{\mu i})_\mu}
+  \]
+
+  In particular,
+  \[
+    \dim \mathcal{M}_\mu^{\operatorname{ss}}
+    = \sum_i
+      \dim (\mathcal{M}_{\mu i + 1})_\mu - \dim (\mathcal{M}_{\mu i})_\mu
+    = \dim \mathcal{M}[\mu]_\mu
+    = \dim \mathcal{M}_\mu
+    = d
+  \]
+
+  Likewise, given \(u \in \mathcal{U}(\mathfrak{g})_0\) the value
+  \[
+    \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\mu^{\operatorname{ss}}})
+    = \sum_i
+      \operatorname{Tr}(u\!\restriction_{(\mathcal{M}_{\mu i + 1})_\mu})
+    - \operatorname{Tr}(u\!\restriction_{(\mathcal{M}_{\mu i})_\mu})
+    = \operatorname{Tr}(u\!\restriction_{\mathcal{M}[\mu]_\mu})
+    = \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\mu})
+  \]
+  is polynomial in \(\mu \in \mathfrak{h}^*\).
+\end{proof}
+
+\begin{note}
+  Although we have provided an explicit construction of
+  \(\mathcal{M}^{\operatorname{ss}}\) in terms of \(\mathcal{M}\), we should
+  point out this construction is not functorial. First, given a
+  \(\mathfrak{g}\)-homomorphism \(f : \mathcal{M} \to \mathcal{N}\) between
+  coherent families, it is unclear what \(f^{\operatorname{ss}} :
+  \mathcal{M}^{\operatorname{ss}} \to \mathcal{N}^{\operatorname{ss}}\) is
+  supposed to be. Secondly, and this is more relevant, our construction depends
+  on the choice of composition series \(0 = \mathcal{M}_{\lambda 0} \subset
+  \cdots \subset \mathcal{M}_{\lambda r_\lambda} = \mathcal{M}[\lambda]\).
+  While different choices of composition series yield isomorphic results, there
+  is no canonical isomorphism. In addition, there is no canonical choice of
+  composition series.
+\end{note}
+
+The proof of Lemma~\ref{thm:component-coh-family-has-finite-length} is
+extremely technical and will not be included in here. It suffices to note that,
+as in Proposition~\ref{thm:ess-supp-is-zariski-dense}, the general case follows
+from the case where \(\mathfrak{g}\) is simple, which may be found in
+\cite{mathieu} -- see Lemma 3.3. As promised, if \(\mathcal{M}\) is a coherent
+extension of \(M\) then so is \(\mathcal{M}^{\operatorname{ss}}\).
+
+\begin{proposition}
+  Let \(M\) be a simple bounded \(\mathfrak{g}\)-module and \(\mathcal{M}\)
+  be a coherent extension of \(M\). Then \(\mathcal{M}^{\operatorname{ss}}\) is
+  a coherent extension of \(M\), and \(M\) is in fact a submodule of
+  \(\mathcal{M}^{\operatorname{ss}}\).
+\end{proposition}
+
+\begin{proof}
+  Since \(M\) is simple, its support is contained in a single \(Q\)-coset.
+  This implies that \(M\) is a subquotient of \(\mathcal{M}[\lambda]\) for any
+  \(\lambda \in \operatorname{supp} M\). If we fix some composition series \(0
+  = \mathcal{M}_0 \subset \mathcal{M}_1 \subset \cdots \subset \mathcal{M}_r =
+  \mathcal{M}[\lambda]\) of \(\mathcal{M}[\lambda]\) with \(M \cong
+  \mfrac{\mathcal{M}_{i + 1}}{\mathcal{M}_i}\), there is a natural inclusion
+  \[
+    M
+    \isoto \mfrac{\mathcal{M}_{i + 1}}{\mathcal{M}_i}
+    \to \bigoplus_j \mfrac{\mathcal{M}_{j + 1}}{\mathcal{M}_j}
+    \cong \mathcal{M}^{\operatorname{ss}}[\lambda]
+  \]
+\end{proof}
+
+Given the uniqueness of the semisimplification, the semisimplification of any
+semisimple coherent extension \(\mathcal{M}\) is \(\mathcal{M}\)
+itself and therefore\dots
+
+\begin{corollary}\label{thm:bounded-is-submod-of-extension}
+  Let \(M\) be a simple bounded \(\mathfrak{g}\)-module and \(\mathcal{M}\)
+  be a semisimple coherent extension of \(M\). Then \(M\) is
+  contained in \(\mathcal{M}\).
+\end{corollary}
+
+These last results provide a partial answer to the question of existence of
+well behaved coherent extensions. As for the uniqueness \(\mathcal{M}\) in
+Corollary~\ref{thm:bounded-is-submod-of-extension}, it suffices to show that
+the multiplicities of the simple weight \(\mathfrak{g}\)-modules in
+\(\mathcal{M}\) are uniquely determined by \(M\). These multiplicities may be
+computed via the following lemma.
+
+\begin{lemma}\label{thm:centralizer-multiplicity}
+  Let \(M\) be a semisimple weight \(\mathfrak{g}\)-module. Then \(M_\lambda\)
+  is a semisimple \(\mathcal{U}(\mathfrak{g})_0\)-module for any \(\lambda \in
+  \operatorname{supp} M\). Moreover, if \(L\) is a simple weight
+  \(\mathfrak{g}\)-module such that \(\lambda \in \operatorname{supp} L\) then
+  \(L_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module and the
+  multiplicity \(L\) in \(M\) coincides with the multiplicity of \(L_\lambda\)
+  in \(M_\lambda\) as a \(\mathcal{U}(\mathfrak{g})_0\)-module.
+\end{lemma}
+
+\begin{proof}
+  We begin by showing that \(L_\lambda\) is simple. Let \(N \subset L_\lambda\)
+  be a nontrivial \(\mathcal{U}(\mathfrak{g})_0\)-submodule. We want to
+  establish that \(N = L_\lambda\).
+
+  If \(\mathcal{U}(\mathfrak{g})_\alpha\) denotes the root space of \(\alpha\)
+  in \(\mathcal{U}(\mathfrak{g})\) under the adjoint action of \(\mathfrak{g}\)
+  as in Example~\ref{ex:adjoint-action-in-universal-enveloping-is-weight},
+  \(\alpha \in Q\), a simple calculation shows
+  \(\mathcal{U}(\mathfrak{g})_\alpha \cdot N \subset L_{\lambda + \alpha}\).
+  Since \(L\) is simple and \(N\) is nonzero, it follows from
+  Example~\ref{ex:adjoint-action-in-universal-enveloping-is-weight} that
+  \[
+    L
+    = \mathcal{U}(\mathfrak{g}) \cdot N
+    = \bigoplus_{\alpha \in Q} \mathcal{U}(\mathfrak{g})_\alpha \cdot N
+  \]
+  and thus \(L_{\lambda + \alpha} = \mathcal{U}(\mathfrak{g})_\alpha \cdot N\).
+  In particular, \(L_\lambda = \mathcal{U}(\mathfrak{g})_0 \cdot N \subset N\)
+  and \(N = L_\lambda\).
+
+  Now given a semisimple weight \(\mathfrak{g}\)-module \(M = \bigoplus_i M_i\)
+  with \(M_i\) simple, it is clear \(M_\lambda = \bigoplus_i (M_i)_\lambda\).
+  Each \((M_i)_\lambda\) is either \(0\) or a simple
+  \(\mathcal{U}(\mathfrak{g})_0\)-module, so that \(M_\lambda\) is a semisimple
+  \(\mathcal{U}(\mathfrak{g})_0\)-module. In addition, to see that the
+  multiplicity of \(L\) in \(M\) coincides with the multiplicity of
+  \(L_\lambda\) in \(M_\lambda\) it suffices to show that if \((M_i)_\lambda
+  \cong (M_j)_\lambda\) are both nonzero then \(M_i \cong M_j\).
+
+  If \(I(M_i) = \mathcal{U}(\mathfrak{g}) \otimes_{\mathcal{U}(\mathfrak{g})_0}
+  (M_i)_\lambda\), the inclusion of \(\mathcal{U}(\mathfrak{g})_0\)-modules
+  \((M_i)_\lambda \to M_i\) induces a \(\mathfrak{g}\)-homomorphism
+  \begin{align*}
+            I(M_i) & \to     M_i       \\
+    u \otimes m & \mapsto u \cdot m
+  \end{align*}
+
+  Since \(M_i\) is simple and \(\lambda \in \operatorname{supp} M_i\), \(M_i =
+  \mathcal{U}(\mathfrak{g}) \cdot (M_i)_\lambda\). The homomorphism \(I(M_i)
+  \to M_i\) is thus surjective. Similarly, if \(I(M_j) =
+  \mathcal{U}(\mathfrak{g}) \otimes_{\mathcal{U}(\mathfrak{g})_0}
+  (M_j)_\lambda\) then there is a natural surjective
+  \(\mathfrak{g}\)-homomorphism \(I(M_j) \to M_j\). Now suppose there is an
+  isomorphism of \(\mathcal{U}(\mathfrak{g})_0\)-modules \(f: (M_i)_\lambda
+  \isoto (M_j)_\lambda\). Such an isomorphism induces an isomorphism of
+  \(\mathfrak{g}\)-modules
+  \begin{align*}
+    \tilde f : I(M_i) & \isoto  I(M_j)            \\
+       u \otimes m & \mapsto u \otimes f(m)
+  \end{align*}
+
+  By composing \(\tilde f\) with the projection \(I(M_j) \to M_j\) we get a
+  surjective homomorphism \(I(M_i) \to M_j\). We claim \(\ker (I(M_i) \to M_i)
+  = \ker (I(M_i) \to M_j)\).  To see this, notice that \(\ker(I(M_i) \to M_i)\)
+  coincides with the largest submodule \(Z(M_i) \subset I(M_i)\) contained in
+  \(\bigoplus_{\alpha \ne 0} \mathcal{U}(\mathfrak{g})_\alpha
+  \otimes_{\mathcal{U}(\mathfrak{g})_0} (M_i)_\lambda\). Indeed, a simple
+  computation shows \(\ker (I(M_i) \to M_i) \cap (\mathcal{U}(\mathfrak{g})_0
+  \otimes_{\mathcal{U}(\mathfrak{g})_0} (M_i)_\lambda) = 0\), which implies
+  \(\ker(I(M_i) \to M_i) \subset Z(M_i)\). Since \(M_i\) is simple, \(\ker
+  (I(M_i) \to M_i)\) is maximal and thus \(\ker(I(M_i) \to M_i) = Z(M_i)\). By
+  the same token, \(\ker (I(M_j) \to M_j)\) is the largest submodule of
+  \(I(M_j)\) contained in \(\bigoplus_{\alpha \ne 0}
+  \mathcal{U}(\mathfrak{g})_\alpha \otimes_{\mathcal{U}(\mathfrak{g})_0}
+  (M_j)_\lambda\) and therefore \(\ker(I(M_i) \to M_i) =
+  \tilde{f}^{-1}(\ker(I(M_j) \to M_j)) = \ker(I(M_i) \to M_j)\).
+
+  Hence there is an isomorphism \(\mfrac{I(M_i)}{\ker(I(M_i) \to M_i)} \isoto
+  M_j\) satisfying
+  \begin{center}
+    \begin{tikzcd}
+      I(M_i) \rar{\tilde f} \dar & I(M_j) \dar \\
+      \mfrac{I(M_i)}{\ker(I(M_i) \to M_i)} \rar{\sim} & M_j
+    \end{tikzcd}
+  \end{center}
+ and finally \(M_i \cong \mfrac{I(M_i)}{\ker(I(M_i) \to M_i)} \cong M_j\).
+\end{proof}
+
+A complementary question now is: which submodules of a \emph{nice} coherent
+family are cuspidal representations?
+
+\begin{proposition}[Mathieu]
+  Let \(\mathcal{M}\) be an irreducible coherent family of degree \(d\) and
+  \(\lambda \in \mathfrak{h}^*\). The following conditions are equivalent.
+  \begin{enumerate}
+    \item \(\mathcal{M}[\lambda]\) is simple.
+    \item \(F_\alpha\!\restriction_{\mathcal{M}[\lambda]}\) is injective for
+      all \(\alpha \in \Delta\).
+    \item \(\mathcal{M}[\lambda]\) is cuspidal.
+  \end{enumerate}
+\end{proposition}
+
+\begin{proof}
+  The fact that \strong{(i)} and \strong{(iii)} are equivalent follows directly
+  from Corollary~\ref{thm:cuspidal-mod-equivs}. Likewise, it is clear from the
+  corollary that \strong{(iii)} implies \strong{(ii)}. All it is left is to
+  show \strong{(ii)} implies \strong{(iii)}. This isn't already clear from
+  Corollary~\ref{thm:cuspidal-mod-equivs} because, at first glance,
+  $\mathcal{M}[\lambda]$ may not be simple for some $\lambda$ satisfying
+  \strong{(ii)}. We will show this is never the case.
+
+  Suppose \(F_\alpha\) acts injectively on the submodule
+  \(\mathcal{M}[\lambda]\), for all \(\alpha \in \Delta\). Since
+  \(\mathcal{M}[\lambda]\) has finite length, \(\mathcal{M}[\lambda]\) contains
+  an infinite-dimensional simple \(\mathfrak{g}\)-submodule \(M\). Moreover,
+  again by Corollary~\ref{thm:cuspidal-mod-equivs} we conclude \(M\) is a
+  cuspidal module, and its degree is bounded by \(d\). We want to show
+  \(\mathcal{M}[\lambda] = M\).
+
+  We claim the set \(U = \{\mu \in \mathfrak{h}^* : \mathcal{M}_\mu \ \text{is
+  a simple $\mathcal{U}(\mathfrak{g})_0$-module}\}\) is Zariski-open. If we
+  suppose this is the case for a moment or two, it follows from the fact that
+  \(M\) is simple and \(\operatorname{supp}_{\operatorname{ess}} M\) is
+  Zariski-dense that \(U \cap \operatorname{supp}_{\operatorname{ess}} M\) is
+  non-empty. In other words, there is some \(\mu \in \mathfrak{h}^*\) such that
+  \(\mathcal{M}_\mu\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module and
+  \(\dim M_\mu = \deg M\).
+
+  In particular, \(M_\mu \ne 0\), so \(M_\mu = \mathcal{M}_\mu\). Now given any
+  simple \(\mathfrak{g}\)-module \(L\), it follows from
+  Lemma~\ref{thm:centralizer-multiplicity} that the multiplicity of \(L\)
+  in \(\mathcal{M}[\lambda]\) is the same as the multiplicity \(L_\mu\) in
+  \(\mathcal{M}_\mu\) as a \(\mathcal{U}(\mathfrak{g})_0\)-module -- which is,
+  of course, \(1\) if \(L \cong M\) and \(0\) otherwise. Hence
+  \(\mathcal{M}[\lambda] = M\) and \(\mathcal{M}[\lambda]\) is cuspidal.
+\end{proof}
+
+To finish the proof, we now show\dots
+
+\begin{lemma}
+  Let \(\mathcal{M}\) be a coherent family. The set \(U = \{\lambda \in
+  \mathfrak{h}^* : \mathcal{M}_\lambda \ \text{is a simple
+  $\mathcal{U}(\mathfrak{g})_0$-module}\}\) is Zariski-open.
+\end{lemma}
+
+\begin{proof}
+  For each \(\lambda \in \mathfrak{h}^*\) we introduce the bilinear form
+  \begin{align*}
+    B_\lambda : \mathcal{U}(\mathfrak{g})_0 \times \mathcal{U}(\mathfrak{g})_0
+    & \to K \\
+    (u, v)
+    & \mapsto \operatorname{Tr}(u v \!\restriction_{\mathcal{M}_\lambda})
+  \end{align*}
+  and consider its rank -- i.e. the dimension of the image of the induced
+  operator
+  \begin{align*}
+    \mathcal{U}(\mathfrak{g})_0 & \to     \mathcal{U}(\mathfrak{g})_0^* \\
+                              u & \mapsto B_\lambda(u, \cdot)
+  \end{align*}
+
+  Our first observation is that \(\operatorname{rank} B_\lambda \le d^2\). This
+  follows from the commutativity of
+  \begin{center}
+    \begin{tikzcd}
+      \mathcal{U}(\mathfrak{g})_0               \rar \dar  &
+      \mathcal{U}(\mathfrak{g})_0^*                        \\
+      \operatorname{End}(\mathcal{M}_\lambda)   \rar{\sim} &
+      \operatorname{End}(\mathcal{M}_\lambda)^* \uar
+    \end{tikzcd},
+  \end{center}
+  where the map \(\mathcal{U}(\mathfrak{g})_0 \to
+  \operatorname{End}(\mathcal{M}_\lambda)\) is given by the action of
+  \(\mathcal{U}(\mathfrak{g})_0\), the map
+  \(\operatorname{End}(\mathcal{M}_\lambda)^* \to
+  \mathcal{U}(\mathfrak{g})_0^*\) is its dual, and the isomorphism
+  \(\operatorname{End}(\mathcal{M}_\lambda) \isoto
+  \operatorname{End}(\mathcal{M}_\lambda)^*\) is induced by the trace form
+  \begin{align*}
+    \operatorname{End}(\mathcal{M}_\lambda) \times
+    \operatorname{End}(\mathcal{M}_\lambda) & \to K \\
+    (T, S) & \mapsto \operatorname{Tr}(T S)
+  \end{align*}
+
+  Indeed, \(\operatorname{rank} B_\lambda \le
+  \operatorname{rank}(\mathcal{U}(\mathfrak{g})_0 \to
+  \operatorname{End}(\mathcal{M}_\lambda)) \le \dim
+  \operatorname{End}(\mathcal{M}_\lambda) = d^2\). Furthermore, if
+  \(\operatorname{rank} B_\lambda = d^2\) then we must have
+  \(\operatorname{rank}(\mathcal{U}(\mathfrak{g})_0 \to
+  \operatorname{End}(\mathcal{M}_\lambda)) = d^2\) -- i.e. the map
+  \(\mathcal{U}(\mathfrak{g})_0 \to \operatorname{End}(\mathcal{M}_\lambda)\)
+  is surjective. In particular, if \(\operatorname{rank} B_\lambda = d^2\) then
+  \(\mathcal{M}_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module,
+  for if \(M \subset \mathcal{M}_\lambda\) is invariant under the action of
+  \(\mathcal{U}(\mathfrak{g})_0\) then \(M\) is invariant under any
+  \(K\)-linear operator \(\mathcal{M}_\lambda \to \mathcal{M}_\lambda\), so
+  that \(M = 0\) or \(M = \mathcal{M}_\lambda\).
+
+  On the other hand, if \(\mathcal{M}_\lambda\) is simple then by Burnside's
+  Theorem on matrix algebras the map \(\mathcal{U}(\mathfrak{g})_0 \to
+  \operatorname{End}(\mathcal{M}_\lambda)\) is surjective. Hence the
+  commutativity of the previously drawn diagram, as well as the fact that
+  \(\operatorname{rank}(\mathcal{U}(\mathfrak{g})_0 \to
+  \operatorname{End}(\mathcal{M}_\lambda)) =
+  \operatorname{rank}(\operatorname{End}(\mathcal{M}_\lambda)^* \to
+  \mathcal{U}(\mathfrak{g})_0^*)\), imply that \(\operatorname{rank} B_\lambda
+  = d^2\). This goes to show that \(U\) is precisely the set of all \(\lambda\)
+  such that \(B_\lambda\) has maximal rank \(d^2\). We now show that \(U\) is
+  Zariski-open. First, notice that
+  \[
+    U =
+    \bigcup_{\substack{V \subset \mathcal{U}(\mathfrak{g})_0 \\ \dim V = d}}
+    U_V,
+  \]
+  where \(U_V = \{\lambda \in \mathfrak{h}^* : \operatorname{rank}
+  B_\lambda\!\restriction_V = d^2 \}\). Here \(V\) ranges over all
+  \(d\)-dimensional subspaces of \(\mathcal{U}(\mathfrak{g})_0\) -- \(V\) is
+  not necessarily a \(\mathcal{U}(\mathfrak{g})_0\)-submodule.
+
+  Indeed, if \(\operatorname{rank} B_\lambda = d^2\) it follows from the
+  subjectivity of the map \(\mathcal{U}(\mathfrak{g})_0 \to
+  \operatorname{End}(\mathcal{M}_\lambda)\) that there is some \(V \subset
+  \mathcal{U}(\mathfrak{g})_0\) with \(\dim V = d\) such that the restriction
+  \(V \to \operatorname{End}(\mathcal{M}_\lambda)\) is surjective. The
+  commutativity of
+  \begin{center}
+    \begin{tikzcd}
+      V                                         \rar \dar  & V^* \\
+      \operatorname{End}(\mathcal{M}_\lambda)   \rar{\sim} &
+      \operatorname{End}(\mathcal{M}_\lambda)^* \uar
+    \end{tikzcd}
+  \end{center}
+  then implies \(\operatorname{rank} B_\lambda\!\restriction_V = d^2\). In
+  other words, \(U \subset \bigcup_V U_V\).
+
+  Likewise, if \(\operatorname{rank} B_\lambda\!\restriction_V = d^2\) for some
+  \(V\), then the commutativity of
+  \begin{center}
+    \begin{tikzcd}
+      V                             \rar \dar & V^* \\
+      \mathcal{U}(\mathfrak{g})_0   \rar      &
+      \mathcal{U}(\mathfrak{g})_0^* \uar
+    \end{tikzcd}
+  \end{center}
+  implies \(\operatorname{rank} B_\lambda \ge d^2\), which goes to show
+  \(\bigcup_V U_V \subset U\).
+
+  Given \(\lambda \in U_V\), the surjectivity of \(V \to
+  \operatorname{End}(\mathcal{M}_\lambda)\) and the fact that \(\dim V <
+  \infty\) imply \(V \to V^*\) is invertible. Since \(\mathcal{M}\) is a
+  coherent family, \(B_\lambda\) depends polynomially in \(\lambda\). Hence so
+  does the induced maps \(V \to V^*\). In particular, there is some Zariski
+  neighborhood \(U'\) of \(\lambda\) such that the map \(V \to V^*\) induced by
+  \(B_\mu\!\restriction_V\) is invertible for all \(\mu \in U'\).
+
+  But the surjectivity of the map induced by \(B_\mu\!\restriction_V\) implies
+  \(\operatorname{rank} B_\mu = d^2\), so \(\mu \in U_V\) and therefore \(U'
+  \subset U_V\). This implies \(U_V\) is open for all \(V\). Finally, \(U\) is
+  the union of Zariski-open subsets and is therefore open. We are done.
+\end{proof}
+
+The major remaining question for us to tackle is that of the existence of
+coherent extensions, which will be the focus of our next section.
+
+\section{Localizations \& the Existence of Coherent Extensions}
+
+Let \(M\) be a simple bounded \(\mathfrak{g}\)-module of degree \(d\). Our
+goal is to prove that \(M\) has a (unique) irreducible semisimple coherent
+extension \(\mathcal{M}\). Since \(M\) is simple, we know \(M \subset
+\mathcal{M}[\lambda]\) for any \(\lambda \in \operatorname{supp} M\). Our first
+task is constructing \(\mathcal{M}[\lambda]\). The issue here is that
+\(\operatorname{supp}_{\operatorname{ess}} M\) may not be all of \(\lambda + Q
+= \operatorname{supp}_{\operatorname{ess}} \mathcal{M}[\lambda]\), so we may
+find \(M \subsetneq \mathcal{M}[\lambda]\). In fact, we may find
+\(\operatorname{supp} M \subsetneq \lambda + Q\).
+
+This wasn't an issue an Example~\ref{ex:laurent-polynomial-mod} because we
+verified that the action of \(f \in \mathfrak{sl}_2(K)\) on \(K[x, x^{-1}]\) is
+injective. Since all weight spaces of \(K[x, x^{-1}]\) are \(1\)-dimensional,
+this implies the action of \(f\) is actually bijective, so we can obtain a
+nonzero vector in \(K[x, x^{-1}]_{2 k} = K x^k\) for any \(k \in \mathbb{Z}\)
+by translating between weight spaced using \(f\) and \(f^{-1}\) -- here
+\(f^{-1}\) denotes the \(K\)-linear operator \((-
+\sfrac{\mathrm{d}}{\mathrm{d}x} + \sfrac{x^{-1}}{2})^{-1}\), which is the
+inverse of the action of \(f\) on \(K[x, x^{-1}]\).
+\begin{center}
+  \begin{tikzcd}
+    \cdots     \rar[bend left=60]{f^{-1}}
+    & K x^{-2} \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
+    & K x^{-1} \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
+    & K        \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
+    & K x      \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
+    & K x^2    \rar[bend left=60]{f^{-1}} \lar[bend left=60]{f}
+    & \cdots   \lar[bend left=60]{f}
+  \end{tikzcd}
+\end{center}
+
+In the general case, the action of some \(F_\alpha \in \mathfrak{g}\) with
+\(\alpha \in \Delta\) in \(M\) may not be injective. In fact, we have seen that
+the action of \(F_\alpha\) is injective for all \(\alpha \in \Delta^+\) if, and
+only if \(M\) is cuspidal. Nevertheless, we could intuitively \emph{make it
+injective} by formally inverting the elements \(F_\alpha \in
+\mathcal{U}(\mathfrak{g})\). This would allow us to obtain nonzero vectors in
+\(M_\mu\) for all \(\mu \in \lambda + Q\) by successively applying elements of
+\(\{F_\alpha^{\pm 1}\}_{\alpha \in \Delta}\) to a nonzero weight vector \(m \in
+M_\lambda\). Moreover, if the actions of the \(F_\alpha\) were to be
+invertible, we would find that all \(M_\mu\) are \(d\)-dimensional for \(\mu
+\in \lambda + Q\).
+
+In a commutative domain, this can be achieved by tensoring our module by the
+field of fractions. However, \(\mathcal{U}(\mathfrak{g})\) is hardly ever
+commutative -- \(\mathcal{U}(\mathfrak{g})\) is commutative if, and only if
+\(\mathfrak{g}\) is Abelian -- and the situation is more delicate in the
+non-commutative case. For starters, a non-commutative \(K\)-algebra \(A\) may
+not even have a ``field of fractions'' -- i.e. an over-ring where all elements
+of \(A\) have inverses. Nevertheless, it is possible to formally invert
+elements of certain subsets of \(A\) via a process known as
+\emph{localization}, which we now describe.
+
+\begin{definition}\index{localization!multiplicative subsets}\index{localization!Ore's condition}
+  Let \(A\) be a \(K\)-algebra. A subset \(S \subset A\) is called
+  \emph{multiplicative} if \(s \cdot t \in S\) for all \(s, t \in S\) and \(0
+  \notin S\). A multiplicative subset \(S\) is said to satisfy \emph{Ore's
+  localization condition} if for each \(a \in A\) and \(s \in S\) there exists
+  \(b, c \in A\) and \(t, t' \in S\) such that \(s a = b t\) and \(a s = t'
+  c\).
+\end{definition}
+
+\begin{theorem}[Ore-Asano]\index{localization!Ore-Asano Theorem}
+  Let \(S \subset A\) be a multiplicative subset satisfying Ore's localization
+  condition. Then there exists a (unique) \(K\)-algebra \(S^{-1} A\), with a
+  canonical algebra homomorphism \(A \to S^{-1} A\), enjoying the universal
+  property that each algebra homomorphism \(f : A \to B\) such that \(f(s)\) is
+  invertible for all \(s \in S\) can be uniquely extended to an algebra
+  homomorphism \(S^{-1} A \to B\). \(S^{-1} A\) is called \emph{the
+  localization of \(A\) by \(S\)}, and the map \(A \to S^{-1} A\) is called
+  \emph{the localization map}.
+  \begin{center}
+    \begin{tikzcd}
+      A        \dar \rar{f}        & B \\
+      S^{-1} A \urar[swap, dotted] &
+    \end{tikzcd}
+  \end{center}
+\end{theorem}
+
+If we identify an element with its image under the localization map, it follows
+directly from Ore's construction that every element of \(S^{-1} A\) has the
+form \(s^{-1} a\) for some \(s \in S\) and \(a \in A\). Likewise, any element
+of \(S^{-1} A\) can also be written as \(b t^{-1}\) for some \(t \in S\), \(b
+\in A\).
+
+Ore's localization condition may seem a bit arbitrary at first, but a more
+thorough investigation reveals the intuition behind it. The issue in question
+here is that in the non-commutative case we can no longer take the existence of
+common denominators for granted. However, the existence of common denominators
+is fundamental to the proof of the fact the field of fractions is a ring -- it
+is used, for example, to define the sum of two elements in the field of
+fractions. We thus need to impose their existence for us to have any hope of
+defining consistent arithmetics in the localization of an algebra, and Ore's
+condition is actually equivalent to the existence of common denominators --
+see the discussion in the introduction of \cite[ch.~6]{goodearl-warfield} for
+further details.
+
+We should also point out that there are numerous other conditions -- which may
+be easier to check than Ore's -- known to imply Ore's condition. For
+instance\dots
+
+\begin{lemma}
+  Let \(S \subset A\) be a multiplicative subset generated by finitely many
+  locally \(\operatorname{ad}\)-nilpotent elements -- i.e. elements \(s \in S\)
+  such that for each \(a \in A\) there exists \(r > 0\) such that
+  \(\operatorname{ad}(s)^r a = [s, [s, \cdots [s, a]]\cdots] = 0\). Then \(S\)
+  satisfies Ore's localization condition.
+\end{lemma}
+
+In our case, we are more interested in formally inverting the action of
+\(F_\alpha\) on \(M\) than in inverting \(F_\alpha\) itself. To that end, we
+introduce one further construction, known as \emph{the localization of a
+module}.
+
+\begin{definition}\index{localization!localization of modules}
+  Let \(S \subset A\) be a multiplicative subset satisfying Ore's localization
+  condition and \(M\) be an \(A\)-module. The \(S^{-1} A\)-module \(S^{-1} M =
+  S^{-1} A \otimes_A M\) is called \emph{the localization of \(M\) by \(S\)},
+  and the homomorphism of \(A\)-modules
+  \begin{align*}
+    M & \to     S^{-1} M    \\
+    m & \mapsto 1 \otimes m
+  \end{align*}
+  is called \emph{the localization map of \(M\)}.
+\end{definition}
+
+Notice that the \(S^{-1} A\)-module \(S^{-1} M\) has the natural structure of
+an \(A\)-module, where the action of \(A\) is given by the localization map \(A
+\to S^{-1} A\).
+
+It is interesting to observe that, unlike in the case of the field of fractions
+of a commutative domain, in general the localization map \(A \to S^{-1} A\) --
+i.e. the map \(a \mapsto \frac{a}{1}\) -- may not be injective. For instance,
+if \(S\) contains a divisor of zero \(s\), its image under the localization map
+is invertible and therefore cannot be a divisor of zero in \(S^{-1} A\). In
+particular, if \(a \in A\) is nonzero and such that \(s a = 0\) or \(a s = 0\)
+then its image under the localization map has to be \(0\). However, the
+existence of divisors of zero in \(S\) turns out to be the only obstruction to
+the injectivity of the localization map, as shown in\dots
+
+\begin{lemma}
+  Let \(S \subset A\) be a multiplicative subset satisfying Ore's localization
+  condition and \(M\) be an \(A\)-module. If \(S\) acts injectively on \(M\)
+  then the localization map \(M \to S^{-1} M\) is injective. In particular, if
+  \(S\) has no zero divisors then \(A\) is a subalgebra of \(S^{-1} A\).
+\end{lemma}
+
+Again, in our case we are interested in inverting the actions of the
+\(F_\alpha\) on \(M\). However, for us to be able to translate between all
+weight spaces associated with elements of \(\lambda + Q\), \(\lambda \in
+\operatorname{supp} M\), we only need to invert the \(F_\alpha\)'s for
+\(\alpha\) in some subset of \(\Delta\) which spans all of \(Q = \mathbb{Z}
+\Delta\). In other words, it suffices to invert \(F_\beta\) for all \(\beta\)
+in some basis \(\Sigma\) for \(\Delta\). We can choose such a basis to be
+well-behaved. For example, we can show\dots
+
+\begin{lemma}\label{thm:nice-basis-for-inversion}
+  Let \(M\) be a simple infinite-dimensional bounded \(\mathfrak{g}\)-module.
+  There is a basis \(\Sigma = \{\beta_1, \ldots, \beta_r\}\) for \(\Delta\)
+  such that the elements \(F_{\beta_i}\) all act injectively on \(M\) and
+  satisfy \([F_{\beta_i}, F_{\beta_j}] = 0\).
+\end{lemma}
+
+\begin{note}
+  The basis \(\Sigma\) in Lemma~\ref{thm:nice-basis-for-inversion} may very
+  well depend on the representation \(M\)! This is another obstruction to the
+  functoriality of our constructions.
+\end{note}
+
+The proof of the previous Lemma is quite technical and was deemed too tedious
+to be included in here. See Lemma 4.4 of \cite{mathieu} for a full proof. Since
+\(F_\alpha\) is locally \(\operatorname{ad}\)-nilpotent for all \(\alpha \in
+\Delta\), we can see\dots
+
+\begin{corollary}
+  Let \(\Sigma\) be as in Lemma~\ref{thm:nice-basis-for-inversion} and
+  \((F_\beta)_{\beta \in \Sigma} \subset \mathcal{U}(\mathfrak{g})\) be the
+  multiplicative subset generated by the \(F_\beta\)'s. The \(K\)-algebra
+  \(\Sigma^{-1} \mathcal{U}(\mathfrak{g}) = (F_\beta)_{\beta \in \Sigma}^{-1}
+  \mathcal{U}(\mathfrak{g})\) is well defined. Moreover, if we denote by
+  \(\Sigma^{-1} M\) the localization of \(M\) by \((F_\beta)_{\beta \in
+  \Sigma}\), the localization map \(M \to \Sigma^{-1} M\) is injective.
+\end{corollary}
+
+From now on let \(\Sigma\) be some fixed basis for \(\Delta\) satisfying the
+hypothesis of Lemma~\ref{thm:nice-basis-for-inversion}. We now show that
+\(\Sigma^{-1} M\) is a weight \(\mathfrak{g}\)-module whose support is an
+entire \(Q\)-coset.
+
+\begin{proposition}\label{thm:irr-bounded-is-contained-in-nice-mod}
+  The restriction of the localization \(\Sigma^{-1} M\) is a bounded
+  \(\mathfrak{g}\)-module of degree \(d\) with \(\operatorname{supp}
+  \Sigma^{-1} M = Q + \operatorname{supp} M\) and \(\dim \Sigma^{-1} M_\lambda
+  = d\) for all \(\lambda \in \operatorname{supp} \Sigma^{-1} M\).
+\end{proposition}
+
+\begin{proof}
+  Fix some \(\beta \in \Sigma\). We begin by showing that \(F_\beta\) and
+  \(F_\beta^{-1}\) map the weight space \(\Sigma^{-1} M_\lambda\) to
+  \(\Sigma^{-1} M_{\lambda - \beta}\) and \(\Sigma^{-1} M_{\lambda + \beta}\),
+  respectively. Indeed, given \(m \in M_\lambda\) and \(H \in \mathfrak{h}\) we
+  have
+  \[
+    H \cdot (F_\beta \cdot m)
+    = ([H, F_\beta] + F_\beta H) \cdot m
+    = F_\beta (-\beta(H) + H) \cdot m
+    = (\lambda - \beta)(H) F_\beta \cdot m
+  \]
+
+  On the other hand,
+  \[
+    0
+    = [H, 1]
+    = [H, F_\beta F_\beta^{-1}]
+    = F_\beta [H, F_\beta^{-1}] + [H, F_\beta] F_\beta^{-1}
+    = F_\beta [H, F_\beta^{-1}] - \beta(H) F_\beta F_\beta^{-1},
+  \]
+  so that \([H, F_\beta^{-1}] = \beta(H) \cdot F_\beta^{-1}\) and therefore
+  \[
+    H \cdot (F_\beta^{-1} \cdot m)
+    = ([H, F_\beta^{-1}] + F_\beta^{-1} H) \cdot m
+    = F_\beta^{-1} (\beta(H) + H) \cdot m
+    = (\lambda + \beta)(H) F_\beta^{-1} \cdot m
+  \]
+
+  From the fact that \(F_\beta^{\pm 1}\) maps \(M_\lambda\) to \(\Sigma^{-1}
+  M_{\lambda \pm \beta}\) follows our first conclusion: since \(M\) is a weight
+  module and every element of \(\Sigma^{-1} M\) has the form \(s^{-1} \cdot m =
+  s^{-1} \otimes m\) for \(s \in (F_\beta)_{\beta \in \Sigma}\) and \(m \in
+  M\), we can see that \(\Sigma^{-1} M = \bigoplus_\lambda \Sigma^{-1}
+  M_\lambda\). Furthermore, since the action of each \(F_\beta\) on
+  \(\Sigma^{-1} M\) is bijective and \(\Sigma\) is a basis for \(Q\) we obtain
+  \(\operatorname{supp} \Sigma^{-1} M = Q + \operatorname{supp} M\).
+
+  Again, because of the bijectivity of the \(F_\beta\)'s, to see that \(\dim
+  \Sigma^{-1} M_\lambda = d\) for all \(\lambda \in \operatorname{supp}
+  \Sigma^{-1} M\) it suffices to show that \(\dim \Sigma^{-1} M_\lambda = d\)
+  for some \(\lambda \in \operatorname{supp} \Sigma^{-1} M\). We may take
+  \(\lambda \in \operatorname{supp} M\) with \(\dim M_\lambda = d\). For any
+  finite-dimensional subspace \(V \subset \Sigma^{-1} M_\lambda\) we can find
+  \(s \in (F_\beta)_{\beta \in \Sigma}\) such that \(s \cdot V \subset M\). If
+  \(s = F_{\beta_{i_1}} \cdots F_{\beta_{i_r}}\), it is clear \(s \cdot V
+  \subset M_{\lambda - \beta_{i_1} - \cdots - \beta_{i_r}}\), so \(\dim V =
+  \dim s \cdot V \le d\). This holds for all finite-dimensional \(V \subset
+  \Sigma^{-1} M_\lambda\), so \(\dim \Sigma^{-1} M_\lambda \le d\). It then
+  follows from the fact that \(M_\lambda \subset \Sigma^{-1} M_\lambda\) that
+  \(M_\lambda = \Sigma^{-1} M_\lambda\) and therefore \(\dim \Sigma^{-1}
+  M_\lambda = d\).
+\end{proof}
+
+We now have a good candidate for a coherent extension of \(M\), but
+\(\Sigma^{-1} M\) is still not a coherent extension since its support is
+contained in a single \(Q\)-coset. In particular, \(\operatorname{supp}
+\Sigma^{-1} M \ne \mathfrak{h}^*\) and \(\Sigma^{-1} M\) is not a coherent
+family. To obtain a coherent family we thus need somehow extend \(\Sigma^{-1}
+M\). To that end, we will attempt to replicate the construction of the coherent
+extension of the \(\mathfrak{sl}_2(K)\)-module \(K[x, x^{-1}]\). Specifically,
+the idea is that if twist \(\Sigma^{-1} M\) by an automorphism which shifts its
+support by some \(\lambda \in \mathfrak{h}^*\), we can construct a coherent
+family by summing these modules over \(\lambda\) as in
+Example~\ref{ex:sl-laurent-family}.
+
+For \(K[x, x^{-1}]\) this was achieved by twisting the
+\(\operatorname{Diff}(K[x, x^{-1}])\)-module \(K[x, x^{-1}]\) by the
+automorphisms \(\varphi_\lambda : \operatorname{Diff}(K[x, x^{-1}]) \to
+\operatorname{Diff}(K[x, x^{-1}])\) and restricting the results to
+\(\mathcal{U}(\mathfrak{sl}_2(K))\) via the map
+\(\mathcal{U}(\mathfrak{sl}_2(K)) \to \operatorname{Diff}(K[x, x^{-1}])\), but
+this approach is inflexible since not every \(\mathfrak{sl}_2(K)\)-module
+factors through \(\operatorname{Diff}(K[x, x^{-1}])\). Nevertheless, we could
+just as well twist \(K[x, x^{-1}]\) by automorphisms of
+\(\mathcal{U}(\mathfrak{sl}_2(K))_f\) directly -- where
+\(\mathcal{U}(\mathfrak{sl}_2(K))_f = (f)^{-1} \mathcal{U}(\mathfrak{g})\) is
+the localization of \(\mathcal{U}(\mathfrak{sl}_2(K))\) by the multiplicative
+subset generated by \(f\).
+
+In general, we may twist the \(\Sigma^{-1} \mathcal{U}(\mathfrak{g})\)-module
+\(\Sigma^{-1} M\) by automorphisms of \(\Sigma^{-1}
+\mathcal{U}(\mathfrak{g})\). For \(\lambda = \beta \in \Sigma\) the map
+\begin{align*}
+  \theta_\beta : \Sigma^{-1} \mathcal{U}(\mathfrak{g}) & \to
+                 \Sigma^{-1} \mathcal{U}(\mathfrak{g}) \\
+                 u & \mapsto F_\beta u F_\beta^{-1}
+\end{align*}
+is a natural candidate for such a twisting automorphism. Indeed, we will soon
+see that \(\twisted{(\Sigma^{-1} M)}{\theta_\beta}_\lambda = \Sigma^{-1}
+M_{\lambda + \beta}\). However, this is hardly useful to us, since \(\beta \in
+Q\) and therefore \(\beta + \operatorname{supp} \Sigma^{-1} M =
+\operatorname{supp} \Sigma^{-1} M\). If we want to expand the support of
+\(\Sigma^{-1} M\) we will have to twist by automorphisms that shift its support
+by \(\lambda \in \mathfrak{h}^*\) lying \emph{outside} of \(Q\).
+
+The situation is much less obvious in this case. Nevertheless, it turns out we
+can extend the family \(\{\theta_\beta\}_{\beta \in \Sigma}\) to a family of
+automorphisms \(\{\theta_\lambda\}_{\lambda \in \mathfrak{h}^*}\).
+Explicitly\dots
+
+\begin{proposition}\label{thm:nice-automorphisms-exist}
+  There is a family of automorphisms \(\{\theta_\lambda : \Sigma^{-1}
+  \mathcal{U}(\mathfrak{g}) \to \Sigma^{-1}
+  \mathcal{U}(\mathfrak{g})\}_{\lambda \in \mathfrak{h}^*}\) such that
+  \begin{enumerate}
+    \item \(\theta_{k_1 \beta_1 + \cdots + k_r \beta_r}(u) = F_{\beta_1}^{k_1}
+      \cdots F_{\beta_r}^{k_r} u F_{\beta_r}^{- k_r} \cdots F_{\beta_1}^{-
+      k_1}\) for all \(u \in \Sigma^{-1} \mathcal{U}(\mathfrak{g})\) and \(k_1,
+      \ldots, k_r \in \mathbb{Z}\).
+
+    \item For each \(u \in \Sigma^{-1} \mathcal{U}(\mathfrak{g})\) the map
+      \begin{align*}
+        \mathfrak{h}^* & \to     \Sigma^{-1} \mathcal{U}(\mathfrak{g}) \\
+               \lambda & \mapsto \theta_\lambda(u)
+      \end{align*}
+      is polynomial.
+
+    \item If \(\lambda, \mu \in \mathfrak{h}^*\), \(N\) is a \(\Sigma^{-1}
+      \mathcal{U}(\mathfrak{g})\)-module whose restriction to
+      \(\mathcal{U}(\mathfrak{g})\) is a weight \(\mathfrak{g}\)-module and
+      \(\twisted{N}{\theta_\lambda}\) is the \(\Sigma^{-1}
+      \mathcal{U}(\mathfrak{g})\)-module \(N\) twisted by the automorphism
+      \(\theta_\lambda\) then \(N_\mu = \twisted{N}{\theta_\lambda}_{\mu +
+      \lambda}\). In particular, \(\operatorname{supp}
+      \twisted{N}{\theta_\lambda} = \lambda + \operatorname{supp} N\).
+  \end{enumerate}
+\end{proposition}
+
+\begin{proof}
+  Since the elements \(F_\beta\), \(\beta \in \Sigma\) commute with one
+  another, the endomorphisms
+  \begin{align*}
+    \theta_{k_1 \beta_1 + \cdots + k_r \beta_r}
+    : \Sigma^{-1} \mathcal{U}(\mathfrak{g}) &
+    \to \Sigma^{-1} \mathcal{U}(\mathfrak{g}) \\
+    u & \mapsto
+    F_{\beta_1}^{k_1} \cdots F_{\beta_r}^{k_r}
+    u
+    F_{\beta_1}^{- k_r} \cdots F_{\beta_r}^{- k_1}
+  \end{align*}
+  are well defined for all \(k_1, \ldots, k_r \in \mathbb{Z}\).
+
+  Fix some \(u \in \Sigma^{-1} \mathcal{U}(\mathfrak{g})\). For any \(s \in
+  (F_\beta)_{\beta \in \Sigma}\) and \(k > 0\) we have \(s^k u = \binom{k}{0}
+  \operatorname{ad}(s)^0 u s^{k - 0} + \cdots + \binom{k}{k}
+  \operatorname{ad}(s)^k u s^{k - k}\). Now if we take \(\ell\) such
+  \(\operatorname{ad}(F_\beta)^{\ell + 1} u = 0\) for all \(\beta \in \Sigma\)
+  we find
+  \[
+    \theta_{k_1 \beta_1 + \cdots + k_r \beta_r}(u)
+    = \sum_{i_1, \ldots, i_r = 1, \ldots, \ell}
+    \binom{k_1}{i_1} \cdots \binom{k_r}{i_r}
+    \operatorname{ad}(F_{\beta_1})^{i_1} \cdots
+    \operatorname{ad}(F_{\beta_r})^{i_r}
+    u
+    F_{\beta_1}^{- i_1} \cdots F_{\beta_r}^{- i_r}
+  \]
+  for all \(k_1, \ldots, k_r \in \mathbb{N}\).
+
+  Since the binomial coefficients \(\binom{x}{k} = \frac{x (x-1) \cdots (x - k
+  + 1)}{k!}\) can be uniquely extended to polynomial functions in \(x \in K\),
+  we may in general define
+  \[
+    \theta_\lambda(u)
+    = \sum_{i_1, \ldots, i_r \ge 0}
+    \binom{\lambda_1}{i_1} \cdots \binom{\lambda_r}{i_r}
+    \operatorname{ad}(F_{\beta_1})^{i_1} \cdots
+    \operatorname{ad}(F_{\beta_r})^{i_r}
+    r
+    F_{\beta_1}^{- i_1} \cdots F_{\beta_r}^{- i_r}
+  \]
+  for \(\lambda_1, \ldots, \lambda_r \in K\), \(\lambda = \lambda_1 \beta_1 +
+  \cdots + \lambda_r \beta_r \in \mathfrak{h}^*\).
+
+  It is clear that the \(\theta_\lambda\) are endomorphisms. To see that the
+  \(\theta_\lambda\) are indeed automorphisms, notice \(\theta_{- k_1 \beta_1 -
+  \cdots - k_r \beta_r} = \theta_{k_1 \beta_1 + \cdots + k_r \beta_r}^{-1}\).
+  The uniqueness of the polynomial extensions then implies \(\theta_{- \lambda}
+  = \theta_\lambda^{-1}\) in general: given \(u \in \Sigma^{-1}
+  \mathcal{U}(\mathfrak{g})\), the map
+  \begin{align*}
+    \mathfrak{h}^* & \to \Sigma^{-1} \mathcal{U}(\mathfrak{g})        \\
+           \lambda & \mapsto \theta_\lambda(\theta_{-\lambda}(u)) - u
+  \end{align*}
+  is a polynomial extension of the zero map \(\mathbb{Z} \beta_1 \oplus \cdots
+  \oplus \mathbb{Z} \beta_r \to \Sigma^{-1} \mathcal{U}(\mathfrak{g})\) and is
+  therefore identically zero.
+
+  Finally, let \(N\) be a \(\Sigma^{-1} \mathcal{U}(\mathfrak{g})\)-module
+  whose restriction is a weight module. If \(n \in N\) then
+  \[
+    n \in \twisted{N}{\theta_\lambda}_{\mu + \lambda}
+    \iff \theta_\lambda(H) \cdot n = (\mu + \lambda)(H) n
+    \, \forall H \in \mathfrak{h}
+  \]
+
+  But
+  \[
+    \theta_\beta(H)
+    = F_\beta H F_\beta^{-1}
+    = ([F_\beta, H] + H F_\beta) F_\beta^{-1}
+    = (\beta(H) + H) F_\beta F_\beta^{-1}
+    = \beta(H) + H
+  \]
+  for all \(H \in \mathfrak{h}\) and \(\beta \in \Sigma\). In general,
+  \(\theta_\lambda(H) = \lambda(H) + H\) for all \(\lambda \in \mathfrak{h}^*\)
+  and hence
+  \[
+    \begin{split}
+      n \in \twisted{N}{\theta_\lambda}_{\mu + \lambda}
+      & \iff (\lambda(H) + H) \cdot n = (\mu + \lambda)(H) n
+        \; \forall H \in \mathfrak{h} \\
+      & \iff H \cdot n = \mu(H) n \; \forall H \in \mathfrak{h} \\
+      & \iff n \in N_\mu
+    \end{split},
+  \]
+  so that \(\twisted{N}{\theta_\lambda}_{\mu + \lambda} = N_\mu\).
+\end{proof}
+
+It should now be obvious\dots
+
+\begin{proposition}[Mathieu]
+  There exists a coherent extension \(\mathcal{M}\) of \(M\).
+\end{proposition}
+
+\begin{proof}
+  Take\footnote{Here we fix some $\lambda_\xi \in \xi$ for each $Q$-coset $\xi
+  \in \mfrac{\mathfrak{h}^*}{Q}$. While there is a natural isomorphism
+  $\twisted{(\Sigma^{-1} M)}{\theta_\lambda} \isoto \twisted{(\Sigma^{-1}
+  M)}{\theta_\mu}$ for each $\mu \in \lambda + Q$, they are not the same
+  \(\mathfrak{g}\)-modules strictly speaking. This is yet another obstruction
+  to the functoriality of our constructions.}
+  \[
+    \mathcal{M}
+    = \bigoplus_{\lambda + Q \in \mfrac{\mathfrak{h}^*}{Q}}
+      \twisted{(\Sigma^{-1} M)}{\theta_\lambda}
+  \]
+
+  It is clear \(M\) lies in \(\Sigma^{-1} M = \twisted{(\Sigma^{-1}
+  M)}{\theta_0}\) and therefore \(M \subset \mathcal{M}\). On the other hand,
+  \(\dim \mathcal{M}_\mu = \dim \twisted{(\Sigma^{-1} M)}{\theta_\lambda}_\mu =
+  \dim \Sigma^{-1} M_{\mu - \lambda} = d\) for all \(\mu \in \lambda + Q\) --
+  \(\lambda\) standing for some fixed representative of its \(Q\)-coset.
+  Furthermore, given \(u \in \mathcal{U}(\mathfrak{g})_0\) and \(\mu \in
+  \lambda + Q\),
+  \[
+    \operatorname{Tr}(u\!\restriction_{\mathcal{M}_\mu})
+    = \operatorname{Tr}
+      (\theta_\lambda(u)\!\restriction_{\Sigma^{-1} M_{\mu - \lambda}})
+  \]
+  is polynomial in \(\mu\) because of the second item of
+  Proposition~\ref{thm:nice-automorphisms-exist}.
+\end{proof}
+
+Lo and behold\dots
+
+\begin{theorem}[Mathieu]\index{coherent family!Mathieu's \(\mExt\) coherent extension}
+  There exists a unique semisimple coherent extension \(\mExt(M)\) of \(M\).
+  More precisely, if \(\mathcal{M}\) is any coherent extension of \(M\), then
+  \(\mathcal{M}^{\operatorname{ss}} \cong \mExt(M)\). Furthermore, \(\mExt(M)\)
+  is a irreducible coherent family.
+\end{theorem}
+
+\begin{proof}
+  The existence part should be clear from the previous discussion: it suffices
+  to fix some coherent extension \(\mathcal{M}\) of \(M\) and take
+  \(\mExt(M) = \mathcal{M}^{\operatorname{ss}}\).
+
+  To see that \(\mExt(M)\) is irreducible, recall from
+  Corollary~\ref{thm:bounded-is-submod-of-extension} that \(M\) is a
+  \(\mathfrak{g}\)-submodule of \(\mExt(M)\). Since the degree of \(M\) is the
+  same as the degree of \(\mExt(M)\), some of its weight spaces have maximal
+  dimension inside of \(\mExt(M)\). In particular, it follows from
+  Lemma~\ref{thm:centralizer-multiplicity} that \(\mExt(M)_\lambda =
+  M_\lambda\) is a simple \(\mathcal{U}(\mathfrak{g})_0\)-module for some
+  \(\lambda \in \operatorname{supp} M\).
+
+  As for the uniqueness of \(\mExt(M)\), fix some other semisimple coherent
+  extension \(\mathcal{N}\) of \(M\). We claim that the multiplicity of a given
+  simple \(\mathfrak{g}\)-module \(L\) in \(\mathcal{N}\) is determined by its
+  \emph{trace function}
+  \begin{align*}
+    \mathfrak{h}^* \times \mathcal{U}(\mathfrak{g})_0 &
+    \to K \\
+    (\lambda, u) &
+    \mapsto \operatorname{Tr}(u\!\restriction_{\mathcal{N}_\lambda})
+  \end{align*}
+
+  It is a well known fact of the theory of modules that, given an associative
+  \(K\)-algebra \(A\), a finite-dimensional semisimple \(A\)-module \(L\) is
+  determined, up to isomorphism, by its \emph{character}
+  \begin{align*}
+    \chi_L : A & \to     K                                    \\
+             a & \mapsto \operatorname{Tr}(a\!\restriction_L)
+  \end{align*}
+
+  In particular, the multiplicity of \(L\) in \(\mathcal{N}\), which is the
+  same as the multiplicity of \(L_\lambda\) in \(\mathcal{N}_\lambda\), is
+  determined by the character \(\chi_{\mathcal{N}_\lambda} :
+  \mathcal{U}(\mathfrak{g})_0 \to K\). Since this holds for all simple weight
+  \(\mathfrak{g}\)-modules, it follows that \(\mathcal{N}\) is determined by
+  its trace function. Of course, the same holds for \(\mExt(M)\). We now claim
+  that the trace function of \(\mathcal{N}\) is the same as that of
+  \(\mExt(M)\). Clearly,
+  \(\operatorname{Tr}(u\!\restriction_{\mExt(M)_\lambda}) =
+  \operatorname{Tr}(u\!\restriction_{M_\lambda}) =
+  \operatorname{Tr}(u\!\restriction_{\mathcal{N}_\lambda})\) for all \(\lambda
+  \in \operatorname{supp}_{\operatorname{ess}} M\), \(u \in
+  \mathcal{U}(\mathfrak{g})_0\). Since the essential support of \(M\) is
+  Zariski-dense and the maps \(\lambda \mapsto
+  \operatorname{Tr}(u\!\restriction_{\mExt(M)_\lambda})\) and \(\lambda \mapsto
+  \operatorname{Tr}(u\!\restriction_{\mathcal{N}_\lambda})\) are polynomial in
+  \(\lambda \in \mathfrak{h}^*\), it follows that these maps coincide for all
+  \(u\).
+
+  In conclusion, \(\mathcal{N} \cong \mExt(M)\) and \(\mExt(M)\) is unique.
+\end{proof}
+
+% This is a very important theorem, but since we won't classify the coherent
+% extensions in here we don't need it, and there is no other motivation behind
+% it. Including this would also require me to explain what central characters
+% are, which is a bit of a pain
+%\begin{proposition}[Mathieu]
+%  The central characters of the irreducible submodules of
+%  \(\operatorname{Ext}(M)\) are all the same.
+%\end{proposition}
+
+We have thus concluded our classification of cuspidal modules in terms of
+coherent families. Of course, to get an explicit construction of all simple
+\(\mathfrak{g}\)-modules we would have to classify the irreducible semisimple
+coherent \(\mathfrak{g}\)-families themselves, which is the subject of sections
+7, 8 and 9 of \cite{mathieu}. In addition, in sections 11 and 12 of
+\cite{mathieu} Mathieu provides an explicit construction of coherent families.
+We unfortunately do not have the necessary space to discuss these results in
+detail, but we will now provide a brief overview.
+
+First and foremost, notice that because of
+Example~\ref{thm:simple-weight-mod-is-tensor-prod} the problem of classifying
+the simple weight \(\mathfrak{g}\)-modules can be reduced to that of
+classifying the simple weight modules of its simple components. In addition, it
+turns out that very few simple Lie algebras admit cuspidal modules at all.
+Specifically\dots
+
+\begin{proposition}[Fernando]\label{thm:only-sl-n-sp-have-cuspidal}
+  Let \(\mathfrak{s}\) be a finite-dimensional simple Lie algebra. Suppose
+  there exists a simple cuspidal \(\mathfrak{s}\)-module. Then \(\mathfrak{s}
+  \cong \mathfrak{sl}_n(K)\) or \(\mathfrak{s} \cong \mathfrak{sp}_{2 n}(K)\).
+\end{proposition}
+
+Hence it suffices to classify the irreducible semisimple coherent families of
+\(\mathfrak{sl}_n(K)\) and \(\mathfrak{sp}_{2 n}(K)\). These can be described
+either algebraically, using combinatorial invariants -- which Mathieu does in
+sections 7, 8 and 9 of his paper -- or geometrically, using algebraic varieties
+and differential forms -- which is done in sections 11 and 12. While rather
+complicated on its own, the geometric construction is more concrete than its
+combinatorial counterpart.
+
+This construction also brings us full circle to the beginning of these notes,
+where we saw in Proposition~\ref{thm:geometric-realization-of-uni-env} that
+\(\mathfrak{g}\)-modules may be understood as geometric objects. In fact,
+throughout the previous four chapters we have seen a tremendous number of
+geometrically motivated examples, which further emphasizes the connection
+between representation theory and geometry. I would personally go as far as
+saying that the beautiful interplay between the algebraic and the geometric is
+precisely what makes representation theory such a fascinating and charming
+subject.
+
+Alas, our journey has come to an end. All it is left is to wonder at the beauty
+of Lie algebras and their representations.
+
+\label{end-47}
diff --git a/tcc.tex b/tcc.tex
@@ -33,9 +33,9 @@
 
 \input{sections/sl2-sl3}
 
-\input{sections/semisimple-algebras}
+\input{sections/fin-dim-simple}
 
-\input{sections/mathieu}
+\input{sections/simple-weight}
 
 \printbibliography
 \printindex