- Commit
- 35b312c77ebe8541c33d26a2e2ca27ec06d96ecb
- Parent
- 68b7ca86f5a81411b6e73e698d79dbea106e2950
- Author
- Pablo <pablo-escobar@riseup.net>
- Date
Incorporated much of the discussion on semisimplicity and complete reducibility to the chapter on sl2
Source code for my notes on representations of semisimple Lie algebras and Olivier Mathieu's classification of simple weight modules
Incorporated much of the discussion on semisimplicity and complete reducibility to the chapter on sl2
1 file changed, 240 insertions, 50 deletions
Status | File Name | N° Changes | Insertions | Deletions |
Modified | sections/complete-reducibility.tex | 290 | 240 | 50 |
diff --git a/sections/complete-reducibility.tex b/sections/complete-reducibility.tex @@ -3,67 +3,257 @@ % TODO: Remove this? \epigraph{Nobody has ever bet enough on a winning horse.}{Some gambler} -% TODOOO: Point out we are now working with finite-dimensional Lie algebras -% over an algebraicly closed field of characteristic zero - -% TODO: Update the 40 pages thing when we're done -% TODO: Have we seen the fact representations are useful? -Having hopefully established in the previous chapter that Lie algebras are -indeed useful, we are now faced with the Herculean task of trying to -understand them. We have seen that representations are a remarkably effective -way to derive information about groups -- and therefore algebras -- but the -question remains: how to we go about classifying the representations of a given -Lie algebra? This is a question that have sparked an entire field of research, -and we cannot hope to provide a comprehensive answer the 40 pages we have left. -Nevertheless, we can work on particular cases. - -Like any sane mathematician would do, we begin by studying a simpler case, -which is that of \emph{semisimple} Lie algebras algebras. The first question we -have is thus: why are semisimple algebras simpler -- or perhaps -\emph{semisimpler} -- to understand than any old Lie algebra? Well, the special -thing about semisimple algebras is that the relationship between their -indecomposable representations and their irreducible representations is much -clearer -- at least in finite dimension. Namely\dots +% TODO: Update the 40 pages thing when we're done TODO: Have we seen the fact +% representations are useful? +Having hopefully established in the previous chapter that Lie algebras and +their representations are indeed useful, we are now faced with the Herculean +task of trying to understand them. We have seen that representations are a +remarkably effective way to derive information about groups -- and therefore +algebras -- but the question remains: how to we go about classifying the +representations of a given Lie algebra? This is a question that have sparked an +entire field of research, and we cannot hope to provide a comprehensive answer +the 40 pages we have left. Nevertheless, we can work on particular cases. + +For instance, one can redily check that a representation \(V\) of the +\(n\)-dimensional Abelian Lie algebra \(K^n\) is nothing more than a choice of +\(n\) commuting operators \(V \to V\) -- correspoding to the action of the +canonical basis elements \(e_1, \ldots, e_n \in K^n\). In other words, +classifying the representations of Abelian algebras is a trivial affair. +Instead, we focus on the the finite-dimensional representations of a +finite-dimensional Lie algebra \(\mathfrak{g}\) over an algebraicly closed +field of characteristic \(0\). But why are the representations semisimple +algebras simpler -- or perhaps \emph{semisimpler} -- to understand than those +of any old Lie algebra? + +We will come back to this question in a moment, but for now we simply note +that, when solving a classification problem, it is often profitable to break +down our structure is smaller peaces. This leads us to the following +definitions. + +\begin{definition} + A representation of \(\mathfrak{g}\) is called \emph{indecomposable} if it is + not isomorphic to the direct sum of two non-zero representations. +\end{definition} + +\begin{definition} + A representation of \(\mathfrak{g}\) is called \emph{irreducible} if it has + no non-zero subrepresentations. +\end{definition} + +\begin{example} + The trivial representation \(K\) is an example of an irreducible + representations. In fact, every \(1\)-dimensional representation \(V\) of a + Lie algebra \(\mathfrak{g}\) is irreducible: \(V\) has no non-zero proper + subspaces, let alone \(\mathfrak{g}\)-invariant subspaces. +\end{example} + +The general strategy for classifying finite-dimensional representations of an +algebra is to classify the indecomposable representations. This is because\dots + +\begin{theorem}[Krull-Schmidt]\label{thm:krull-schmidt} + Every finite-dimensional representation of a Lie algebra can be uniquely -- + up the isomorphisms and reordering of the summands -- decomposed into a + direct sum of indecomposable representations. +\end{theorem} + +Hence finding the indecomposable representations suffices to find \emph{all} +finite-dimensional representations: they are the direct sum of indecomposable +representations. The existence of the decomposition should be clear from the +definitions. Indeed, if \(V\) is representation of \(\mathfrak{g}\) a simple +argument via induction in \(\dim V\) suffices to prove the existence: if \(V\) +is indecomposable then there is nothing to prove, and if \(V\) is not +indecomposable then \(V = W \oplus U\) for some \(W, U \subsetneq V\) non-zero +subrepresentations, so that their dimensions are both strictly smaller than +\(\dim V\) and the existence follows from the induction hypothesis. For a proof +of uniqueness please refer to \cite{etingof}. + +Finding the indecomposable representations of an arbitrary Lie algebra, +however, turns out to be a bit of a circular problem: the indecomposable +representations are the ones that cannot be decomposed, which is to say, those +that are \emph{not} decomposable. Ideally, we would like to find some other +condition, equivalent to indecomposability, but which is easier to work with. +It is clear from the definitions that every irreducible representation is +indecomposable, but there is no reason to believe the converse is true. Indeed, +this is not always the case. For instance\dots + +\begin{example}\label{ex:indecomposable-not-irr} + The space \(V = K^2\) endowed with the homorphism of Lie algebras + \begin{align*} + \rho : K[x] & \to \mathfrak{gl}(V) \\ + x & \mapsto \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix} + \end{align*} + is a representation of the Lie algebra \(K[x]\). Notice \(V\) has a single + non-zero proper subrepresentation, which is spanned by the vector \((1, 0)\). + This is because if \((a + b, b) = \rho(x) \ (a, b) = \lambda (a, b)\) for + some \(\lambda \in \CC\) then \(\lambda = 1\) and \(b = 0\). Hence \(V\) is + indecomposable -- it cannot be broken into a direct sum of \(1\)-dimensional + subrepresentations -- but it is evidently not irreducible. +\end{example} + +This counterexample poses an interesting question: are there conditions one can +impose on an algebra \(\mathfrak{g}\) under which every indecomposable +representation of \(\mathfrak{g}\) is irreducible? This is what is known in +representation theory as \emph{complete reducibility}. + +\begin{definition} + A \(\mathfrak{g}\)-module \(V\) is called \emph{completely reducible} if it + the direct sum of irreducible representations. +\end{definition} + +In case the relationship between complete reducibility and the irreducibility +of indecomposable representations is unclear, the following results should +clear things up. \begin{proposition}\label{thm:complete-reducibility-equiv} - Given a finite-dimensional Lie algebra \(\mathfrak{g}\) over \(K\), the - following conditions are equivalent. + The following conditions are equivalent. \begin{enumerate} - \item \(\mathfrak{g}\) is semisimple. - - \item Given a finite-dimensional representation \(V\) of \(\mathfrak{g}\) - and a subrepresentation \(W \subset V\), \(W\) has a - \(\mathfrak{g}\)-invariant complement in \(V\). + \item Every subrepresentation of a finite-dimensional representation of + \(\mathfrak{g}\) has a \(\mathfrak{g}\)-invariant complement -- i.e. + given \(W \subset V\) there is a subrepresentation \(U \subset V\) such + that \(V = W \oplus U\). \item Every exact sequence of finite-dimensional representations of \(\mathfrak{g}\) splits. - \item Every finite-dimensional indecomposable representation of + \item Every indecomposable finite-dimensional representation of \(\mathfrak{g}\) is irreducible. - \item Every finite-dimensional representation of \(\mathfrak{g}\) can be - uniquely decomposed as a direct sum of irreducible representations. + \item Every finite-dimensional representation of \(\mathfrak{g}\) is + completely reducible. \end{enumerate} \end{proposition} -Condition \textbf{(ii)} is known as \emph{complete reducibility}. The -equivalence between conditions \textbf{(ii)} to \textbf{(iv)} follows at once -from simple arguments. Furthermore, the equivalence between \textbf{(ii)} and -\textbf{(v)} is a direct consequence of the Krull-Schmidt theorem. On the other -hand, the equivalence between \textbf{(i)} and the other items is more subtle. -We are particularly interested in the proof that \textbf{(i)} implies -\textbf{(ii)}. In other words, we are interested in the fact that every -finite-dimensional representation of a semisimple Lie algebra is -\emph{completely reducible}. - -This is because if every finite-dimensional representation of \(\mathfrak{g}\) -is completely reducible, the equivalence between \textbf{(ii)} and \textbf{(v)} -implies a classification of the finite-dimensional irreducible representations -of \(\mathfrak{g}\) leads to a classification of \emph{all} finite-dimensional -representation of \(\mathfrak{g}\) -- it suffices to take direct sums of the -already classified irreducible modules. This leads us to the third restriction -we will impose: for now, we will focus our attention exclusively on -finite-dimensional representations. +\begin{proof} + We begin by \(\textbf{(i)} \Rightarrow \textbf{(ii)}\). Let + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + W \arrow{r}{f} & + V \arrow{r}{g} & + U \arrow{r} & + 0 + \end{tikzcd} + \end{center} + be an exact sequence of representations of \(\mathfrak{g}\). We can suppose + without any loss of generality that \(W \subset V\) is a subrepresentation + and \(f\) is the usual inclusion, for if this is not the case there is an + isomorphism of sequences + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + W \arrow{r}{f} \arrow[swap]{d}{f} & + V \arrow{r}{g} \arrow[Rightarrow, no head]{d} & + U \arrow{r} \arrow[Rightarrow, no head]{d} & + 0 \\ + 0 \arrow{r} & + f(W) \arrow{r} & + V \arrow[swap]{r}{g} & + U \arrow{r} & + 0 + \end{tikzcd} + \end{center} + + It then follows from \textbf{(i)} that there exists a subrepresentation \(U' + \subset V\) such that \(V = W \oplus U'\). Finally, the projection \(\pi : V + \to W\) is an intertwiner satisfying + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + W \arrow{r}{f} & + V \arrow{r}{g} \arrow[bend left=30]{l}{\pi} & + U \arrow{r} & + 0 + \end{tikzcd} + \end{center} + + Next is \(\textbf{(ii)} \Rightarrow \textbf{(iii)}\). If \(V\) is an + indecomposable \(\mathfrak{g}\)-module and \(W \subset V\) is a + subrepresentation, we have an exact sequence + \begin{center} + \begin{tikzcd} + 0 \arrow{r} & + W \arrow{r}{i} & + V \arrow{r}{\pi} & + \mfrac{V}{W} \arrow{r} & + 0 + \end{tikzcd} + \end{center} + of representations of \(\mathfrak{g}\). + + Since our sequence splits, we must have \(V \cong W \oplus \mfrac{V}{W}\). + But \(V\) is indecomposable, so that either \(W = V\) or \(W = 0\). Since + this holds for all \(W \subset V\), \(V\) is irreducible. For + \(\textbf{(iii)} \Rightarrow \textbf{(iv)}\) it suffices to apply + theorem~\ref{thm:krull-schmidt}. + + Finally, for \(\textbf{(iv)} \Rightarrow \textbf{(i)}\), if we assume + \(\textbf{(iii)}\) and let \(V\) be a representation of \(\mathfrak{g}\) with + decomposition into irreducible subrepresentations + \[ + V = \bigoplus_i V_i + \] + and \(W \subset V\) is a subrepresentation. Take some maximal set of indexes + \(\{i_1, \ldots, i_n\}\) so that \(\left( \bigoplus_k V_{i_k} + \right) \cap W = 0\) and let \(U = \bigoplus_k V_{i_k}\). We want to + establish \(V = W \oplus U\). + + Suppose without any loss in generality that \(i_k = k\) for all \(k\) and + let \(j > n\). By the maximality of our set of indexes, there is some + non-zero \(w \in (V_j \oplus U) \cap W\). Say \(w = v_j + v_1 + \cdots + + v_n\) with each \(v_i \in V_i\). Then \(v_j = w - v_1 - \cdots - v_n \in V_j + \cap (W \oplus U)\) is non-zero. Indeed, if this is not the case we find \(0 + \ne w = v_1 + \cdots + v_n \in \left( \bigoplus_{i = 1}^n V_i \right) \cap + W\), a contradiction. This implies \(V_j \cap (W \oplus U)\) is a non-zero + subrepresentation of \(V_j\). Since \(V_j\) is irreducible, \(V_j = V_j \cap + (W \oplus U)\) and therefore \(V_j \subset W \oplus U\). Given the arbitrary + choice of \(j\), it then follows \(V = W \oplus U\). +\end{proof} + +The advantage of working with irreducible representations as opposed to +indecomposable ones is that they are generally much easier to find. The +relationship between irreducible representations is also well understood. This +is because of the following result, known as \emph{Schur's lemma}. + +\begin{lemma}[Schur] + Let \(V\) and \(W\) be irreducible representations of \(\mathfrak{g}\) and + \(T : V \to W\) be an intertwiner. Then \(T\) is either \(0\) or an + isomorphism. Furtheremore, if \(V = W\) then \(T\) is a scalar operator. +\end{lemma} + +\begin{proof} + For the first statement, it suffices to notice that \(\ker T\) and + \(\operatorname{im} T\) are both subrepresentations. In particular, either + \(\ker T = 0\) and \(\operatorname{im} T = W\) or \(\ker T = V\) and + \(\operatorname{im} T = 0\). Now suppose \(V = W\). Let \(\lambda \in K\) be + an eigenvalue of \(T\) -- which exists because \(K\) is algebraicly closed -- + and \(V_\lambda\) be its corresponding eigenspace. Given \(v \in V_\lambda\), + \(T X v = X T v = \lambda \cdot X v\). In other words, \(V_\lambda\) is a + subrepresentation. It then follows \(V_\lambda = V\), given that \(V_\lambda + \ne 0\). +\end{proof} + +We are now ready to answer our first question: the special thing about +semisimple algebras is that the relationship between their indecomposable +representations and their irreducible representations is much clearer. +Namely\dots + +\begin{proposition}\label{thm:complete-reducibility-equiv} + Given a finite-dimensional Lie algebra \(\mathfrak{g}\) over \(K\), + \(\mathfrak{g}\) is semisimple if, and only if every finite-dimensional + representation of \(\mathfrak{g}\) is completely reducible. +\end{proposition} + +The proof of the fact that a finite-dimensional Lie algebra \(\mathfrak{g}\) +whose finite-dimensional representations are completely reducible is semisimple +is actually pretty simple. Namely, it suffices to note that the adjoint +representation \(\mathfrak{g}\) is the direct sum of irreducible +subrepresentations, which are all simple ideals of \(\mathfrak{g}\) -- so +\(\mathfrak{g}\) is the direct sum of simple Lie algebras. The proof of the +converse is more nuanced, and this will be our next milestone. Before proceding +to the proof of complete reducibility, however, we would like to review some +basic tools which will come in handy later on, known as\dots + +\section{Invariant Bilinear Forms} Another interesting characterization of semisimple Lie algebras, which will come in handy later on, is the following. @@ -92,7 +282,7 @@ come in handy later on, is the following. We refer the reader for \cite[ch. 5]{humphreys} for a proof of this last result. Without further ado, we may proceed to a proof of\dots -\section{Complete Reducibility} +\section{Proof of Complete Reducibility} Historically, complete reducibility was first proved by Herman Weyl for \(K = \mathbb{C}\), using his knowledge of smooth representations of compact Lie