Browse Source

port initial notes on commutative algebra: ideals & modules

commutative-alg
arnaucube 1 week ago
parent
commit
0de6f6c815
4 changed files with 545 additions and 0 deletions
  1. +1
    -0
      .gitignore
  2. +1
    -0
      README.md
  3. BIN
      commutative-algebra-notes.pdf
  4. +543
    -0
      commutative-algebra-notes.tex

+ 1
- 0
.gitignore

@ -12,3 +12,4 @@
*.snm
*.vrb
galois-theory-notes.bib
commutative-algebra-notes.bib

+ 1
- 0
README.md

@ -7,6 +7,7 @@ Notes, code and documents done while reading books and papers.
- [Notes on "Abstract Algebra" book, by Charles C. Pinter](abstract-algebra-charles-pinter-notes.pdf)
- [Notes on Weil pairing](weil-pairing.pdf)
- [Notes on Galois Theory](galois-theory-notes.pdf)
- [Notes on Commutative Algebra](commutative-algebra-notes.pdf)
In-between math & crypto:

BIN
commutative-algebra-notes.pdf


+ 543
- 0
commutative-algebra-notes.tex

@ -0,0 +1,543 @@
\documentclass{article}
\usepackage[utf8]{inputenc}
\usepackage{amsfonts}
% \usepackage{yfonts} % WIP
\usepackage{amsthm}
\usepackage{amsmath}
\usepackage{enumerate}
\usepackage{hyperref}
\usepackage{amssymb}
\usepackage{tikz} % diagram
\begin{filecontents}[overwrite]{commutative-algebra-notes.bib}
@misc{am,
author = {M. F. Atiyah and I. G. MacDonald},
title = {{Introduction to Commutative Algebra}},
year = {1969}
}
@misc{reid,
author = {Miles Reid},
title = {{Undergraduate Commutative Algebra}},
year = {1995}
}
@misc{mit-course,
author = {Steven Kleiman},
title = {{Commutative Algebra - MIT OpenCourseWare}},
year = {2008},
note = {\url{https://ocw.mit.edu/courses/18-705-commutative-algebra-fall-2008/}},
url = {https://ocw.mit.edu/courses/18-705-commutative-algebra-fall-2008/}
}
\end{filecontents}
\nocite{*}
\theoremstyle{definition}
\newtheorem{innerdefn}{Definition}
\newenvironment{defn}[1]
{\renewcommand\theinnerdefn{#1}\innerdefn}
{\endinnerdefn}
\newtheorem{innerthm}{Theorem}
\newenvironment{thm}[1]
{\renewcommand\theinnerthm{#1}\innerthm}
{\endinnerthm}
\newtheorem{innerlemma}{Lemma}
\newenvironment{lemma}[1]
{\renewcommand\theinnerlemma{#1}\innerlemma}
{\endinnerlemma}
\newtheorem{innerprop}{Proposition}
\newenvironment{prop}[1]
{\renewcommand\theinnerprop{#1}\innerprop}
{\endinnerprop}
\newtheorem{innercor}{Corollary}
\newenvironment{cor}[1]
{\renewcommand\theinnercor{#1}\innercor}
{\endinnercor}
\newtheorem{innereg}{Example}
\newenvironment{eg}[1]
{\renewcommand\theinnereg{#1}\innereg}
{\endinnereg}
\newtheorem{innerex}{Exercise}
\newenvironment{ex}[1]
{\renewcommand\theinnerex{#1}\innerex}
{\endinnerex}
\newcommand{\aA}{\mathfrak{a}} % TODO: use goth font
\newcommand{\mM}{\mathfrak{m}}
\title{Commutative Algebra notes}
\author{arnaucube}
\date{}
\begin{document}
\maketitle
\begin{abstract}
Notes taken while studying Commutative Algebra, mostly from Atiyah \& MacDonald book \cite{am} and Reid's book \cite{reid}.
Usually while reading books and papers I take handwritten notes in a notebook, this document contains some of them re-written to $LaTeX$.
The proofs may slightly differ from the ones from the books, since I try to extend them for a deeper understanding.
\end{abstract}
\tableofcontents
\section{Ideals}
\subsection{Definitions}
% WIP
\subsection{Lemmas, propositions and corollaries}
\begin{thm}{1.X}{Zorn's lemma} \label{zorn}
TODO
\end{thm}
\begin{thm}{1.3} \label{1.3}
Every ring $A \neq 0$ has at lleast one maximal ideal.
\end{thm}
\begin{proof}
By Zorn's lemma \ref{zorn}.
\end{proof}
\begin{cor}{1.4} \label{1.4}
if $I \neq (1)$ an ideal of $A$, $\exists$ a maximal ideal of $A$ containing $I$.
\end{cor}
\begin{cor}{1.5} \label{1.5}
Every non-unit of $A$ is contained in a maximal ideal.
\end{cor}
\begin{defn}{Jacobson radical}
The \emph{Jacobson radical} of a ring $A$ is the intersection of all the maximal ideals of $A$.
Denoted $Jac(A)$.
$Jac(A)$ is an ideal of $A$.
\end{defn}
\begin{prop}{1.9} \label{1.9}
$x \in Jac(A)$ iff $(1 - xy)$ is a unit in $A$, $\forall y \in A$.
\end{prop}
\begin{proof}
Suppose $1-xy$ not a unit.
By \ref{1.5}, $1-xy \in \mM$ for $\mM$ some maximal ideal.
But $x \in Jac(A) \subseteq \mM$, since $Jac(A)$ is the intersection of all maximal ideals of $A$.
Hence $xy \in \mM$, and therefore $1 \in \mM$, which is absurd, thus $1-xy$ is a unit.
Conversely:\\
Suppose $x \not\in \mM$ for some maximal ideal $\mM$.
Then $\mM$ and $x$ generte the unit ideal $(1)$, so that we have $u + xy = 1$ for some $u \in \mM$ and some $y \in A$.
Hence $1 -xy \in \mM$, and is therefore not a unit.
\end{proof}
\section{Modules}
\subsection{Modules}
\subsection{Cayley-Hamilton theorem, Nakayama lemma, and corollaries}
\begin{prop}{2.4}(Cayley-Hamilton Theorem) \label{2.4}
Let $M$ a fingen $A$-module. Let $\aA$ an ideal of $A$, let $\psi$ an
$A$-module endomorphism of $M$ such that $\psi(M) \subseteq \aA M$.
Then $\psi$ satisfies
$$\psi^n + a_1 \psi^{n-1} + \ldots + a_{n-1} \psi + a_n = 0$$
with $a_i \in \aA$.
\end{prop}
\begin{proof}
Since $M$ fingen, let $\{ x_1, \ldots, x_n \}$ be generators of $M$.\\
By hypothesis, $\psi(M) \subseteq \aA M$; so for any generator $x_i$, it's image $\psi(x_i) \in \aA M$.
Any element in $\aA M$ is a linear combination of the generators with coefficients in the ideal $\aA$, thus
$$\psi(x_i)= \sum_{j=1}^n a_{ij} x_j$$
with $a_{ij} \in \aA$.
Thus, for a module with $n$ generators, we have $n$ different $\psi(x_i)$ equations:
$$
\left.
\begin{aligned}
\psi(x_1) &= a_{1,1} x_1 + a_{1,2} x_2 + \ldots + a_{1,n} x_n\\
\psi(x_2) &= a_{2,1} x_1 + a_{2,2} x_2 + \ldots + a_{2,n} x_n\\
\ldots\\
\psi(x_n) &= a_{n,1} x_1 + a_{n,2} x_2 + \ldots + a_{n,n} x_n
\end{aligned}
\right\}
\begin{aligned}
&\text{n elements $\psi(x_i) \in \aA M$ which}\\
&\text{are linear combinations of the}\\
&\text{$n$ generators of $M$}
\end{aligned}
$$
Next step: rearrange in order to use matrix algebra.
Observe that each row equals $0$, and rearranging the elements at each row we get
\begin{align*}
&\psi(x_1) - (a_{1,1} x_1 + a_{1,2} x_2 + \ldots + a_{1,n} x_n) = 0\\
&\psi(x_2) - (a_{2,1} x_1 + a_{2,2} x_2 + \ldots + a_{2,n} x_n) = 0\\
&\ldots\\
&\psi(x_n) - (a_{n,1} x_1 + a_{n,2} x_2 + \ldots + a_{n,n} x_n) = 0
\end{align*}
Then, group the $x_i$ terms together; as example, take the row $i=1$:
$$(\psi - a_{1,1})x_1 - a_{1,2} x_2 - \ldots - a_{1,n} x_n = 0$$
for $i=2$:
$$-a_{2,1} x_1 + (\psi - a_{2,2}) x_2 - \ldots - a_{2,n} x_n = 0$$
So, $\forall i \in [n]$, as a matrix:
$$
\begin{pmatrix}
\psi - a_{1,1} & -a_{1,2} & \ldots & -a_{1,n}\\
-a_{2,1} & \psi-a_{2,2} & \ldots & -a_{2,n}\\
\vdots\\
-a_{n,1} & -a_{n,2} & \ldots & \psi-a_{n,n}\\
\end{pmatrix}
\begin{pmatrix}
x_1\\ x_2\\ \vdots\\ x_n
\end{pmatrix}
=
\begin{pmatrix}
0\\ 0\\ \vdots\\ 0
\end{pmatrix}
$$
Kronecker delta:
$\delta_{ij} =
\begin{cases}
1 & \text{if } i = j,\\
0 & \text{otherwise}
\end{cases}$
With the Kronecker delta, $\psi(x_i)$ can be expressed as
$$\psi(x_i) = \sum_{j=1}^n \delta_{ij} \psi(x_j)$$
so the previous matrix ccan be characterized as
$$\sum_{j=1}^n (\delta_{ij} \psi - a_{ij}) x_j = 0$$
The entries of the matrix are \emph{endomorphisms} (elements of the ring $A[\psi]$)
\begin{itemize}
\item the term $(\psi - a_{11})$ is an operator that acts on $x_1$; as $(\psi(x_1)-a_{11}\cdot x_1)$
\item the term $(-a_{12})$ is an operator that acts on $x_2$; as multiplication by it, ie. $(-a_{12} \cdot x_2)$
\end{itemize}
Since $A$ is a commutative ring, and $\psi$ commutes with any $a \in A$,
the ring of operators $A[\psi]$ is a commutative ring.
$\Longrightarrow~$ so we can treat the matrix as a matrix of real numbers and calculate its determinant.
We're interested in the determinant because it is the only way to turn a system of multiple equations in a single scalar-like equation that describes the endomorphism $\psi$.\\
$\rightarrow$ Because in module theory, we lack of "division", so can not "solve for $\psi$" the system of equations.\\
$\rightarrow$ The determinant provides a way to find a polynomial that \emph{annihilates} the module; the \emph{characteristic polynomial}, which related $\psi$ to the ideal $\aA$
$$det(M) \cdot x_i = 0~~ \forall i$$
where $x_i$ are the generators of $M$.
Since $det(M)$ kills every generator, it must kill every element in $M$\\
$\Longrightarrow~~ det(M)$ is the zero map.
Leibniz formula of the determinant of an $n \times n$ matrix:
$$
det(M) = \sum_{\sigma \in S_n} sign(\sigma) \prod_{i=1}^n M_{i, \sigma(i)}
$$
so,
$$(\psi - a_{11}) (\psi - a_{22}) \ldots (\psi - a_{nn})$$
expanding it,
\begin{itemize}
\item highest power is $1 \cdot \psi^n$
\item coefficient of $\psi^{n-1}$ is $-( \underbrace{ a_{11} + a_{22} + \ldots + a_{nn} }_{a_1})$,\\
where, since each $a_{ii} \in \aA,~~ a_1 \in \aA$
\item the rest of coefficients of $\psi^k$ are also elements in $\aA$
\end{itemize}
So we have
$$p(\psi) = \psi^n + a_1 \psi^{n-1} + a_2 \psi^{n-2} + \ldots + a_{n-1} \psi + a_n$$
with $a_i \in \aA$.
Since this determinant annihilates the generators (ie. $det(M)x_i=0$), the resulting enddomorphism $p(\psi)$ is the zero map on the entire module $M$, so:
$$\psi^n + a_1 \psi^{n-1} + a_2 \psi^{n-2} + \ldots + a_{n-1} \psi + a_n = 0$$
with $a_i \in \aA$, as stated in the Cayley-Hamilton theorem.
\end{proof}
\begin{cor}{2.5} \label{2.5}
Let $M$ a fingen $A$-module, let $\aA$ an ideal of $A$ such that $\aA M = M$.
Then, $\exists~ x \equiv 1 \pmod \aA$ such that $xM = 0$.
\end{cor}
\begin{proof}
take $\psi = \text{identity}$. Then in Cayley-Hamilton (\ref{2.4}):
\begin{align*}
&\psi^n + a_1 \psi^{n-1} + a_2 \psi^{n-2} + \ldots + a_{n-1} \psi + a_n = 0\\
\Longrightarrow~ &id_M + a_1 id_M + a_2 id_M + \ldots + a_{n-1} id_M + a_n = 0\\
\Longrightarrow~ &(1 + a_1 + \ldots + a_n) id_M = 0
\end{align*}
apply it to $m \in M$, where since $id_M(m)=m$ (by definition of the identity), we then have
$$(1 + a_1 + \ldots + a_n) \cdot m = 0$$
with $a_i \in \aA$.
\begin{enumerate}[\text{part} i.]
\item $xM=0$:\\
Thus the scalar $x = (1 + a_1 + \ldots + a_n)$ annihilates every $m \in M$, ie. the entire module $M$.
\item $x \equiv 1 \pmod \aA$:\\
$x \equiv 1 \pmod \aA ~~ \Longleftrightarrow (x-1) \in \aA$\\
then from $x = (1 + \underbrace{a_1 + \ldots + a_n}_b) \in \aA$, set $b=a_1 + \ldots + a_n$,\\
so that $x=(1+b) \in \aA$.\\
Then $x-1 = (1+b)-1 = b \in \aA$\\
so $x-1 \in \aA$, thus $x \equiv 1 \pmod \aA$ as stated.
\end{enumerate}
\end{proof}
\begin{prop}{2.6}{Nakayama's lemma} \label{2.6}
Let $M$ a fingen $A$-module, let $\aA$ an ideal of $A$ such that $\aA \subseteq Jac(A)$.
Then $\aA M = M$ implies $M=0$.
\end{prop}
\begin{proof}
By \ref{2.5}: since $\aA M = M$, we have $x M =0$ for some $x \equiv 1 \pmod {Jac(A)}$. (notice that at \ref{2.5} is $\pmod \aA$ but here we use $\pmod {Jac(A)}$, since we have $\aA \subseteq Jac(A)$).
By \ref{1.9}, $x$ is a unit in $A$.
Hence $M = x^{-1} \cdot \underbrace{x \cdot M}_{=0~ \text{(by \ref{2.5})}} = 0$.
Thus, if $\aA M = M$ then $M=0$.
\end{proof}
\begin{cor}{2.7} \label{2.7}
Let $M$ a fingen $A$-module, let $N \subseteq M$ a submodule of $M$, let $\aA \subseteq Jac(A)$ an ideal.
Then $M = \aA M + N \stackrel{\text{implies}}{\Longrightarrow} M=N$.
\end{cor}
\begin{proof}
The idea is to apply Nakayama (\ref{2.6}) to $M/N$.
Since $M$ fingen $\Longrightarrow~~ M/N$ is fingen and an $A$-module.
Since $\aA \subseteq Jac(A) ~\Longrightarrow~$ Nakayama applies to $M/N$ too.
By definition,
$$\aA M = \left\{ \sum a_i \cdot m_i ~~|~~ a_i \in \aA, m_i \in M \right\}$$
where $m_i$ are the generators of $M$.
Then, for $M/N$,
$$\aA (\frac{M}{N}) = \left\{ \sum a_i \cdot (m_i + N) ~~|~~ a_i \in \aA, m_i \in M \right\}$$
observe that $a_i(m_i+N)= a_i m_i +N$, thus
$$\sum_i a_i \cdot (m_i + N) = \underbrace{(\sum_i a_i \cdot m_i)}_{\in \aA M} + N \in \aA M + N$$
Hence,
\begin{equation}
\aA (\frac{M}{N}) = \left\{ x + N ~~|~~ x \in \aA M \right\} = \aA M + N
\label{eq:2.7.1}
\end{equation}
By definition, if we take $\frac{\aA M + N}{N}$, then
$$\frac{\aA M + N}{N} = \left\{ y + N ~~|~~ y \in \aA M +N \right\} = \aA M + N$$
thus every $y \in \aA M +N$ can be written as
$$y=x+n,~~ \text{with} x \in \aA M,~ n\in M$$
which comes from \eqref{eq:2.7.1}.
Thus, $y + N = (x+n)+N = x+N$, since $n \in N$ is zero in the quotient.
Hence, every element of $\frac{\aA M +N}{N}$ has the form
$$\frac{\aA M + N}{N} = \left\{ x + N ~~|~~ x \in \aA M \right\}$$
as in \eqref{eq:2.7.1}.
Thus
\begin{equation}
\aA (\frac{M}{N}) = \aA M + N = \frac{\aA M +N}{N}
\label{eq:2.7.2}
\end{equation}
By the Collorary assumption, $M = \aA M + N$; quotient it by $N$:
\begin{equation}
\frac{M}{N} = \frac{\aA M +N}{N}
\label{eq:2.7.3}
\end{equation}
So, from \eqref{eq:2.7.2} and \eqref{eq:2.7.3}:
$$\aA (\frac{M}{N}) = \aA M +N = \frac{\aA M +N}{N} = \frac{M}{N}$$
thus, $\aA (\frac{M}{N}) = \frac{M}{N}$.
By Nakayama's lemma \ref{2.6}, if $\aA (\frac{M}{N}) = \frac{M}{N} ~\stackrel{implies}{\Longrightarrow}~ \frac{M}{N}=0$
Note that
$$\frac{M}{N} = \{ m + N ~|~ m \in M \}$$
(the zero element in $\frac{M}{N}$ is the coset $N=0+N$)
Then, $\frac{M}{N}=0$ means that the quotient has exactly one element, the zero coset $N$.
Thus, every coset $m + N$ equals the zero coset $N$, so $m-0 \in N ~\Longrightarrow~ m \in N$.
Hence every $m \in M$ lies in $N$, ie. $\forall m \in M,~ m \in N$.
So $M \subseteq N$. But notice that by the Corollary, we had $N \subseteq M$, therefore $M = N$.
Thus, if $M = \aA M + N \stackrel{implies}{\Longrightarrow} M = N$.
\end{proof}
\begin{prop}{2.8} \label{2.8}
Let $x_i \forall i \in [n]$ be elements of $M$ whose images $\frac{M}{m M}$ from a basis of this vecctor space. Then the $x_i$ generate $M$.
\end{prop}
\begin{proof}
Let $N$ submodule $M$, generated by the $x_i$.
Then the composite map $N \longrightarrow M \longrightarrow \frac{M}{m M}$ maps $N$ onto $\frac{M}{m M}$, hence $N + \aA M = M$, which by \ref{2.7} implies $N = M$.
\end{proof}
\subsection{Noetherean rings}
\section{Exercises}
For the exercises, I follow the assignements listed at \cite{mit-course}.
The exercises that start with \textbf{R} are the ones from the book \cite{reid}, and the ones starting with \textbf{AM} are the ones from the book \cite{am}.
\subsection{Exercises Chapter 1}
\begin{ex}{R.1.1}
Ring $A$ and ideals $I, J$ such that $I \cup J$ is not an ideal. What's the smallest ideal containing $I$ and $J$?
\end{ex}
\begin{proof}
Take ring $A= \mathbb{Z}$. Set $I = 2 \mathbb{Z},~ J=3 \mathbb{Z}$.
$I,~J$ are ideals of $A$ ($=\mathbb{Z}$). And $I \cup J = 2 \mathbb{Z} \cup 3 \mathbb{Z}$.\\
Observe that for $2 \in I,~ 3 \in J ~\Longrightarrow~ 2,3 \in I \cup J$, but $2+3 = 5 \not\in I \cup J$.
Thus $I \cup J$ is not closed under addition; thus is not an ideal.
Smallest ideal of $\mathbb{Z}$ ($=A$) containing $I$ and $J$ is their sum:
$$I+J = \{ a+b | a \in I, b \in J \}$$
$gcd(2,3)=1$, so $I+J = \mathbb{Z}$.
Therefore, smallest ideal containing $I$ and $J$ is the whole ring $\mathbb{Z}$.
\end{proof}
\begin{ex}{R.1.5}
let $\psi: A \longrightarrow B$ a ring homomorphism. Prove that $\psi^{-1}$ takes prime ideals of $B$ to prime ideals of $A$.\\
In particular if $A \subset B$ and $P$ a prime ideal of $B$, then $A \cap P$ is a prime ideal of $A$.
\end{ex}
\begin{proof}
(Recall: prime ideal is if $a,b \in R$ and $a \cdot b \in P$ (with $R \neq P$), implies $a \in P$ or $b \in P$).
Let
$$\psi^{-1}(P) = \{ a \in A | \psi(a) \in P \} = A \cap P$$
The claim is that $\psi^{-1}(P)$ is prime iddeal of $A$.
\begin{enumerate}[i.]
\item show that $\psi^{-1}(P)$ is an ideal of $A$:\\
$0_A \in \psi^{-1}(P)$, since $\psi(0_A)=0_B \in P$ (since every ideal contains $0$).
If $a,b \in \psi^{-1}(P)$, then $\psi(a), \psi(b) \in P$, so
$$\psi(a-b)= \psi(a) - \psi(b) \in P$$
hence $a-b \in \psi^{-1}(P)$.
If $a \in \psi^{-1}(P)$ and $r \in A$, then $\psi(ra) = \psi(r) \psi(a) \in P$, since $P$ is an ideal.\\
Thus $ra \in \psi^{-1}(P)$.
$\Longrightarrow$ so $\psi^{-1}$ is an ideal of $A$.
\item show that $\psi^{-1}(P)$ is prime:\\
$\psi^{-1}(P) \neq A$, since if $\psi^{-1}(P)=A$, then $1_A \in \psi^{-1}(P)$, so $\psi(1_A)=1_B \in P$, which would mean that $P=B$, a contradiction since $P$ is prime ideal of $B$.
Take $a,b \in A$ with $ab \in \psi^{-1}(P)$; then $\psi(ab) \in P$, and since $\psi$ is a ring homomorphism, $\psi(ab) = \psi(a)\psi(b)$.
Since $P$ prime ideal, then $\psi(a)\psi(b) \in P$ implies either $\psi(a) \in P$ or $\psi(b) \in P$.\\
Thus $a \in \psi^{-1}(P)$ or $b \in \psi^{-1}(P)$.
Hence $\psi^{-1}(P)$ ($=A \cap P$) is a prime ideal of $A$.
\end{enumerate}
\end{proof}
\begin{ex}{R.1.6}
prove or give a counter example:
\begin{enumerate}[a.]
\item the intersection of two prime ideals is prime
\item the ideal $P_1+P_2$ generated by $2$ prime ideals $P_1,P_2$ is prime
\item if $\psi: A \longrightarrow B$ ring homomorphism, then $\psi^{-1}$ takes maximal ideals of $B$ to maximal ideals of $A$
\end{enumerate}
\end{ex}
\begin{proof}
\begin{enumerate}[a.]
\item let $I = 2 \mathbb{Z} = (2)$, $J = 3 \mathbb{Z} = (3)$ be ideals of $\mathbb{Z}$, both prime.
Then $I \cap J = (2) \cap (3) = (6)$.
The ideal $(6)$ is not prime in $\mathbb{Z}$, since $2 \cdot 3 \in (6)$, but $2 \neq (6)$ and $3 \neq (6)$.
Thus the intersection of two primes can not be prime.
\item $P_1=(2),~ P_2=(3)$, both prime.
Then,
$$P_1 + P_2 = (2)+(3)=\{ a+b | a \in P_1, b \in P_2 \}$$
$\longrightarrow~$ in a principal ideal domain (like $\mathbb{Z}$), the sum of two principal ideals is again principal, and given by $(m)+(n)=(gcd(m,n))$.
(recall: principal= generated by a single element)
So, $P_1+P_2= (2)+(3) = (gcd(2,3))=(1)=\mathbb{Z}$.
The whole ring is not a prime ideal (by the definition of the prime ideal), so $P_1+P_2$ is not a prime ideal.
Henceforth, the sum of two prime ideals is not necessarily prime.
\item let $A=\mathbb{Z},~ B=\mathbb{Q},~ \psi: A \longrightarrow B$.
Since $\mathbb{Q}$ is a field, its only maximal ideal is $(0)$.
Then
\begin{align*}
\psi^{-1}( (0) ) &= (0) \subset \mathbb{Z}\\
\text{ie.}~ \psi^{-1}( m_B ) &= (m_B) \subset A
\end{align*}
But $(0)$ is not maximal in $\mathbb{Z}$, because $\mathbb{Z}/(0) \cong \mathbb{Z}$ is not a field.
Thus the preimages of maximal ideals under arbitrary ring homomorphisms need not be maximal.
\end{enumerate}
\end{proof}
\subsection{Exercises Chapter 2}
\bibliographystyle{unsrt}
\bibliography{commutative-algebra-notes.bib}
\end{document}

Loading…
Cancel
Save