\classheader{2012-10-12} Last time, we proved that for an algebraic element $a\in A$ in an algebra, \[\spec(a)=\{\text{roots of the minimal polynomial }p_a\in k[t]\}.\] Let $a^n=0$, so that $a$ is nilpotent. Then $p_a=t^n$, so that $\spec(a)=\{0\}$. This implies that $\lambda-a$ is invertible for any $\lambda\neq 0$. Let's see if we can find an explicit inverse. \begin{align*} (\lambda-a)^{-1}&=[\lambda(1-\lambda^{-1}a)]^{-1}\\ &=\lambda^{-1}(1-\lambda^{-1}a)^{-1}\\ &=\lambda^{-1}(1+(\lambda^{-1}a)+(\lambda^{-1}a)^2+\cdots)\\ &=\sum_{i=0}^\infty\lambda^{-(i+1)}a^i\\ &=\sum_{i=0}^{n-1}\lambda^{-(i+1)}a^i \end{align*} The intermediate steps aren't really allowed, but it gets us the right answer. From now on we consider $k$-algebras over an algebraically closed field $k$. \begin{proposition} Let $A$ be a finite dimensional $k$-algebra. \begin{enumerate} \item $\spec(a)\neq \varnothing$ for any $a\in A$. \item If $A$ is a division algebra, then $A\cong k$. \end{enumerate} \end{proposition} \begin{proof} For part 1, note that $\dim(A)<\infty$ implies that any $a\in A$ is algebraic, so that $\spec(a)=\{\text{roots of }p_a\}$, which is non-empty because $k$ is algebraically closed. For part 2, note that for any $k$-algebra, we have an inclusion $k\hookrightarrow A$ by $\lambda\mapsto \lambda\cdot 1_A$. Suppose that $a\in A\setminus k$. Then for any $\lambda\in k$, we have that $\lambda-a\neq 0$, hence $\lambda-a$ is invertible for all $a\in k$, hence $\spec(a)=\varnothing$; but this contradicts part 1. \end{proof} \begin{proposition}[Schur lemma for algebras] Let $M$ be a finite dimensional (over $k$) simple $A$-module. Then $\End_A(M)=k$, i.e. any endomorphism $f:M\to M$ is of the form $\lambda\cdot\id_M$ for some $\lambda\in k$. \end{proposition} \begin{proof} We know that $\End_A(M)$ is a division algebra. The $A$-action on $M$ is the same as a map $A\to \End_k(M)$, which is a finite dimensional division algebra. Let $A'=\im(A)$, so $\dim(A')<\infty$ and $M$ is a simple $A'$-module, which implies $\End_{A'}(M)=k$, but $\End_A(M)=\End_{A'}(M)$. \end{proof} \begin{corollary} $\text{}$ \begin{enumerate} \item The center $Z(A)$ of $A$ acts by scalars in any finite-dimensional simple $A$-module. \item If $A$ is commutative\footnote{A reader points out that $A$ also needs to be finite dimensional; otherwise a transcendental field extension of $k$ would provide a counterexample.}, then any simple $A$-module has dimension 1 over $k$. \end{enumerate} \end{corollary} \begin{proof} Let $M$ be a finite-dimensional simple $A$-module. Then consider the action map $\operatorname{act}:A\to\End_k(M)$. If $z\in Z(A)$, then $\operatorname{act}(z)\in\End_A(M)=k$ by the Schur lemma for algebras. Part 1 follows. If $A=Z(A)$, then any element of $A$ acts on $M$ by scalars, so any vector subspace $N\subseteq M$ is $A$-stable. Thus, $\dim_k(M)=1$. \end{proof} Let $S_A$ be the isomorphism classes of simple $A$-modules. Let $M$ be a finite direct sum of simple modules, so \[M\cong\bigoplus_{L\in S_A}L^{n(L)}\] For any $A$-module $N$, we have an evaluation map $\operatorname{ev}:\Hom_A(N,M)\otimes_k N\to M$, defined by sending $f\otimes n$ to $f(n)$. Now $M$ is a finite direct sum of simple modules, so we get an evaluation map \[\bigoplus_{L\in S^A}\Hom_A(L,M)\otimes_k L\xrightarrow{\;\;\cong\;\;} M\] by the Schur lemma. \begin{theorem}[Wedderburn Theorem for Algebras] $A$ is a finite-dimensional semi-simple $k$-algebra if and only if \[A\cong \M_{r_1}(k)\oplus\cdots\oplus\M_{r_n}(k).\] \end{theorem} \begin{proof} By Wedderburn's theorem, we have \[A\cong \M_{r_1}(D_1)\oplus\cdots\oplus\M_{r_n}(D_n)\] where the $D_i$ are division rings. The fact that $\dim_k(A)<\infty$ implies $\dim_k(D_i)<\infty$ which implies that $D_i=k$. \end{proof} \begin{corollary} Let $A$ be a finite-dimensional semisimple algebra. Then \begin{enumerate} \item $S_A$ is a finite set and any $M\in S_A$ is finite dimensional over $k$. Moreover, $[A:L]=\dim(L)$ for all $L\in S_A$. \item $\dim(A)=\sum_{L\in S_A}\dim(L)^2$ \end{enumerate} \end{corollary} \begin{proof} Any simple $A$-module is cyclic, so that it is isomorphic to $A/J$ for some $J$. We have $\dim_k(A/J)\leq\dim_k(A)<\infty$, so any simple $A$-module is finite dimensional. Note that $A$ satisfies a universal property, $\Hom_A(A,L)\xrightarrow{\cong}L$ is an isomorphism of $k$-vector spaces. We can decompose $A=\bigoplus_{L\in S_A}L^{n_L}$. We get that, for any simple $L'$, \[\Hom_A(L',A)=\Hom_A\bigg(L',\bigoplus_{L\in S_A} L^{n_L}\bigg)=\bigoplus_{L\in S_A}\Hom(L',L)^{n_L}=\bigoplus_{L\in S_A}\left(\genfrac{}{}{0pt}{}{k\text{ if }L=L'}{0\text{ if }L\neq L'}\right)^{n_L}= k^{n_{L'}}.\] By the same argument we also have \[\Hom_A(A,L')=\Hom_A\bigg(\bigoplus_{L\in S_A} L^{n_L},L'\bigg)=\bigoplus_{L\in S_A}\Hom(L,L')^{n_L}=\bigoplus_{L\in S_A}\left(\genfrac{}{}{0pt}{}{k\text{ if }L=L'}{0\text{ if }L\neq L'}\right)^{n_L}= k^{n_{L'}}.\] Because $\Hom_A(A,L)\cong L$ as $k$-vector spaces, we therefore have that $\dim(L')=\dim(\Hom_A(A,L'))=n_{L'}$, and hence \[\dim(A)=\sum n_L\dim(\Hom_A(L,A))=\sum n_L\dim(L)=\sum \dim(L)^2.\qedhere\] \end{proof}