\classheader{2012-10-08} Today we'll talk about simple and semisimple modules. \subsection*{Simple modules} \begin{definition} A module is simple if it is non-trivial and has no submodules except 0 and itself. \end{definition} \begin{example} If our ring is a field $k$, then a $k$-module $M$ is simple if and only if $M$ is cyclic, which is the case if and only if $\dim_k(M)=1$. \end{example} Note that an $A$-module is simple if and only if $M=Am$ for any non-zero $m\in M$. \begin{definition} A (left) ideal $J\subsetneq A$ is called maximal if, for any ideal $I\supseteq J$, we have $A=I$. \end{definition} \begin{lemma} $M$ is a simple $A$-module if and only if $M\cong A/J$ for some maximal (left) ideal $J$ of $A$. \end{lemma} \begin{proof} There is a bijection between left submodules $I/J\subseteq A/J$ and ideals $J\subseteq I\subseteq A$: \begin{center} \begin{tikzpicture}[scale=1.5,>=angle 90] \node (a) at (0,2) {$A$}; \node (b) at (1,2) {$A/J$}; \node (c) at (0,1) {$I$}; \node (d) at (1,1) {$I/J$}; \node (e) at (0,0) {$J$}; \node (f) at (1,0) {$0$}; \draw[->] (a) to (b); \draw[->] (c) to (d); \draw[->] (e) to (f); \node[rotate=90] at (0,1.5) {$\subseteq$}; \node[rotate=90] at (0,0.5) {$\subseteq$}; \node[rotate=90] at (1,1.5) {$\subseteq$}; \node[rotate=90] at (1,0.5) {$\subseteq$}; \end{tikzpicture} \end{center} \vspace{-0.3in} %\[\xymatrix{A\ar@{}[d]|{\cup} \ar[r] & A/J \ar@{}[d]|{\cup} \\ I\ar@{}[d]|{\cup} \ar@{-->}[r] & I/J \ar@{}[d]|{\cup}\\ J \ar[r] & 0}\]\qedhere \end{proof} \begin{proposition} Any (left) ideal $I\subsetneq A$ is contained in a maximal ideal. \end{proposition} \begin{proof} Zorn's lemma. \end{proof} \begin{corollary} Any cyclic module has a simple quotient. \end{corollary} \begin{warning} If $N$ is a submodule in $M$, it isn't necessarily true that there exists a maximal submodule $N'$ of $M$ with $N\subseteq N'$. \end{warning} \begin{exercise} Show that $(\Q/\Z,+)$ has no maximal subgroups. \end{exercise} \begin{theorem}[Schur's lemma] If $M$, $N$ are simple $A$-modules, then any morphism $f:M\to N$ is either 0 or is an isomorphism. \end{theorem} \begin{proof} Assume that $f\neq 0$. Then $\im(f)\neq 0$ is a submodule in $N$, hence $\im(f)=N$, i.e. $f$ is surjective. Because $\ker(f)\neq M$, we must have $\ker(f)=0$, hence $f$ is injective. \end{proof} \begin{corollary} For a simple module $M$, the ring $\End_A(M)$ is a division ring. \end{corollary} \begin{proof} Every $f\neq 0$ in $\End_A(M)$ is an isomorphism. \end{proof} \begin{corollary} If $A$ is commutative, then an ideal $I\subset A$ is maximal if and only if $A/I$ is a field. \end{corollary} \begin{proof} If $I$ is maximal, then $A/I$ is a simple $A$-module, so that $\End_A(A/I)\cong A/I$ is a division ring, and hence a field. If $A/I$ is a field, then $A/I$ is generated by every non-zero element, hence $A/I$ is simple, hence $I$ is maximal (by our lemma). \end{proof} \begin{corollary} If $A$ is commutative, and $n\neq m$, then $A^n\not\cong A^m$ as $A$-modules. \end{corollary} \begin{proof} Pick a maximal ideal $I\subset A$. If $M$ is an $A$-module, we can consider $M/IM$ as an $A/I$-module. In particular we get \[(A/I)^n\cong A^n/I\cdot A^n\cong A^m/I\cdot A^m\cong (A/I)^m\] but these are vector spaces over $A/I$, so we must have $m=n$. \end{proof} This is false for non-commutative rings in general. A good counterexample can be obtained by taking an infinite dimensional vector space $V$ over a field $k$, and letting $A=\End_k(V)$. Then $A^n\cong A$ for all $n\geq 1$. \subsection*{Semisimple modules} \begin{definition}[and Proposition] For $M$ an $A$-module, the following are equivalent: \begin{enumerate} \item $M\cong \oplus \text{ simple modules}$ \item $M=\sum \text{simple submodules}$ \item $M$ is completely reducible, i.e. for any submodule $N\subset M$, there is an $N'\subset M$ such that $N\oplus N'=M$. \end{enumerate} \end{definition} \begin{proof} Zorn's lemma. \end{proof} If (1)-(3) hold then $M$ is called semisimple. \begin{examples} If $A=k$ is a field, then any module is semisimple. If $A$ is a division ring, then any module is semisimple. If $A=\M_n(D)$ where $D$ is a division ring, then any $A$-module is semisimple. \end{examples} \begin{corollary} Any direct sum, quotient, or submodule of a semisimple module is semisimple. \end{corollary} \begin{proof} The case of submodules follows from (3). The case of quotients follows from (2). The case of direct sums follows from (1). \end{proof} \begin{notation} If $M=\bigoplus_i M_i$, where the $M_i$'s are simple, then we set for a simple $L$ \[[M:L]=\#\{i\mid M_i\cong L\}.\] \end{notation} It's not necessarily clear this is well-defined; an alternative is to write $M$ as \[M=\bigoplus_{L\in S_A}L^{[M:L]}\] where $S_A$ is the set of isomorphism classes of simple $A$-modules. For any $A$-modules $M$ and $N$, the ring $\End_A(N)$ acts (on the right) by composition on $\Hom_A(M,N)$. Let $L$ be a simple module. Then $D_L:=\End_A(L)$ is a division ring by Schur's lemma, so that $\Hom_A(M,L)$ is a $D_L$-module. \begin{proposition}[Multiplicity formula] For a semisimple module $M$, \[[M:L]=\operatorname{rank}_{D_L}(\Hom(M,L)).\] \end{proposition} \begin{proof} Assume $M=\bigoplus M_i$, where $M_i$ are simple. Then \[\Hom_A(M,L)=\Hom_A\left(\bigoplus M_i,L\right)=\bigoplus \Hom_A(M_i,L)\cong\bigoplus_{\{i\mid M_i\cong L\}}\End_A(L)=D_L^{[M:L]}.\qedhere\] \end{proof} Let $M=L_1^{r_1}\oplus\cdots\oplus L_n^{r_n}$ where the $L_i$ are simple and $L_i\not\cong L_j$. Then \begin{align*} \End_A(M)&=\Hom(M,L_1^{r_1})\oplus\cdots\oplus \Hom(M,L_n^{r_n})\\ &=\Hom(M,L_1)^{r_1}\oplus\cdots\oplus\Hom(M,L_n)^{r_n}\\ &= \M_{r_1}(D_{L_1})\oplus \cdots\oplus \M_{r_n}(D_{L_n}). \end{align*}