\classheader{2012-10-01} Today we'll be talking about the Chinese Remainder Theorem and some of its consequences. \newtheorem*{theorem-lecture1-lagrange}{Lagrange Interpolation Formula} \begin{theorem-lecture1-lagrange} Let $k$ be a field. Then for any $c_1,\ldots,c_n\in k$, and any distinct $x_1,\ldots,x_n\in k$, there exists a $p\in k[t]$ such that $p(x_i)=c_i$ for all $i$. \end{theorem-lecture1-lagrange} \begin{proof} The proof comes in three steps. \textit{Step 1:} For each $i=1,\ldots,n$, define \[p_{ij}(t)=\frac{t-x_j}{x_i-x_j}.\] Note that \[p_{ij}(t)=\begin{cases}1 & \text{if } t= x_i,\\ 0 & \text{if }t=x_j.\end{cases}\] \textit{Step 2:} For each $i=1,\ldots,n$, define \[p_i(t)=\prod_{j\neq i}p_{ij}(t).\] Note that \[p_{i}(t)=\begin{cases}1 & \text{if } t= x_i,\\ 0 & \text{if }t=x_j\text{ for any }j\neq i.\end{cases}\] \textit{Step 3:} Define \[p(t)=\sum_{i=1}^n c_ip_i(t).\] It is easy to verify that this $p$ satisfies the claimed property. \end{proof} In this course, rings will always be unital, though they will not necessarily be commutative. Given a ring $A$ and two-sided ideals $I_1,\ldots,I_n\subset A$, there are some important ways we can create new two-sided ideals. The sum of the $I_i$ is defined by \[I_1+\cdots+I_n=\{x_1+\cdots+x_n\mid x_i\in I_i\}.\] The product of the $I_i$ is defined by \[I_1\cdots I_n=\{\sum x_1\cdots x_n\mid x_i\in I_i\}.\] The intersection of the $I_i$ is just their intersection as subsets of $A$. Note that we always have $I_1\cdots I_n\subseteq I_1\cap \cdots \cap I_n$. We can re-express the Lagrange Interpolation Formula as follows: \newtheorem*{theorem-lecture1-lagrange2}{Lagrange Interpolation Formula} \begin{theorem-lecture1-lagrange2} Let $k$ be a field, and let $A=k[t]$. For any $c_1,\ldots,c_n\in k$, and any distinct $x_1,\ldots,x_n\in k$, let $I_i=(t-x_i)$. There exists a $p\in A$ such that $p-c_i\in I_i$ for all $i$. \end{theorem-lecture1-lagrange2} Note that \begin{align*}\text{the }x_i\text{ are distinct}&\iff \text{for any }i\neq j,\,-(t-x_i)+(t-x_j)=x_i-x_j\neq 0\\ &\iff \text{for any }i\neq j,\,I_i+I_j=A. \end{align*} \newtheorem*{theorem-lecture1-crt}{Chinese Remainder Theorem} \begin{theorem-lecture1-crt} Let $A$ be a not-necessarily-commutative ring, and $I_1,\ldots,I_n$ two-sided ideals of $A$ such that $I_i+I_j=A$ for all $i\neq j$. Then \begin{enumerate} \item For any $c_1,\ldots,c_n\in A$, there exists $p\in A$ such that $p-c_i\in I_i$ for all $i$. \item If $A$ is commutative, then $I_1\cdots I_n=I_1\cap \cdots \cap I_n$. \end{enumerate} \end{theorem-lecture1-crt} \begin{proof}[Proof of Part 1] The proof comes in three steps. \textit{Step 1:} For any $i\neq j$, we have that $I_i+I_j=A$. Thus, for any $i\neq j$ there are some $q_{ij}\in I_i$ and $p_{ij}\in I_j$ such that $q_{ij}+p_{ij}=1$. Note that $1-p_{ij}=q_{ij}\in I_i$. \textit{Step 2:} For each $i$, let \[p_i=\prod_{j\neq i}p_{ij}\in\prod_{j\neq i} I_j\subset I_j\text{ for all }j\neq i.\] We claim that $1-p_i\in I_i$. This follows from observing that \[p_i=\prod_{j\neq i}(1-q_{ij})=1+(\text{terms involving the }q_{ij})\] and that for all $j\neq i$, $q_{ij}\in I_i$. \textit{Step 3:} Let \[p=\sum_{i=1}^n c_ip_i.\] Checking that this $p$ works, we see that for any $i$, \[p-c_i=c_i(1-p_i)+\sum_{j\neq i}c_jp_j,\] and $(1-p_i)\in I_i$ and each $p_j\in I_i$ by Step 2, so that $p-c_i\in I_i$ for all $i$. \end{proof} \begin{proof}[Proof of Part 2] We proceed by induction. For the case of $n=2$, note that because $I_1+I_2=A$, there are some $u_1\in I_1$, $u_2\in I_2$ such that $u_1+u_2=1$. For any $a\in I_1\cap I_2$, we then have \[a=au_1+au_2.\] Because $a\in I_2$ and $u_1\in I_1$, we have that $au_1\in I_2I_1$. Because $a\in I_1$ and $u_2\in I_2$, we have that $au_2\in I_1I_2$. Now using the assumption that $A$ is commutative, we have that $I_2I_1=I_1I_2$ and therefore $a\in I_1I_2$. This proves that $I_1\cap I_2\subseteq I_1I_2$, and hence $I_1\cap I_2=I_1I_2$. Now for the inductive step. By the inductive hypothesis, we know that \[I_2\cap \cdots \cap I_n=I_2\cdots I_n,\] and therefore \[I_1\cap (I_2\cap \cdots \cap I_n)=I_1\cap(I_2\cdots I_n).\] We would like to show that \[I_1\cap(I_2\cdots I_n)=I_1I_2\cdots I_n.\] This will follow from the $n=2$ case, provided that we can show that \[I_1+(I_2\cdots I_n)=A.\] Recall that in the proof of part 1, we constructed a $p_1\in A$ such that $1-p_1\in I_1$ and \[p_1\in \prod_{j\neq i}I_j= I_2\cdots I_n.\] Then the fact that \[1=p_1+(1-p_1)\] implies that $1\in I_1+(I_2\cdots I_n)$, and hence $I_1+(I_2\cdots I_n)=A$. \end{proof} Because the $I_i$ are two-sided ideals, we can quotient the ring $A$ by them. Define the homomorphisms $\pi_i:A\to A/I_i$ to be the quotient maps, so that $\ker(\pi_i)=I_i$. Define $\pi$ to be the composition \[\pi:A\xrightarrow{\;\text{diag}\;}\bigoplus_{i=1}^n A\xrightarrow{\;\oplus \pi_i\;}\bigoplus_{i=1}^n A/I_i,\] so that \[\pi(a)=(a\bmod I_1,\ldots,a\bmod I_n).\] Clearly, $\ker(\pi)=\bigcap \ker(\pi_i)=\bigcap I_i$. We can now restate the Chinese Remainder Theorem in a more abstract form. \newtheorem*{theorem-lecture1-crt2}{Chinese Remainder Theorem} \begin{theorem-lecture1-crt2} Let $A$ be a not-necessarily-commutative ring, and $I_1,\ldots,I_n$ two-sided ideals of $A$ such that $I_i+I_j=A$ for all $i\neq j$. Then the map $\pi$ is surjective, and induces an isomorphism \[\overline{\pi}:A/(I_1\cap\cdots\cap I_2)\to \bigoplus_{i=1}^n A/I_i.\] \end{theorem-lecture1-crt2} \begin{proof} Because we induced the map $\overline{\pi}$ by quotienting out by the kernel of $\pi$, we have that $\ker(\overline{\pi})=0$. Therefore, $\overline{\pi}$ is injective. Also, $\overline{\pi}$ is surjective, because for any choice of $\overline{c_i}\in A/I_i$ for each $i$, there is some $a\in A$ such that $a\bmod I_i=\overline{c_i}$ by the Chinese Remainder Theorem. \end{proof} Now we will review some basics about PIDs. A ring $A$ is a PID when \begin{enumerate} \item $A$ is commutative, \item $A$ has no zero-divisors, and \item Any ideal of $A$ is principal. \end{enumerate} Examples of PIDs include the ring $\Z$, and the polynomial ring $k[t]$ over a field $k$. Let $A$ be a PID. For any $a,b\in A\setminus\{0\}$, we say that $d\in A$ is a $\gcd$ of $a$ and $b$ when $d\mid a$ and $d\mid b$, and when for any $g\in A$ such that $g\mid a$ and $g\mid b$, we also have $g\mid d$. The following is a fundamental result about PIDs. \begin{theorem} Let $A$ be a PID. For any $a,b\in A\setminus\{0\}$, \[Aa+Ab=A\gcd(a,b).\] \end{theorem} \begin{proof} Since $A$ is a PID, we know the ideal $Aa+Ab$ is equal to $Ad$ for some $d$. Note that \[a\in Aa\subset Aa+Ab=Ad\] implies that $d\mid a$. By symmetry, $d\mid b$ as well. If $g\mid a$ and $g\mid b$, then $a,b\in Ag$, so that \[Ad=Aa+Ab\subset Ag,\] which implies that $d\in Ag$ and hence $g\mid d$. \end{proof} \begin{corollary} Let $A$ be a PID. Then \[\gcd(a,b)=1\iff Aa+Ab=A\iff\text{ there exist }u,v\in A\text{ such that }au+bv=1.\] \end{corollary} \begin{corollary} Let $A$ be a PID, and let $a=a_1\cdots a_n$ where $\gcd(a_i,a_j)=1$ for all $i\neq j$. Then \[A/(a)\cong A/(a_1)\oplus\cdots\oplus A/(a_n).\] \end{corollary}