- Commit
- fcce6df81e52f7f57c9aece3394770f265e18eb9
- Parent
- e8eaf8a1546120705c9455f9a1de77a3db93530c
- Author
- Pablo <pablo-escobar@riseup.net>
- Date
Comecei a trabalhar na prova de redutibilidade completa
Source code for my notes on representations of semisimple Lie algebras and Olivier Mathieu's classification of simple weight modules
Comecei a trabalhar na prova de redutibilidade completa
1 file changed, 373 insertions, 27 deletions
Status | File Name | N° Changes | Insertions | Deletions |
Modified | sections/semisimple-algebras.tex | 400 | 373 | 27 |
diff --git a/sections/semisimple-algebras.tex b/sections/semisimple-algebras.tex @@ -1,7 +1,6 @@ \chapter{Semisimple Lie Algebras \& their Representations}\label{ch:lie-algebras} -\epigraph{Nobody has ever bet enough on a winning horse.}{\textit{Some -gambler}} +\epigraph{Nobody has ever bet enough on a winning horse.}{Some gambler} % TODO: Update the 40 pages thing when we're done Having hopefully stablished in the previous chapter that Lia algebras are @@ -21,10 +20,11 @@ themselves. First of all, we will work exclusively with finite-dimensional Lie algebras over an algebraicly closed field \(K\) of characteristic \(0\). This is a restriction we will cary throught these notes. Moreover, as indicated by the title of this chapter, we will initially focus on the so called -\emph{semisimple} Lie algebras algebras\footnote{We will later relax this -restriction a bit in the next chapter.}. There are multiple equivalent ways to -define what a semisimple Lie algebra is. Perhaps the most common definition -is\dots +\emph{semisimple} Lie algebras algebras -- we will later relax this restriction +a bit in the next chapter when we dive into \emph{reductive} Lie algebras. + +There are multiple equivalent ways to define what a semisimple Lie algebra is. +Perhaps the most common definition is\dots \begin{definition}\label{thm:sesimple-algebra} A Lie algebra \(\mathfrak{g}\) over \(K\) is called \emph{semisimple} if it @@ -47,9 +47,12 @@ is\dots A popular alternative to definition~\ref{thm:sesimple-algebra} is\dots \begin{definition}\label{def:semisimple-is-direct-sum} - A Lie algebra \(\mathfrak{s}\) over \(K\) is called \emph{simple} if its only - ideals are \(0\) and \(\mathfrak{s}\). A Lie algebra \(\mathfrak{g}\) is - called \emph{semisimple} if it is the direct sum of simple Lie algebras. + A non-Abelian Lie algebra \(\mathfrak{s}\) over \(K\) is called \emph{simple} + if its only ideals are \(0\) and \(\mathfrak{s}\). A Lie algebra + \(\mathfrak{g}\) is called \emph{semisimple} if it is the direct sum of + simple Lie algebras. Furthermore, a Lie algebra \(\mathfrak{g}\) is called + reductive if \(\mathfrak{g}\) is the direct sum of a reductive Lie algebra + and an Abelian Lie algebra. \end{definition} % TODO: Give a small proof? (At least for n = 2) @@ -59,6 +62,8 @@ A popular alternative to definition~\ref{thm:sesimple-algebra} is\dots forms and the semisimplicity of classical Lie algebras. \end{example} +% TODO: Add gl_n(K) as an example of a reductive algebra + I suppose this last definition explains the nomenclature, but the reason why semisimple Lie algebras are interesting at all is still unclear. In particual, why are they simpler -- or perhaps \emph{semisimpler} -- to understnad than any @@ -122,26 +127,368 @@ handy later on, is\dots \section{Some Homological Algebra} \begin{theorem} + There is a sequence of bifunctors \(\operatorname{Ext}^i : + \mathfrak{g}\text{-}\mathbf{Mod} \times \mathfrak{g}\text{-}\mathbf{Mod} \to + K\text{-}\mathbf{Vect}\), \(i \ge 0\) such that every exact + sequence of \(\mathfrak{g}\)-modules + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & W \arrow{r}{i} & V \arrow{r}{\pi} & U \arrow{r} & 0 + \end{tikzcd} + \end{center} + induces long exact sequences + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + \operatorname{Hom}_{\mathfrak{g}}(S, W) \arrow{r}{i \circ -} & + \operatorname{Hom}_{\mathfrak{g}}(S, V) \arrow{r}{\pi \circ -} & + \operatorname{Hom}_{\mathfrak{g}}(S, U) \arrow{r} & + \hphantom{0} \\ + \hphantom{0} \arrow{r} & + \operatorname{Ext}^1(S, W) \arrow{r} & + \operatorname{Ext}^1(S, V) \arrow{r} & + \operatorname{Ext}^1(S, U) \arrow{r} & + \hphantom{0} \\ + \hphantom{0} \arrow{r} & + \operatorname{Ext}^2(S, W) \arrow{r} & + \operatorname{Ext}^2(S, V) \arrow{r} & + \operatorname{Ext}^2(S, U) \arrow{r} & + \cdots + \end{tikzcd} + \end{center} + and + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + \operatorname{Hom}_{\mathfrak{g}}(U, S) \arrow{r}{- \circ \pi} & + \operatorname{Hom}_{\mathfrak{g}}(V, S) \arrow{r}{- \circ i} & + \operatorname{Hom}_{\mathfrak{g}}(W, S) \arrow{r} & + \hphantom{0} \\ + \hphantom{0} \arrow{r} & + \operatorname{Ext}^1(U, S) \arrow{r} & + \operatorname{Ext}^1(V, S) \arrow{r} & + \operatorname{Ext}^1(W, S) \arrow{r} & + \hphantom{0} \\ + \hphantom{0} \arrow{r} & + \operatorname{Ext}^2(U, S) \arrow{r} & + \operatorname{Ext}^2(V, S) \arrow{r} & + \operatorname{Ext}^2(W, S) \arrow{r} & + \cdots + \end{tikzcd} + \end{center} +\end{theorem} + +% TODO: Make the correspondance more precise? +\begin{theorem}\label{thm:ext-1-classify-short-seqs} + Given \(\mathfrak{g}\)-modules \(W\) and \(U\), there is a one-to-one + correspondance between elements of \(\operatorname{Ext}^1(W, U)\) and + isomorphism classes of short exact sequences + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & W \arrow{r} & V \arrow{r} & U \arrow{r} & 0 + \end{tikzcd} + \end{center} + + In particular, \(\operatorname{Ext}^1(W, U) = 0\) if, and only if every short + exact sequence of \(\mathfrak{g}\)-modules with \(W\) and \(U\) in the + extremes splits. +\end{theorem} + +We are particular interested in the case where \(S = K\) is the trivial +representation of \(\mathfrak{g}\). + +\begin{definition} + Given a \(\mathfrak{g}\)-module \(V\), we refer to the Abelian group + \(H^i(\mathfrak{g}, V) = \operatorname{Ext}^i(K, V)\) as \emph{the \(i\)-th + Lie algebra cohomology group of \(V\)}. +\end{definition} + +% TODO: Prove this +% TODO: Define invariants beforehand +\begin{corollary} + Every short exact sequence of \(\mathfrak{g}\)-modules + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & W \arrow{r}{i} & V \arrow{r}{\pi} & U \arrow{r} & 0 + \end{tikzcd} + \end{center} + induces a long exact sequence + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + W^{\mathfrak{g}} \arrow{r}{i} & + V^{\mathfrak{g}} \arrow{r}{\pi} & + U^{\mathfrak{g}} \arrow{r} & + H^1(\mathfrak{g}, W) \arrow{r} & + H^1(\mathfrak{g}, V) \arrow{r} & + H^1(\mathfrak{g}, U) \arrow{r} & + \cdots + \end{tikzcd} + \end{center} +\end{corollary} + +\begin{definition}\label{def:casimir-element} + Let \(\{X_i\}_i\) be a basis for \(\mathfrak{g}\) and denote by \(\{X^i\}_i\) + its dual basis -- i.e. the unique basis for \(\mathfrak{g}\) satisfying + \(B(X_i, X^j) = \delta_{i j}\). We call + \[ + C = X_1 X^1 + \cdots + X_n X^n \in \mathcal{U}(\mathfrak{g}) + \] + the \emph{Casimir element of \(\mathcal{U}(\mathfrak{g})\)}. +\end{definition} + +\begin{lemma} + The definition of \(C\) is independant of the choice of basis \(\{X_i\}_i\). +\end{lemma} + +\begin{proof} + Whatever basis \(\{X_i\}_i\) we choose, the image of \(C\) under the + canonical isomorphism \(\mathfrak{g} \otimes \mathfrak{g} \isoto \mathfrak{g} + \otimes \mathfrak{g}^* \isoto \operatorname{End}(\mathfrak{g})\) is the + identity operator -- here the isomorphism \(\mathfrak{g} \otimes \mathfrak{g} + \isoto \mathfrak{g} \otimes \mathfrak{g}^*\) is given by tensoring the + identity \(\mathfrak{g} \to \mathfrak{g}\) with the isomorphism + \(\mathfrak{g} \isoto \mathfrak{g}^*\) induced by the Killing form \(B\). +\end{proof} + +\begin{proposition} + The Casimir element \(C \in \mathcal{U}(\mathfrak{g})\) is central, so that + \(C : V \to V\) is an intertnining operator for any \(\mathfrak{g}\)-module + \(V\). Furthermore, \(C\) acts as a non-zero scalar operator whenever \(V\) + is a non-trivial finite-dimensional irreducible representation of + \(\mathfrak{g}\). +\end{proposition} + +\begin{proof} + To see that \(C\) is central fix a basis \(\{X_i\}_i\) for \(\mathfrak{g}\) + and denote by \(\{X^i\}_i\) its dual basis as in + definition~\ref{def:casimir-element}. Let \(X \in \mathfrak{g}\) and denote + by \(\lambda_{i j}, \mu_{i j} \in K\) the coefficients of \(X_j\) and \(X^j\) + in \([X, X_i]\) and \([X, X^i]\), respectively. + + % TODO: Comment on the invariance of the Killing form beforehand + The invariance of \(B\) implies + \[ + \lambda_{i k} + = B([X, X_i], X^k) + = B(-[X_i, X], X^k) + = B(X_i, -[X, X^k]) + = - \mu_{k i} + \] + + Hence + \[ + \begin{split} + [X, C] + & = \sum_i [X, X_i X^i] \\ + & = \sum_i [X, X_i] X^i + \sum_i X_i [X, X^i] \\ + & = \sum_{i j} \lambda_{i j} X_j X^i + \sum_{i j} \mu_{i j} X_i X^j \\ + & = 0 + \end{split}, + \] + and \(C\) is central. This implies that \(C : V \to V\) is an intertwiner for + all representations \(V\) of \(\mathfrak{g}\): its action commutes with the + action of any other element of \(\mathfrak{g}\). + + % TODO: Prove that the action is not zero when V is non-trivial + In particular, it follows from Schur's lemma that if \(V\) is + finite-dimensional and irreducible then \(C\) acts in \(V\) as a scalar + operator. +\end{proof} + +\begin{proposition}\label{thm:first-cohomology-vanishes} + Let \(V\) be a finite-dimensional representation of \(\mathfrak{g}\). Then + \(H^1(\mathfrak{g}, V) = 0\). +\end{proposition} + +\begin{proof} + We begin by the case where \(V\) is irreducible. Due to + theorem~\ref{thm:ext-1-classify-short-seqs}, it suffices to show that any + exact sequence of the form + \begin{equation}\label{eq:exact-seq-h1-vanishes} + \begin{tikzcd} + 0 \arrow{r} & K \arrow{r} & W \arrow{r}{\pi} & V \arrow{r} & 0 + \end{tikzcd} + \end{equation} + splits. + + If \(V = K\) is the trivial representation then the exactness of + \begin{equation}\label{eq:trivial-extrems-exact-seq} + \begin{tikzcd} + 0 \arrow{r} & K \arrow{r} & W \arrow{r}{\pi} & K \arrow{r} & 0 + \end{tikzcd} + \end{equation} + implies \(W\) is 2-dimensional. Take any non-zero \(w_2 \in W\) outside of + the image of the inclusion \(K \to W\). + + Since \(\dim W = 2\), the irreducible component \(\mathcal{U}(\mathfrak{g}) + \cdot w\) of \(w\) in \(W\) is either \(K w\) or \(W\) itself. But this + component cannot be \(W\), since the image the inclusion \(K \to W\) is a + 1-dimensional representation -- i.e. a proper non-zero subrepresentation. + Hence \(K w\) is invariant under the action of \(\mathfrak{g}\). In + particular, \(X w = 0\) for all \(X \in \mathfrak{g}\). Since \(w\) lies + outside the image of the inclusion \(K \to W\), \(\pi(w) \ne 0\) -- which is + to say, \(w \notin \ker \pi\). This implies the map \(K \to W\) that takes + \(1\) to \(\sfrac{w}{\pi(w)}\) is a spliting of + (\ref{eq:trivial-extrems-exact-seq}). + + Now suppose that \(V\) is non-trivial, so that \(C\) acts on \(V\) as + \(\lambda \operatorname{Id}\) for some \(\lambda \ne 0\). Given an eigenvalue + \(\mu \in K\) of the action of \(C\) in \(W\), denote by \(W^\mu\) its + associated generalized eigespace. We claim \(W^0\) is the image of the + inclusion \(K \to W\). Since \(C\) acts as zero in \(K\), this image is + clearly contained in \(W^0\). On the other hand, if \(w \in W\) is such that + \(C^n w = 0\) then + \[ + \lambda^n \pi(w) + = C^n \pi(w) + = \pi(C^n w) + = 0, + \] + so that \(w \in \ker \pi\) -- because \(\lambda^n \ne 0\). The exactness of + (\ref{eq:exact-seq-h1-vanishes}) then implies the desired conclusion. + + We furthermore claim that the only eigenvalues of \(C\) in \(W\) are \(0\) + and \(\lambda\). Indeed, if \(\mu \ne 0\) is eigenvalue and \(w\) is an + associated eigenvector, then + \[ + \mu \pi(w) = \pi(C w) = C \pi(w) = \lambda \pi(w) + \] + + Since \(w \notin W^0\), \(\pi(w) \ne 0\) and therefore \(\mu = \lambda\). + Hence \(W = W^0 \oplus W^\lambda\). The fact that \(C\) is central implies + \((C - \lambda \operatorname{Id})^n X v = X (C - \lambda \operatorname{Id})^n + v\) for all \(v \in V\), \(X \in \mathfrak{g}\) and \(n > 0\). In particular, + \(W^\lambda\) is stable under the action of \(\mathfrak{g}\) -- i.e. + \(W^\lambda\) is a subrepresentation. Since \(W^0\) is precisely the kernel + of \(\pi\), we have an isomorphism of representations \(W^\lambda \cong + \sfrac{W}{W^0} \isoto V\), which induces a splitting \(W \cong K \oplus V\). + + Finally, we consider the case where \(V\) is not irreducible. Suppose + \(H^1(\mathfrak{g}, W) = 0\) for all \(\mathfrak{g}\)-modules with \(\dim W < + \dim V\) and let \(W \subset V\) be a proper non-zero subrepresentation. Then + the exact sequence + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & W \arrow{r} & V \arrow{r} & \sfrac{V}{W} \arrow{r} & 0 + \end{tikzcd} + \end{center} + induces a long exact sequence of the form + \begin{center} + \begin{tikzcd} + \cdots \arrow{r} & + H^1(\mathfrak{g}, W) \arrow{r} & + H^1(\mathfrak{g}, V) \arrow{r} & + H^1(\mathfrak{g}, \sfrac{V}{W}) \arrow{r} & + \cdots + \end{tikzcd} + \end{center} + + Since \(0 < \dim W, \dim \sfrac{V}{W} < \dim V\) it follows + \(H^1(\mathfrak{g}, W) = H^1(\mathfrak{g}, \sfrac{V}{W}) = 0\). The exactness + of + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + H^1(\mathfrak{g}, V) \arrow{r} & + 0 + \end{tikzcd} + \end{center} + then implies \(H^1(\mathfrak{g}, V) = 0\). Hence by induction in \(\dim V\) + we find \(H^1(\mathfrak{g}, V) = 0\) for all finite-dimensional \(V\). We are + done. +\end{proof} + +\begin{theorem} Every representation of a semisimple Lie algebra is completely reducible. \end{theorem} -%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% -% TODO: Move this to the introduction -%By the same token, most of other aspects of representation -%theory of compact groups must hold in the context of semisimple algebras. For -%instance, we have\dots -% -%\begin{lemma}[Schur] -% Let \(V\) and \(W\) be two irreducible representations of a complex -% semisimple Lie algebra \(\mathfrak{g}\) and \(T : V \to W\) be an -% intertwining operator. Then either \(T = 0\) or \(T\) is an isomorphism. -% Furthermore, if \(V = W\) then \(T\) is scalar multiple of the identity. -%\end{lemma} -% -%\begin{corollary} -% Every irreducible representation of an Abelian Lie group is 1-dimensional. -%\end{corollary} -%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% +\begin{proof} + Let + \begin{equation}\label{eq:generict-exact-sequence} + \begin{tikzcd} + 0 \arrow{r} & W \arrow{r} & V \arrow{r}{\pi} & U \arrow{r} & 0 + \end{tikzcd} + \end{equation} + be a short exact sequence of finite-dimensional representations of + \(\mathfrak{g}\). We want to establish that + (\ref{eq:generict-exact-sequence}) splits. + + We have an exact sequence + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + \operatorname{Hom}(U, W) \arrow{r} & + \operatorname{Hom}(U, V) \arrow{r}{\pi \circ -} & + \operatorname{Hom}(U, U) \arrow{r} & 0 + \end{tikzcd} + \end{center} + of vector spaces. Since all maps involved are intertwiners, this is an exact + sequence of \(\mathfrak{g}\)-modules. Now by applying the functor + \(\operatorname{Hom}_{\mathfrak{g}}(K, -)\) to our sequence, we get long + exact sequence + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + \operatorname{Hom}(U, W)^{\mathfrak{g}} \arrow{r} & + \operatorname{Hom}(U, V)^{\mathfrak{g}} \arrow{r}{\pi \circ -} & + \operatorname{Hom}(U, U)^{\mathfrak{g}} \arrow{r} & + \hphantom{0} + \\ + \hphantom{0} \arrow{r} & + H^1(\mathfrak{g}, \operatorname{Hom}(U, W)) \arrow{r} & + H^1(\mathfrak{g}, \operatorname{Hom}(U, V)) \arrow{r} & + H^1(\mathfrak{g}, \operatorname{Hom}(U, U)) \arrow{r} & + \cdots + \end{tikzcd} + \end{center} + of vector spaces. But \(H^1(\mathfrak{g}, \operatorname{Hom}(U, W))\) + vanishes because of proposition~\ref{thm:first-cohomology-vanishes}. Hence we + have an exact sequence + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + \operatorname{Hom}(U, W)^{\mathfrak{g}} \arrow{r} & + \operatorname{Hom}(U, V)^{\mathfrak{g}} \arrow{r}{\pi \circ -} & + \operatorname{Hom}(U, U)^{\mathfrak{g}} \arrow{r} & + 0 + \end{tikzcd} + \end{center} + + Now notice \(\operatorname{Hom}(U, -)^{\mathfrak{g}} = + \operatorname{Hom}_{\mathfrak{g}}(U, -)\). Indeed, + \[ + \begin{split} + T \in \operatorname{Hom}(U, S)^{\mathfrak{g}} + & \iff X T - T X = 0 \quad \forall X \in \mathfrak{g} \\ + & \iff X T = T X \quad \forall X \in \mathfrak{g} \\ + & \iff T \in \operatorname{Hom}_{\mathfrak{g}}(U, S) + \end{split} + \] + for all \(\mathfrak{g}\)-module \(S\). We thus have a short exact sequence + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + \operatorname{Hom}_{\mathfrak{g}}(U, W) \arrow{r} & + \operatorname{Hom}_{\mathfrak{g}}(U, V) \arrow{r}{\pi \circ -} & + \operatorname{Hom}_{\mathfrak{g}}(U, U) \arrow{r} & + 0 + \end{tikzcd} + \end{center} + + In particular, there is some intertwiner \(T : U \to V\) such that \(\pi + \circ T : U \to U\) is the identity operator. In other words + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + W \arrow{r} & + V \arrow{r}{\pi} & + U \arrow{r} \arrow[bend left]{l}{T} & + 0 + \end{tikzcd} + \end{center} + is a splitting of (\ref{eq:generict-exact-sequence}). +\end{proof} % TODO: Turn this into a proper proof Alternatively, one could prove the same statement in a purely algebraic manner @@ -1418,7 +1765,6 @@ 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. -% TODO: Define reductive Lie algebras beforehand? % TODO: Define the associated Borel subalgebra as soon as possible (we need to % fix an ordering of the roots beforehand) \begin{definition}