\classheader{2012-10-10} A ring $A$ is said to be semisimple when any $A$-module is semisimple. \begin{theorem}[Wedderburn Theorem] For a ring $A$, the following conditions are equivalent: \begin{enumerate} \item $A$ is a finite direct sum \[A=\M_{r_1}(D_1)\oplus\cdots\oplus \M_{r_n}(D_n)\] where the $D_i$ are division rings. \item $A$ is semisimple. \item The rank 1 free $A$-module is semisimple. \item Any left ideal in $A$ has the form $Ae$ where $e^2=e\in A$. \item $A=Ae_1+\cdots+Ae_n$ where $e_i^2=e_i$ and $e_ie_j=0$ for $i\neq j$, and each $Ae_i$ is a simple $A$-module. \end{enumerate} \end{theorem} \begin{proof}[(1 $\Rightarrow$ 2)] The fact that any module over $\M_r(D)$ is semisimple is something that is from your homework. Now note that if $A=A_1\oplus\cdots\oplus A_n$, then an $A$-module is equivalent to an $n$-tuple $\{M_i\}$ where $M_i$ is an $A_i$-module, so that because each $\M_{r_i}(D_i)$ is semisimple, so is $\M_{r_1}(D_1)\oplus\cdots\oplus \M_{r_n}(D_n)$. \end{proof} \begin{proof}[(2 $\Rightarrow$ 3)] Clear. \end{proof} \begin{proof}[(3 $\Rightarrow$ 4)] $A$ is completely reducible as an $A$-module, so for any left ideal $I\subset A$, there exists a left ideal $I'\subset A$ such that $A=I\oplus I'$. Thus, there are $e\in I$ and $e'\in I'$ such that $1=e+e'$. Now note that for any $x\in I$, \[\underbrace{x}_{\in I} = x\cdot 1=\underbrace{\;x\cdot e\;}_{\in I}+\underbrace{\;x\cdot e'\;}_{\in I'}\] which implies that $x\cdot e'$ and $x=xe$. Thus, $I$ is of the form $Ae$, and $e=e^2$. \end{proof} \begin{proof}[(4 $\Rightarrow$ 5)] Since any left ideal $I\subset A$ has the form $I=Ae$ where $e^2=e$, we have for any ideal $Ae$ that $A=Ae\oplus A(1-e)$. Thus, $A$ is completely reducible. This implies that $A=\bigoplus L_i$, a possibly infinite direct sum of simple $A$-modules. But then $1=e_1+\cdots+e_n$, hence $x=x\cdot 1= xe_1+\cdots+xe_n$ where $xe_i\in L_i$. Thus $A=L_1\oplus\cdots\oplus L_n$. \end{proof} \begin{proof}[(5 $\Rightarrow$ 1)] Reexpressing the statement of 5, we have that \[A=\bigoplus_{j=1}^m L_i^{r_i},\qquad L_i\not\cong L_j\text{ for }i\neq j.\] Therefore \[A^{\text{op}}=\End_A(A)=\bigoplus_{j=1}^m\End(L_j^{r_j})=\bigoplus_{j=1}^m\M_{r_j}(\End_A(L_i))\] and by Schur's lemma, $D_i=\End_A(L_i)$ is a division ring. But this gives a decomposition of $A^{\text{op}}$, not $A$. To fix this, note that \[A=(A^{\text{op}})^{\text{op}}=\bigoplus_{j=1}^m \M_{r_j}(D_j)^{\text{op}}\] that $\M_r(D)^{\text{op}}=\M_r(D^{\text{op}})$, and that $D^{\text{op}}$ is a division ring when $D$ is. \end{proof} \begin{corollary} A commutative ring $A$ is semisimple if and only if it is a finite direct sum of fields. \end{corollary} \begin{proof} There is no way to get a commutative ring if any of the $r_i>1$, nor if any of the $D_j$ are non-commutative division algebras. \end{proof} From this point forward, let $A$ be a $k$-algebra where $k$ is a field. \begin{definition} An element $a\in A$ is called algebraic if there is some monic $p\in k[t]$ such that $p(a)=0$. \end{definition} \begin{examples} $\text{}$ \begin{itemize} \item Any idempotent or nilpotent element is algebraic. \item If $\dim_k(A)<\infty$ then any $a\in A$ is algebraic, because $1,a,a^2,\ldots$ cannot be linearly independent over $k$. Thus, there exist some $\lambda_i\in k$ such that $\sum \lambda_i a^{n_i}=0$. \end{itemize} \end{examples} Let $a\in A$ be an algebraic element. Consider the $k$-algebra homomorphism $j:k[t]\to A$ defined by $j(f)=f(a)$. Because $k[t]$ is a PID, we have that $\ker(j)=(p_a)$ is a principal ideal. The monic polynomial $p_a$ is called the minimal polynomial for $a$. Note that we get an induced homomorphism $j:k[t]/(p_a)\hookrightarrow A$. \begin{definition} For any $a\in A$, we define \[\spec(a)=\{\lambda\in k\mid \lambda-a\text{ is not invertible}\}.\] \end{definition} \begin{examples} $\text{}$ \begin{itemize} \item If $A=k\{X\}$, then $\spec(a)=\{\text{the values of }a\}$. \item If $k$ is algebraically closed and $A=\M_n(k)$, then for an $a\in A$, $\spec(a)=\{\text{eigenvalues of }a\}$. \end{itemize} \end{examples} \begin{lemma} Let $a\in A$ be an algebraic element with minimal polynomial $p_a\in k[t]$. Then \[\lambda-a\text{ is not invertible}\iff \lambda-a\text{ is a zero divisor}\iff p_a(\lambda)=0.\] Thus, $\spec(a)=\{\text{roots of }p_a\}$. \end{lemma} \begin{proof} First, we make a general remark: for any $\lambda\in k$, we have \[p(t)-p(\lambda)=q(t)(\lambda-t),\] and because $\deg(q)<\deg(p)$, we must have $q\notin (p)$, hence $q(a)\neq 0$. We have an injective homomorphism $j:k[t]/(p)\hookrightarrow A$ where we send $f$ to $f(a)$. If $p(\lambda)=0$, then \[0=p(a)=q(a)(\lambda-a)\] implies that $\lambda-a$ is a zero-divisor. Clearly, if $\lambda-a$ is a zero-divisor, it is not invertible. Now we want to show that $\lambda-a$ is not invertible $\implies$ $p(\lambda)=0$. Assume for the sake of contradiction that $p(\lambda)\neq 0$; then \[\underbrace{p(a)}_{=0} - \underbrace{p(\lambda)}_{\neq 0}=q(a)(\lambda-a)\] demonstrates that $\lambda-a$ is invertible. \end{proof}