- Commit
- 121602d7475bc83034c7194adb34ea2ba339e80e
- Parent
- bf3eff9730bb0b7a15ff78a323b6a5e04bf8a353
- Author
- Pablo <pablo-escobar@riseup.net>
- Date
Revised the forth chapter
Also implemented Kashuba's suggestions
Source code for my notes on representations of semisimple Lie algebras and Olivier Mathieu's classification of simple weight modules
Revised the forth chapter
Also implemented Kashuba's suggestions
1 file changed, 121 insertions, 129 deletions
Status | File Name | N° Changes | Insertions | Deletions |
Modified | sections/semisimple-algebras.tex | 250 | 121 | 129 |
diff --git a/sections/semisimple-algebras.tex b/sections/semisimple-algebras.tex @@ -12,11 +12,11 @@ This was simple enough to do in the case of \(\mathfrak{sl}_2(K)\), but the reasoning behind it, as well as the mere fact equation (\ref{sym-diag}) holds, are harder to explain in the case of \(\mathfrak{sl}_3(K)\). The eigenspace decomposition associated with an operator \(V \to V\) is a very well-known -tool, and this type of argument should be familiar to anyone familiar with -basic concepts of linear algebra. On the other hand, the eigenspace -decomposition of \(V\) with respect to the action of an arbitrary subalgebra -\(\mathfrak{h} \subset \mathfrak{gl}(V)\) is neither well-known nor does it -hold in general: as previously stated, it may very well be that +tool, and readers familizared with basic concepts of linear algebra should be +used to this type of argument. On the other hand, the eigenspace decomposition +of \(V\) with respect to the action of an arbitrary subalgebra \(\mathfrak{h} +\subset \mathfrak{gl}(V)\) is neither well-known nor does it hold in general: +as previously stated, it may very well be that \[ \bigoplus_{\lambda \in \mathfrak{h}^*} V_\lambda \subsetneq V \] @@ -35,13 +35,13 @@ 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 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 representation \(V\) of -\(\mathfrak{g}\) into subspaces invariant under the action of \(\mathfrak{h}\), -and then analyze how the remaining elements of \(\mathfrak{g}\) act on this -subspaces. The bigger \(\mathfrak{h}\) the simpler our problem gets, because -there are fewer elements outside of \(\mathfrak{h}\) left to analyze. +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 representation \(V\) of \(\mathfrak{g}\) into subspaces +invariant under the action of \(\mathfrak{h}\), and then analyze how the +remaining elements of \(\mathfrak{g}\) act on this 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. @@ -64,9 +64,9 @@ Hence we are generally interested in maximal Abelian subalgebras \(\mathfrak{h} \begin{proof} Notice that \(0 \subset \mathfrak{g}\) is an Abelian subalgebra whose elements act as diagonal operators via the adjoint representation. Indeed, - \(0\) -- the only element of \(0 \subset \mathfrak{g}\) -- is such that - \(\operatorname{ad}(0) = 0\). Furthermore, given a chain of Abelian - subalgebras + \(0\), 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 \[ 0 \subset \mathfrak{h}_1 \subset \mathfrak{h}_2 \subset \cdots \] @@ -75,7 +75,7 @@ Hence we are generally interested in maximal Abelian subalgebras \(\mathfrak{h} \mathfrak{g}\) 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 -- i.e. a Cartan subalgebra. + properties, also known as a Cartan subalgebra. \end{proof} We have already seen some concrete examples. For instance, one can readily @@ -100,9 +100,9 @@ self-normalizing. Hence \(\mathfrak{h}\) is a Cartan subalgebra of 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 two sections. The remaining question then -is: if \(\mathfrak{h} \subset \mathfrak{g}\) is a Cartan subalgebra and \(V\) -is a representation of \(\mathfrak{g}\), does the eigenspace decomposition +subalgebras described the previous chapter. The remaining question then is: if +\(\mathfrak{h} \subset \mathfrak{g}\) is a Cartan subalgebra and \(V\) is a +representation of \(\mathfrak{g}\), does the eigenspace decomposition \[ V = \bigoplus_{\lambda \in \mathfrak{h}^*} V_\lambda \] @@ -126,9 +126,7 @@ What is simultaneous diagonalization all about then? 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 --- where \(n\) is, of course, finite. - +framed as an equivalent statement about diagonalizable \(n \times n\) matrices. Simultaneous diagonalization implies that to show \(V = \bigoplus_\lambda V_\lambda\) it suffices to show that \(H\!\restriction_V : V \to V\) is a diagonalizable operator for each \(H \in \mathfrak{h}\). To that end, we @@ -165,12 +163,8 @@ words\dots \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}. We should point out -this fails spectacularly in positive characteristic. Furthermore, the statement -of proposition~\ref{thm:preservation-jordan-form} only makes sense for -\emph{semisimple} Lie algebras -- i.e. the algebras \(\mathfrak{g}\) for which -the abstract Jordan decomposition of \(\mathfrak{g}\) is defined. Nevertheless, -as promised this implies\dots +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 @@ -214,7 +208,7 @@ Cartan subalgebra of \(\mathfrak{g}\) is \(\mathfrak{g}\) itself, and a \(\mathfrak{g}\)-module is simply a vector space \(V\) endowed with an operator \(V \to V\) -- which corresponds to the action of \(1 \in \mathfrak{g}\) on \(V\). In particular, if we choose an operator \(V \to V\) which is \emph{not} -diagonalizable we find \(V \ne \bigoplus_{\lambda \in \mathfrak{h}^*} +diagonalizable we find \(V \ne 0 = \bigoplus_{\lambda \in \mathfrak{h}^*} V_\lambda\). However, corollary~\ref{thm:finite-dim-is-weight-mod} does work for reductive @@ -224,14 +218,16 @@ representations as scalar operators. The hypothesis of finite-dimensionality is also of huge importance. In the next chapter we will encounter infinite-dimensional \(\mathfrak{g}\)-modules for which the eigenspace decomposition \(V = \bigoplus_{\lambda \in \mathfrak{h}^*} V_\lambda\) fails. -As a first consequence of corollary~\ref{thm:finite-dim-is-weight-mod} +As a first consequence of corollary~\ref{thm:finite-dim-is-weight-mod} we +show\dots \begin{corollary} - The restriction of \(B\) to \(\mathfrak{h}\) is non-degenerate. + The restriction of the Killing form \(B\) to \(\mathfrak{h}\) is + non-degenerate. \end{corollary} \begin{proof} - Consider the eigenspace decomposition \(\mathfrak{g} = \mathfrak{g}_0 \oplus + Consider the root space decomposition \(\mathfrak{g} = \mathfrak{g}_0 \oplus \bigoplus_\alpha \mathfrak{g}_\alpha\) of the adjoint representation, where \(\alpha\) ranges over all nonzero eigenvalues of the adjoint action of \(\mathfrak{h}\). We claim \(\mathfrak{g}_0 = \mathfrak{h}\). @@ -270,7 +266,7 @@ hardly ever a non-degenerate form. \begin{note} Since \(B\) induces an isomorphism \(\mathfrak{h} \isoto \mathfrak{h}^*\), it induces a bilinear form \((B(X, \cdot), B(Y, \cdot)) \mapsto B(X, Y)\) in - \(\mathfrak{h}^*\). We denote this form by \(B\). + \(\mathfrak{h}^*\). We denote this form by \(B\) as well. \end{note} We now have most of the necessary tools to reproduce the results of the @@ -278,8 +274,8 @@ previous chapter in a general setting. Let \(\mathfrak{g}\) be a finite-dimensional semisimple algebra with a Cartan subalgebra \(\mathfrak{h}\) and let \(V\) be a finite-dimensional irreducible representation of \(\mathfrak{g}\). We will proceed, as we did before, by generalizing the -results about of the previous two sections in order. By now the pattern should -be starting become clear, so we will mostly omit technical details and proofs +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}. @@ -290,24 +286,23 @@ We begin our analysis, as we did for \(\mathfrak{sl}_2(K)\) and \(\mathfrak{g}\). Throughout chapter~\ref{ch:sl3} we've seen that the weights of any given finite-dimensional representation of \(\mathfrak{sl}_2(K)\) or \(\mathfrak{sl}_3(K)\) can only assume very rigid configurations. For instance, -we've seen that the weights of any given representation are symmetric with -respect to the origin. In this chapter we will generalize most results from -chapter~\ref{ch:sl3} the rigidity of the geometry of the set of weights of a -given representations. +we've 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 representations. As for the afford mentioned 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 eigenvalues \(\alpha\) of the adjoint action of \(\mathfrak{h}\) on - \(\mathfrak{g}\) are symmetrical about the origin -- i.e. \(- \alpha\) is - also an eigenvalue -- and they span all of \(\mathfrak{h}^*\). + 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'll start with the first claim. Let \(\alpha\) and \(\beta\) be two - eigenvalues of the adjoint action of \(\mathfrak{h}\). Notice + 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 @@ -329,17 +324,16 @@ restriction of the Killing form to the Cartan subalgebra. for \(n\) large enough. In particular, \(B(X, Y) = \operatorname{Tr}(\operatorname{ad}(X) \operatorname{ad}(Y)) = 0\). Now if \(- \alpha\) is not an eigenvalue we find \(B(X, \mathfrak{g}_\beta) = 0\) - for all eigenvalues \(\beta\), which contradicts the non-degeneracy of \(B\). + for all roots \(\beta\), which contradicts the non-degeneracy of \(B\). Hence \(- \alpha\) must be an eigenvalue of the adjoint action of \(\mathfrak{h}\). - For the second statement, note that if the eigenvalues of \(\mathfrak{h}\) do - not span all of \(\mathfrak{h}^*\) then there is some \(H \in \mathfrak{h}\) - nonzero such that \(\alpha(H) = 0\) for all eigenvalues \(\alpha\), which is + 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 \(\mathfrak{z}\) of \(\mathfrak{g}\), which is zero by the - semisimplicity -- a contradiction. + the center \(\mathfrak{z} = 0\) of \(\mathfrak{g}\), a contradiction. \end{proof} Furthermore, as in the case of \(\mathfrak{sl}_2(K)\) and @@ -351,26 +345,24 @@ Furthermore, as in the case of \(\mathfrak{sl}_2(K)\) and The proof of the first statement of proposition~\ref{thm:weights-symmetric-span} highlights something interesting: -if we fix some 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 \(v \in -V_\lambda\) we find +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 \(v \in V_\lambda\) we find \[ H (X v) = X (H v) + [H, X] v = (\lambda + \alpha)(H) \cdot X v \] so that \(X\) carries \(v\) to \(V_{\lambda + \alpha}\). We have encountered -this formula twice in this chapter: again, we find \(\mathfrak{g}_\alpha\) -\emph{acts on \(V\) by translating vectors between eigenspaces}. In other -words, if we denote by \(\Delta\) the set of all roots of \(\mathfrak{g}\) -then\dots +this formula twice in these notes: again, we find \(\mathfrak{g}_\alpha\) +\emph{acts on \(V\) 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} The weights of an irreducible representation \(V\) of \(\mathfrak{g}\) are all congruent module the root lattice \(Q = \ZZ \Delta\) of \(\mathfrak{g}\). - In other words, all weights of \(V\) lie in the same \(Q\)-coset in - \(\mfrac{\mathfrak{h}^*}{Q}\). + In other words, all weights of \(V\) lie in the same \(Q\)-coset + \(t \in \mfrac{\mathfrak{h}^*}{Q}\). \end{theorem} Again, we may leverage our knowledge of \(\mathfrak{sl}_2(K)\) to obtain further @@ -418,7 +410,8 @@ restrictions on the weights of \(V\). Namely, if \(\lambda\) is a weight, \begin{definition}\label{def: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 \(\mathfrak{g}\)}. + \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} @@ -426,20 +419,17 @@ restrictions on the weights of \(V\). Namely, if \(\lambda\) is a weight, in the weight lattice \(P\). \end{proposition} -We call the elements of \(P\) \emph{integral}. 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 \(\mathbb{Z} \alpha_1 \oplus \mathbb{Z} \alpha_2 \oplus -\mathbb{Z} \alpha_3\). - -To proceed further, we would like to take \emph{the highest weight of \(V\)} 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 \(V\) that maximizes this functional, but at this point -it is convenient to introduce some additional tools to our arsenal. This tools -are called \emph{basis}. +\mathbb{Z} \alpha_3\). To proceed further, we would like to take \emph{the +highest weight of \(V\)} 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 \(V\) that +maximizes this functional, but at this point it is convenient to introduce some +additional tools to our arsenal. This tools are called \emph{basis}. \begin{definition}\label{def:basis-of-root} A subset \(\Sigma = \{\beta_1, \ldots, \beta_k\} \subset \Delta\) of linearly @@ -450,9 +440,9 @@ are called \emph{basis}. The interesting thing about basis for \(\Delta\) is that they allow us to compare weights of a given representation. At this point the reader should be -asking himself: how? Definition~\ref{def:basis-of-root} doesn't exactly scream -``comparison''. Well, the thing is that any choice of basis induces a partial -order in \(Q\), where elements are ordered by their \emph{heights}. +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} Let \(\Sigma = \{\beta_1, \ldots, \beta_k\}\) be a basis for \(\Delta\). @@ -484,8 +474,8 @@ order in \(Q\), where elements are ordered by their \emph{heights}. 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, -we say \(\lambda \preceq \mu\) if \(\lambda - \mu \preceq 0\) for \(\lambda \in -\mu + Q\). In particular, since the weights of \(V\) all lie in a single +given \(\lambda \in \mu + Q\), we say \(\lambda \preceq \mu\) if \(\lambda - +\mu \preceq 0\). In particular, since the weights of \(V\) all lie in a single \(Q\)-coset, we may compare them in this fashion. Given a basis \(\Sigma\) for \(\Delta\) we may take ``the highest weight of \(V\)'' as a maximal weight \(\lambda\) of \(V\). The obvious question then is: can we always find a basis @@ -499,14 +489,15 @@ 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 \(B(\alpha, \beta) \in \mathbb{Z}\) for all \(\alpha, \beta \in \Delta\), so that the Killing form -\(B\) restricts to a nondegenerate 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. +\(B\) restricts to a nondegenerate \(\mathbb{Q}\)-linear 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 @@ -558,8 +549,8 @@ This has a number of important consequences. For instance\dots \end{proof} This is entirely analogous to the situation of \(\mathfrak{sl}_3(K)\), where we -found that the weights of the irreducible representations formed a continuous -string symmetric with respect to the lines \(K \alpha\) with \(B(\alpha_i - +found that the weights of the irreducible representations formed continuous +strings symmetric with respect to the lines \(K \alpha\) with \(B(\alpha_i - \alpha_j, \alpha) = 0\). As in the case of \(\mathfrak{sl}_3(K)\), the same class of arguments leads us to the conclusion\dots @@ -577,19 +568,19 @@ class of arguments leads us to the conclusion\dots \(\mathcal{W}\). \end{theorem} -Aside from showing up the previous theorem, the Weyl group will also play an +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 \(\mathcal{W}\) on \(\mathfrak{h}\). By its very nature +canonical action of \(\mathcal{W}\) on \(\mathfrak{h}\). By its very nature, \(\mathcal{W}\) acts in \(\mathfrak{h}^*\). If we conjugate the action \(\sigma\!\restriction_{\mathfrak{h}^*} : \mathfrak{h}^* \isoto \mathfrak{h}^*\) of some \(\sigma \in \mathcal{W}\) by the isomorphism \(\mathfrak{h}^* \isoto \mathfrak{h}\) afforded by the restriction of the -Killing for to \(\mathfrak{h}\) we get a linear automorphism \(\mathfrak{h} -\isoto \mathfrak{h}\). As it turns out, the automorphism -\(\sigma\!\restriction_{\mathfrak{h}} : \mathfrak{h} \isoto \mathfrak{h}\) 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}. +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 $\mathfrak{g}$ @@ -741,7 +732,7 @@ Moreover, we find\dots \[ M(\lambda) \subset - \bigoplus_{\substack{k_i \in \ZZ \\ k_i \ge 0}} + \bigoplus_{k_i \in \mathbb{N}} M(\lambda)_{\lambda + k_1 \cdot \alpha_1 + \cdots + k_n \cdot \alpha_n} \] where \(\{\alpha_1, \ldots, \alpha_m\} = \Delta^-\), so that all weights of @@ -786,9 +777,9 @@ Moreover, we find\dots \bigoplus_{k \ge 0} K f^k v^+\), and the action of \(\mathfrak{sl}_2(K)\) on \(M(\lambda)\) is given by \begin{align*} - f^{k + 1} v^+ & \overset{e}{\mapsto} (2 - k (k - 1)) f^k v^+ & - f^{k + 1} v^+ & \overset{f}{\mapsto} f^{k + 2} v^+ & - f^{k + 1} v^+ & \overset{h}{\mapsto} - 2 k f^{k + 1} v^+ & + f^k v^+ & \overset{e}{\mapsto} (2 - k (k + 1)) f^{k - 1} v^+ & + f^k v^+ & \overset{f}{\mapsto} f^{k + 1} v^+ & + f^k v^+ & \overset{h}{\mapsto} - 2 (k - 1) f^k v^+ & \end{align*} In the language of the diagrams used in chapter~\ref{ch:sl3}, we write @@ -819,7 +810,7 @@ whose highest weight is \(\lambda\). Every subrepresentation \(V \subset M(\lambda)\) is the direct sum of its weight spaces. In particular, \(M(\lambda)\) has a unique maximal subrepresentation \(N(\lambda)\) and a unique irreducible quotient - \(\sfrac{M(\lambda)}{N(\lambda)}\). + \(L(\lambda) = \sfrac{M(\lambda)}{N(\lambda)}\). \end{proposition} \begin{proof} @@ -856,11 +847,11 @@ whose highest weight is \(\lambda\). \in V \end{multline*} - By applying the same procedure over and over again we can see that \(v_1 = X - v \in V\) for some \(X \in \mathcal{U}(\mathfrak{g})\). Furthermore, if we + By applying the same procedure over and over again we can see that \(v_1 = u + v \in V\) for some \(u \in \mathcal{U}(\mathfrak{g})\). Furthermore, if we reproduce all this for \(v_2 + \cdots + v_n = v - v_1 \in V\) we get that - \(v_2 \in V\). Now by applying the same procedure over and over we find - \(v_1, \ldots, v_n \in V\). Hence + \(v_2 \in V\). Now by interating this procedure we find \(v_1, \ldots, v_n + \in V\). Hence \[ V = \bigoplus_\mu V_\mu = \bigoplus_\mu M(\lambda)_\mu \cap V \] @@ -868,57 +859,58 @@ whose highest weight is \(\lambda\). Since \(M(\lambda) = \mathcal{U}(\mathfrak{g}) \cdot v^+\), \(V\) is a proper subrepresentation then \(v^+ \notin V\). 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. In fact, this implies + \(N(\lambda)\) of all such submodules is still proper. This implies \(N(\lambda)\) is the unique maximal subrepresentation of \(M(\lambda)\) and - \(\sfrac{M(\lambda)}{N(\lambda)}\) is its unique irreducible quotient. + \(L(\lambda) = \sfrac{M(\lambda)}{N(\lambda)}\) is its unique irreducible + 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 \(N(\lambda) = \bigoplus_{k \ge - 3} K f^k v^+\), so that \(\sfrac{M(\lambda)}{N(\lambda)}\) is the - \(3\)-dimensional irreducible representation of \(\mathfrak{sl}_2(K)\) -- - i.e. the finite-dimensional irreducible representation with highest weight - \(\lambda\) constructed in chapter~\ref{ch:sl3}. + 3} K f^k v^+\), so that \(L(\lambda)\) is the \(3\)-dimensional irreducible + representation of \(\mathfrak{sl}_2(K)\) -- i.e. the finite-dimensional + irreducible representation with highest weight \(\lambda\) constructed in + chapter~\ref{ch:sl3}. \end{example} This last example is particularly interesting to us, since it indicates that the finite-dimensional irreducible representations of \(\mathfrak{sl}_2(K)\) as quotients of Verma modules. This is because the quotient \(\sfrac{M(\lambda)}{N(\lambda)}\) in example~\ref{ex:sl2-verma-quotient} -happened to be finite-dimensional. As it turns out, this is always the case for -semisimple \(\mathfrak{g}\). Namely\dots +is finite-dimensional. As it turns out, this is not a coincidence. \begin{proposition}\label{thm:verma-is-finite-dim} - If \(\lambda\) is dominant integral then the unique irreducible quotient of - \(M(\lambda)\) is finite-dimensional. + If \(\mathfrak{g}\) is semisimple and \(\lambda\) is dominant integral then + the unique irreducible quotient 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 \(\sfrac{M(\lambda)}{N(\lambda)}\) is stable under the natural -action of the Weyl group \(\mathcal{W}\) on \(\mathfrak{h}^*\). One can then -show that the every weight of \(V\) is conjugate to a single dominant integral -weight of \(\sfrac{M(\lambda)}{N(\lambda)}\), and that the set of dominant -integral weights of such irreducible quotient is finite. Since \(W\) is -finitely generated, this implies the set of weights of the unique irreducible -quotient of \(M(\lambda)\) is finite. But each weight space is -finite-dimensional. Hence so is the irreducible quotient. +weights of \(L(\lambda)\) is stable under the natural action of the Weyl group +\(\mathcal{W}\) on \(\mathfrak{h}^*\). One can then show that the every weight +of \(V\) is conjugate to a single dominant integral weight of +\(\sfrac{M(\lambda)}{N(\lambda)}\), and that the set of dominant integral +weights of such irreducible quotient is finite. Since \(W\) is finitely +generated, this implies the set of weights of the unique irreducible quotient +of \(M(\lambda)\) is finite. But each weight space is finite-dimensional. Hence +so is the irreducible quotient. We refer the reader to \cite[ch. 21]{humphreys} for further details. What we are really interested in is\dots \begin{corollary} - There is a finite-dimensional irreducible \(\mathfrak{g}\)-module \(V\) whose + Let \(\lambda\) be a dominant integral weight of \(\mathfrak{g}\). Then there + is a finite-dimensional irreducible \(\mathfrak{g}\)-module \(V\) whose highest weight is \(\lambda\). \end{corollary} \begin{proof} - Let \(V = \mfrac{M(\lambda)}{N(\lambda)}\). It suffices to show that its - highest weight is \(\lambda\). We have already seen that \(v^+ \in - M(\lambda)_\lambda\) is a highest weight vector. Now since \(v\) lies outside - of the maximal subrepresentation of \(M(\lambda)\), the projection \(v^+ + - N(\lambda) \in V\) is nonzero. + Let \(V = L(\lambda)\). It suffices to show that its highest weight is + \(\lambda\). We have already seen that \(v^+ \in M(\lambda)_\lambda\) is a + highest weight vector. Now since \(v\) lies outside of the maximal + subrepresentation of \(M(\lambda)\), the projection \(v^+ + N(\lambda) \in + V\) is nonzero. We now claim that \(v^+ + N(\lambda) \in V_\lambda\). Indeed, \[ @@ -929,15 +921,15 @@ are really interested in is\dots for all \(H \in \mathfrak{h}\). Hence \(\lambda\) is a weight of \(V\), with weight vector \(v^+ + N(\lambda)\). Finally, we remark that \(\lambda\) is the highest weight of \(V\), for if this was not the case we could find a - weight \(\mu\) of \(M(\lambda)\) which is higher than \(\lambda\). + weight \(\mu\) of \(M(\lambda)\) with \(\mu \succ \lambda\). \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 show that if \(\lambda \in P\) is not dominant then \(N(\lambda) = 0\) and \(M(\lambda)\) is irreducible. For instance, if -\(\mathfrak{g} = \mathfrak{sl}_2(K)\) and \(\lambda = -2\) then the action of -\(\mathfrak{g}\) on \(M(\lambda)\) is given by +\(\mathfrak{g} = \mathfrak{sl}_2(K)\) and \(\lambda : h \mapsto -2\) then the +action of \(\mathfrak{g}\) on \(M(\lambda)\) is given by \begin{center} \begin{tikzcd} \cdots \arrow[bend left=60]{r}{-20}