Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Rings and ideals} | |
\section{Some motivational metaphors about rings vs groups} | |
In this chapter we'll introduce the notion of | |
a \textbf{commutative ring} $R$. | |
It is a larger structure than a group: | |
it will have two operations addition and multiplication, | |
rather than just one. | |
We will then immediately define a \textbf{ring homomorphism} | |
$R \to S$ between pairs of rings. | |
This time, instead of having normal subgroups $H \normalin G$, | |
rings will instead have subsets $I \subseteq R$ called \textbf{ideals}, | |
which are not themselves rings but satisfy some niceness conditions. | |
We will then show how you to define $R/I$, | |
in analogy to $G/H$ as before. | |
Finally, like with groups, we will talk a bit about how to generate ideals. | |
Here is a possibly helpful table of analogies to help you keep track: | |
\begin{center} | |
\begin{tabular}[h]{lcc} | |
& Group & Ring \\ | |
\hline | |
Notation & $G$ & $R$ \\ | |
Operations & $\cdot$ & $+$, $\times$ \\ | |
Commutativity & only if abelian & for us, always \\ | |
Sub-structure & subgroup & (not discussed) \\ | |
Homomorphism & grp hom.\ $G \to H$ & ring hom.\ $R \to S$ \\ | |
Kernel & normal subgroup & ideal \\ | |
Quotient & $G/H$ & $R/I$ \\ | |
\end{tabular} | |
\end{center} | |
\section{(Optional) Pedagogical notes on motivation} | |
I wrote most of these examples with a number theoretic eye in mind; | |
thus if you liked elementary number theory, | |
a lot of your intuition will carry over. | |
Basically, we'll try to generalize properties of the ring $\ZZ$ to | |
any abelian structure in which we can also multiply. | |
That's why, for example, you can talk about | |
``irreducible polynomials in $\QQ[x]$'' in the same | |
way you can talk about ``primes in $\ZZ$'', | |
or about ``factoring polynomials modulo $p$'' | |
in the same way we can talk ``unique factorization in $\ZZ$''. | |
Even if you only care about $\ZZ$ | |
(say, you're a number theorist), this has a lot of value: | |
I assure you that trying to solve $x^n+y^n = z^n$ (for $n > 2$) | |
requires going into a ring other than $\ZZ$! | |
Thus for all the sections that follow, keep $\ZZ$ in mind as your prototype. | |
I mention this here because | |
commutative algebra is \emph{also} closely tied to algebraic geometry. | |
Lots of the ideas in commutative algebra have nice | |
``geometric'' interpretations that motivate the definitions, | |
and these connections are explored in the corresponding part later. | |
So, I want to admit outright that this is not | |
the only good way (perhaps not even the most natural one) | |
of motivating what is to follow. | |
\section{Definition and examples of rings} | |
\prototype{$\ZZ$ all the way! Also $R[x]$ and various fields (next section).} | |
Well, I guess I'll define a ring\footnote{Or, | |
according to some authors, a ``ring with identity''; | |
some authors don't require rings to have multiplicative identity. | |
For us, ``ring'' always means ``ring with $1$''.}. | |
\begin{definition} | |
A \vocab{ring} is a triple $(R, +, \times)$, | |
the two operations usually called addition and multiplication, such that | |
\begin{enumerate}[(i)] | |
\ii $(R,+)$ is an abelian group, with identity $0_R$, or just $0$. | |
\ii $\times$ is an associative, binary operation on $R$ with some | |
identity, written $1_R$ or just $1$. | |
\ii Multiplication distributes over addition. | |
\end{enumerate} | |
The ring $R$ is \vocab{commutative} if $\times$ is commutative. | |
\end{definition} | |
\begin{abuse} | |
As usual, we will abbreviate $(R, +, \times)$ to $R$. | |
\end{abuse} | |
\begin{abuse} | |
For simplicity, assume all rings are commutative | |
for the rest of this chapter. | |
We'll run into some noncommutative rings eventually, | |
but for such rings we won't need the full theory of this chapter anyways. | |
\end{abuse} | |
These definitions are just here for completeness. | |
The examples are much more important. | |
\begin{example}[Typical rings] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii The sets $\ZZ$, $\QQ$, $\RR$ and $\CC$ are all rings | |
with the usual addition and multiplication. | |
\ii The integers modulo $n$ are also a ring | |
with the usual addition and multiplication. | |
We also denote it by $\Zc n$. | |
\end{enumerate} | |
\end{example} | |
Here is also a trivial example. | |
\begin{definition} | |
The \vocab{zero ring} is the ring $R$ with a single element. | |
We denote the zero ring by $0$. | |
A ring is \vocab{nontrivial} if it is not the zero ring. | |
\end{definition} | |
\begin{exercise} | |
[Comedic] | |
Show that a ring is nontrivial if and only if $0_R \ne 1_R$. | |
\end{exercise} | |
Since I've defined this structure, I may as well state the obligatory facts about it. | |
\begin{fact} | |
For any ring $R$ and $r \in R$, $r \cdot 0_R = 0_R$. | |
Moreover, $r \cdot (-1_R) = -r$. | |
\end{fact} | |
Here are some more examples of rings. | |
\begin{example} | |
[Product ring] | |
\label{ex:product_ring} | |
Given two rings $R$ and $S$ the \vocab{product ring}, | |
denoted $R \times S$, is defined as ordered pairs $(r,s)$ | |
with both operations done component-wise. | |
For example, the Chinese remainder theorem says | |
that \[ \Zc{15} \cong \Zc3 \times \Zc5 \] | |
with the isomorphism $n \bmod{15} \mapsto (n \bmod 3, n \bmod 5)$. | |
\end{example} | |
\begin{remark} | |
Equivalently, we can define $R \times S$ as the abelian group $R \oplus S$, | |
and endow it with the multiplication where $r \cdot s = 0$ | |
for $r \in R$, $s \in S$. | |
\end{remark} | |
\begin{ques} | |
Which $(r,s)$ is the identity element of the product ring $R \times S$? | |
\end{ques} | |
\begin{example}[Polynomial ring] | |
Given any ring $R$, | |
the \vocab{polynomial ring} $R[x]$ is defined as the set of polynomials | |
with coefficients in $R$: | |
\[ R[x] = \left\{ a_n x^n+a_{n-1}x^{n-1}+\dots+a_0 | |
\mid a_0, \dots, a_n \in R \right\}. \] | |
This is pronounced ``$R$ adjoin $x$''. | |
Addition and multiplication are done exactly in the way you would expect. | |
\end{example} | |
\begin{remark} | |
[Digression on division] | |
Happily, polynomial division also does what we expect: | |
if $p \in R[x]$ is a polynomial, and $p(a) = 0$, | |
then $(x-a)q(x) = p(x)$ for some polynomial $q$. | |
Proof: do polynomial long division. | |
With that, note the caveat that | |
\[ x^2-1 \equiv (x-1)(x+1) \pmod 8 \] | |
has \emph{four} roots $1$, $3$, $5$, $7$ in $\Zc8$. | |
The problem is that $2 \cdot 4 = 0$ even though $2$ and $4$ are not zero; | |
we call $2$ and $4$ \emph{zero divisors} for that reason. | |
In an \emph{integral domain} (a ring without zero divisors), | |
this pathology goes away, | |
and just about everything you know about polynomials carries over. | |
(I'll say this all again next section.) | |
\end{remark} | |
\begin{example} | |
[Multi-variable polynomial ring] | |
We can consider polynomials in $n$ variables with coefficients in $R$, | |
denoted $R[x_1, \dots, x_n]$. | |
(We can even adjoin infinitely many $x$'s if we like!) | |
\end{example} | |
\begin{example} | |
[Gaussian integers are a ring] | |
The \vocab{Gaussian integers} are the set of complex numbers | |
with integer real and imaginary parts, that is | |
\[ \ZZ[i] = \left\{ a+bi \mid a,b \in \ZZ \right\}. \] | |
\end{example} | |
\begin{abuse} | |
[Liberal use of adjoinment] | |
Careful readers will detect some abuse in notation here. | |
$\ZZ[i]$ should officially be | |
``integer-coefficient polynomials in a variable $i$''. | |
However, it is understood from context that $i^2=-1$; | |
and a polynomial in $i = \sqrt{-1}$ ``is'' a Gaussian integer. | |
\end{abuse} | |
\begin{example} | |
[Cube root of $2$] | |
As another example (using the same abuse of notation): | |
\[ \ZZ[\sqrt[3]{2}] = \left\{ a + b\sqrt[3]{2} + c\sqrt[3]4 | |
\mid a,b,c \in \ZZ \right\}. \] | |
\end{example} | |
\section{Fields} | |
\prototype{$\QQ$ is a field, but $\ZZ$ is not.} | |
Although we won't need to know what a field is until next chapter, | |
they're so convenient for examples I will go ahead and introduce them now. | |
As you might already know, if the multiplication is invertible, | |
then we call the ring a field. | |
To be explicit, let me write the relevant definitions. | |
\begin{definition} | |
\label{def:unit} | |
A \vocab{unit} of a ring $R$ | |
is an element $u \in R$ which is invertible: | |
for some $x \in R$ we have $ux = 1_R$. | |
\end{definition} | |
\begin{example} | |
[Examples of units] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii The units of $\ZZ$ are $\pm 1$, | |
because these are the only things which ``divide $1$'' | |
(which is the reason for the name ``unit''). | |
\ii On the other hand, in $\QQ$ everything is a unit (except $0$). | |
For example, $\frac 35$ is a unit since | |
$\frac 35 \cdot \frac 53 = 1$. | |
\ii The Gaussian integers $\ZZ[i]$ have four units: | |
$\pm 1$ and $\pm i$. | |
\end{enumerate} | |
\end{example} | |
\begin{definition} | |
A nontrivial (commutative) ring is a \vocab{field} | |
when all its nonzero elements are units. | |
\end{definition} | |
Colloquially, we say that | |
\begin{moral} | |
A field is a structure where you can add, subtract, multiply, and divide. | |
\end{moral} | |
Depending on context, they are often denoted | |
either $k$, $K$, $F$. | |
\begin{example} | |
[First examples of fields] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii $\QQ$, $\RR$, $\CC$ are fields, | |
since the notion $\frac 1c$ makes sense in them. | |
\ii If $p$ is a prime, then $\Zc p$ is a field, | |
which we denote will usually denote by $\FF_p$. | |
\end{enumerate} | |
The trivial ring $0$ is \emph{not} considered a field, | |
since we require fields to be nontrivial. | |
\end{example} | |
%\begin{remark} | |
% You might say at this point that ``fields are nicer than rings'', | |
% but as you'll see in this chapter, the conditions for | |
% being a field are somehow ``too strong''. | |
% To give an example of what I mean: | |
% if you try to think about the concept of ``divisibility'' | |
% in $\ZZ$, you've stepped into the vast and bizarre realm of | |
% number theory. Try to do the same thing in $\QQ$ and you get nothing: | |
% any nonzero $a$ ``divides'' any nonzero $b$ | |
% because $b = a \cdot \frac ba$. | |
% | |
% I know at least one person who instead | |
% thinks of this as an argument for why people | |
% shouldn't care about number theory | |
% (studying chaos rather than order). | |
%\end{remark} | |
\section{Homomorphisms} | |
\prototype{$\ZZ \to \Zc5$ by modding out by $5$.} | |
This section is going to go briskly -- | |
it's the obvious generalization of all the stuff | |
we did with quotient groups.\footnote{I once found an | |
abstract algebra textbook which teaches rings before groups. | |
At the time I didn't understand why, | |
but now I think I get it -- modding out by things in | |
commutative rings is far more natural, and you can start talking | |
about all the various flavors of rings and fields. | |
You also have (in my opinion) more vivid first examples | |
for rings than for groups. | |
I actually sympathize a lot with this approach --- maybe I'll convert | |
Napkin to follow it one day.} | |
First, we define a homomorphism and isomorphism. | |
\begin{definition} | |
Let $R = (R, +_R, \times_R)$ and $S = (S, +_S, \times_S)$ be rings. | |
A \vocab{ring homomorphism} is a map $\phi : R \to S$ | |
such that | |
\begin{enumerate}[(i)] | |
\ii $\phi(x +_R y) = \phi(x) +_S \phi(y)$ for each $x,y \in R$. | |
\ii $\phi(x \times_R y) = \phi(x) \times_S \phi(y)$ for each $x,y \in R$. | |
\ii $\phi(1_R) = 1_S$. | |
\end{enumerate} | |
If $\phi$ is a bijection then $\phi$ is an \vocab{isomorphism} | |
and we say that rings $R$ and $S$ are \vocab{isomorphic}. | |
\end{definition} | |
Just what you would expect. | |
The only surprise is that we also demand $\phi(1_R)$ to go to $1_S$. | |
This condition is not extraneous: | |
consider the map $\ZZ \to \ZZ$ called ``multiply by zero''. | |
\begin{example} | |
[Examples of homomorphisms] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii The identity map, as always. | |
\ii The map $\ZZ \to \Zc5$ modding out by $5$. | |
\ii The map $\RR[x] \to \RR$ by $p(x) \mapsto p(0)$ | |
by taking the constant term. | |
\ii For any ring $R$, there is a trivial ring homomorphism $R \to 0$. | |
\end{enumerate} | |
\end{example} | |
\begin{example} | |
[Non-examples of homomorphisms] | |
Because we require $1_R$ to $1_S$, some maps that you | |
might have thought were homomorphisms will fail. | |
\begin{enumerate}[(a)] | |
\ii The map $\ZZ \to \ZZ$ by $x \mapsto 2x$ | |
is not a ring homomorphism. | |
Aside from the fact it sends $1$ to $2$, | |
it also does not preserve multiplication. | |
\ii If $S$ is a nontrivial ring, | |
the map $R \to S$ by $x \mapsto 0$ is not | |
a ring homomorphism, even though it preserves multiplication. | |
\ii There is no ring homomorphism $\Zc{2016} \to \ZZ$ at all. | |
\end{enumerate} | |
In particular, whereas for groups $G$ and $H$ | |
there was always a trivial group homomorphism sending | |
everything in $G$ to $1_H$, this is not the case for rings. | |
\end{example} | |
\section{Ideals} | |
\prototype{The multiples of $5$ are an ideal of $\ZZ$.} | |
Now, just like we were able to mod out by groups, | |
we'd also like to define quotient rings. | |
So once again, | |
\begin{definition} | |
The \vocab{kernel} of a ring homomorphism $\phi \colon R \to S$, | |
denoted $\ker \phi$, is the set of $r \in R$ such that $\phi(r) = 0$. | |
\end{definition} | |
In group theory, we were able to characterize the ``normal'' subgroups by a few | |
obviously necessary conditions (namely, $gHg\inv = H$). | |
We can do the same thing for rings, and it's in fact easier because our operations are commutative. | |
First, note two obvious facts: | |
\begin{itemize} | |
\ii If $\phi(x) = \phi(y) = 0$, then $\phi(x+y) = 0$ as well. | |
So $\ker \phi$ should be closed under addition. | |
\ii If $\phi(x) = 0$, then for any $r \in R$ we have | |
$\phi(rx) = \phi(r)\phi(x) = 0$ too. | |
So for $x \in \ker \phi$ and \emph{any} $r \in R$, | |
we have $rx \in \ker\phi$. | |
\end{itemize} | |
A (nonempty) subset $I \subseteq R$ is called | |
an ideal if it satisfies these properties. | |
That is, | |
\begin{definition} | |
A nonempty subset $I \subseteq R$ is an \vocab{ideal} | |
if it is closed under addition, and for each $x \in I$, | |
$rx \in I$ for all $r \in R$. | |
It is \vocab{proper} if $I \neq R$. | |
\end{definition} | |
Note that in the second condition, $r$ need not be in $I$! | |
So this is stronger than merely saying $I$ is closed under multiplication. | |
\begin{remark} | |
If $R$ is not commutative, we also need the condition $xr \in I$. | |
That is, the ideal is \emph{two-sided}: it absorbs multiplication | |
from both the left and the right. | |
But since rings in Napkin are commutative | |
we needn't worry with this distinction. | |
\end{remark} | |
\begin{example} | |
[Prototypical example of an ideal] | |
Consider the set $I = 5\ZZ = \{\dots,-10,-5,0,5,10,\dots\}$ as an ideal in $\ZZ$. | |
We indeed see $I$ is the kernel of the ``take mod $5$'' homomorphism: | |
\[ \ZZ \surjto \ZZ/5\ZZ. \] | |
It's clearly closed under addition, | |
but it absorbs multiplication from \emph{all} elements of $\ZZ$: | |
given $15 \in I$, $999 \in \ZZ$, we get $15 \cdot 999 \in I$. | |
\end{example} | |
\begin{exercise} | |
[Mandatory: fields have two ideals] | |
If $K$ is a field, show that $K$ has exactly two ideals. | |
What are they? | |
\label{exer:field_ideal} | |
\end{exercise} | |
Now we claim that these conditions are sufficient. | |
More explicitly, | |
\begin{theorem} | |
[Ring analog of normal subgroups] | |
Let $R$ be a ring and $I \subsetneq R$. | |
Then $I$ is the kernel of some homomorphism if and only if it's an ideal. | |
\end{theorem} | |
\begin{proof} | |
It's quite similar to the proof for the normal subgroup thing, | |
and you might try it yourself as an exercise. | |
Obviously the conditions are necessary. | |
To see they're sufficient, we \emph{define} a ring by ``cosets'' | |
\[ S = \left\{ r + I \mid r \in R \right\}. \] | |
These are the equivalence classes under $r_1 \sim r_2$ if and only if $r_1 - r_2 \in I$ | |
(think of this as taking ``mod $I$''). | |
To see that these form a ring, we have to check that the addition | |
and multiplication we put on them is well-defined. | |
Specifically, we want to check that if $r_1 \sim s_1$ and $r_2 \sim s_2$, | |
then $r_1 + r_2 \sim s_1 + s_2$ and $r_1r_2 \sim s_1s_2$. | |
We actually already did the first part | |
-- just think of $R$ and $S$ as abelian | |
groups, forgetting for the moment that we can multiply. | |
The multiplication is more interesting. | |
\begin{exercise} | |
[Recommended] | |
Show that if $r_1 \sim s_1$ and $r_2 \sim s_2$, then $r_1r_2 \sim s_1s_2$. | |
You will need to use the fact that $I$ absorbs multiplication | |
from \emph{any} elements of $R$, not just those in $I$. | |
\end{exercise} | |
Anyways, since this addition and multiplication is well-defined there | |
is now a surjective homomorphism $R \to S$ with kernel exactly $I$. | |
\end{proof} | |
\begin{definition} | |
Given an ideal $I$, we define as above the \vocab{quotient ring} | |
\[ R/I \defeq \left\{ r+I \mid r \in R \right\}. \] | |
It's the ring of these equivalence classes. | |
This ring is pronounced ``$R$ mod $I$''. | |
\end{definition} | |
\begin{example}[$\ZZ/5\ZZ$] | |
The integers modulo $5$ formed by ``modding out additively by $5$'' | |
are the $\Zc 5$ we have already met. | |
\end{example} | |
But here's an important point: | |
just as we don't actually think of $\ZZ/5\ZZ$ as consisting of | |
$k + 5\ZZ$ for $k=0,\dots,4$, | |
we also don't really want to think about $R/I$ as elements $r+I$. | |
The better way to think about it is | |
\begin{moral} | |
$R/I$ is the result when we declare that elements of $I$ are all zero; | |
that is, we ``mod out by elements of $I$''. | |
\end{moral} | |
For example, modding out by $5\ZZ$ means that we consider | |
all elements in $\ZZ$ divisible by $5$ to be zero. | |
This gives you the usual modular arithmetic! | |
\begin{exercise} | |
Earlier, we wrote $\ZZ[i]$ for the Gaussian integers, | |
which was a slight abuse of notation. | |
Convince yourself that this ring | |
could instead be written as $\ZZ[x] / (x^2+1)$, | |
if we wanted to be perfectly formal. | |
(We will stick with $\ZZ[i]$ though --- it's more natural.) | |
Figure out the analogous formalization of $\ZZ[\sqrt[3]{2}]$. | |
\end{exercise} | |
\section{Generating ideals} | |
\prototype{In $\ZZ$, the ideals are all of the form $(n)$.} | |
Let's give you some practice with ideals. | |
An important piece of intuition is that once an ideal | |
contains a unit, it contains $1$, and | |
thus must contain the entire ring. | |
That's why the notion of ``proper ideal'' | |
is useful language. | |
To expand on that: | |
\begin{proposition} | |
[Proper ideal $\iff$ no units] | |
Let $R$ be a ring and $I \subseteq R$ an ideal. | |
Then $I$ is proper (i.e.\ $I \ne R$) | |
if and only if it contains no units of $R$. | |
\end{proposition} | |
\begin{proof} | |
Suppose $I$ contains a unit $u$, i.e.\ an element $u$ | |
with an inverse $u\inv$. | |
Then it contains $u \cdot u\inv = 1$, and thus $I = R$. | |
Conversely, if $I$ contains no units, it is obviously proper. | |
\end{proof} | |
As a consequence, if $K$ is a field, | |
then its only ideals are $(0)$ and $K$ | |
(this was \Cref{exer:field_ideal}). | |
So for our practice purposes, we'll be working with rings that aren't fields. | |
First practice: $\ZZ$. | |
\begin{exercise} | |
Show that the only ideals of $\ZZ$ are precisely those | |
sets of the form $n\ZZ$, where $n$ is a nonnegative integer. | |
\end{exercise} | |
Thus, while ideals of fields are not terribly interesting, | |
ideals of $\ZZ$ look eerily like elements of $\ZZ$. | |
Let's make this more precise. | |
\begin{definition} | |
Let $R$ be a ring. | |
The \vocab{ideal generated} by a set of elements $x_1, \dots, x_n \in R$ | |
is denoted by $I = (x_1, x_2, \dots, x_n)$ | |
and given by | |
\[ I = \left\{ r_1 x_1 + \dots + r_n x_n \mid r_i \in R \right\}. \] | |
One can think of this as ``the smallest ideal containing all the $x_i$''. | |
\end{definition} | |
The analogy of putting the $\{x_i\}$ in a sealed box and shaking vigorously | |
kind of works here too. | |
\begin{remark} | |
[Linear algebra digression] | |
If you know linear algebra, | |
you can summarize this as: an ideal is an $R$-module. | |
The ideal $(x_1, \dots, x_n)$ is the submodule spanned by $x_1, \dots, x_n$. | |
\end{remark} | |
In particular, if $I = (x)$ then $I$ consists of exactly the | |
``multiples of $x$'', i.e.\ numbers of the form $rx$ for $r \in R$. | |
\begin{remark} | |
We can also apply this definition to infinite generating sets, | |
as long as only finitely many of the $r_i$ are not zero | |
(since infinite sums don't make sense in general). | |
\end{remark} | |
\begin{example}[Examples of generated ideals] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii As $(n) = n\ZZ$ for all $n \in \ZZ$, | |
every ideal in $\ZZ$ is of the form $(n)$. | |
\ii In $\ZZ[i]$, we have | |
$(5) = \left\{ 5a + 5b i \mid a,b \in \ZZ \right\}$. | |
\ii In $\ZZ[x]$, the ideal $(x)$ consists of polynomials | |
with zero constant terms. | |
\ii In $\ZZ[x,y]$, the ideal $(x,y)$ again consists | |
of polynomials with zero constant terms. | |
\ii In $\ZZ[x]$, the ideal $(x,5)$ consists of polynomials | |
whose constant term is divisible by $5$. | |
\end{enumerate} | |
\end{example} | |
\begin{ques} | |
Please check that the set | |
$I = \left\{ r_1 x_1 + \dots + r_n x_n \mid r_i \in R \right\}$ | |
is indeed always an ideal (closed under addition, | |
and absorbs multiplication). | |
\end{ques} | |
Now suppose $I = (x_1, \dots, x_n)$. | |
What does $R/I$ look like? | |
According to what I said at the end of the last section, | |
it's what happens when we ``mod out'' by each of the elements $x_i$. | |
For example\dots | |
\begin{example} | |
[Modding out by generated ideals] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii Let $R = \ZZ$ and $I = (5)$. Then $R/I$ is literally | |
$\ZZ/5\ZZ$, or the ``integers modulo $5$'': | |
it is the result of declaring $5 = 0$. | |
\ii Let $R = \ZZ[x]$ and $I = (x)$. | |
Then $R/I$ means we send $x$ to zero; hence $R/I \cong \ZZ$ | |
as given any polynomial $p(x) \in R$, | |
we simply get its constant term. | |
\ii Let $R = \ZZ[x]$ again and now let $I = (x-3)$. | |
Then $R/I$ should be thought of as the quotient when $x-3 \equiv 0$, | |
that is, $x \equiv 3$. | |
So given a polynomial $p(x)$ its image after | |
we mod out should be thought of as $p(3)$. | |
Again $R/I \cong \ZZ$, but in a different way. | |
\ii Finally, let $I = (x-3,5)$. | |
Then $R/I$ not only sends $x$ to three, but also $5$ to zero. | |
So given $p \in R$, we get $p(3) \pmod 5$. | |
Then $R/I \cong \ZZ/5\ZZ$. | |
\end{enumerate} | |
\end{example} | |
\begin{remark} | |
[Mod notation] | |
By the way, given an ideal $I$ of a ring $R$, it's totally legit to write | |
\[ x \equiv y \pmod I \] | |
to mean that $x-y \in I$. | |
Everything you learned about modular arithmetic carries over. | |
\end{remark} | |
\section{Principal ideal domains} | |
\prototype{$\ZZ$ is a PID, $\ZZ[x]$ is not. | |
$\CC[x]$ is a PID, $\CC[x,y]$ is not.} | |
What happens if we put multiple generators in an ideal, | |
like $(10,15) \subseteq \ZZ$? | |
Well, we have by definition that $(10,15)$ is given as a set by | |
\[ (10,15) \defeq \left\{ 10x + 15y \mid x,y \in \ZZ \right\}. \] | |
If you're good at number theory you'll instantly | |
recognize this as $5\ZZ = (5)$. | |
Surprise! In $\ZZ$, the ideal $(a,b)$ is exactly $\gcd(a,b) \ZZ$. | |
And that's exactly the reason you often see the GCD of two numbers denoted $(a,b)$. | |
We call such an ideal (one generated by a single element) a \vocab{principal ideal}. | |
So, in $\ZZ$, every ideal is principal. | |
But the same is not true in more general rings. | |
\begin{example} | |
[A non-principal ideal] | |
In $\ZZ[x]$, $I = (x,2015)$ is \emph{not} a principal ideal. | |
For if $I = (f)$ for some polynomial $f \in I$ | |
then $f$ divides $x$ and $2015$. | |
This can only occur if $f = \pm 1$, | |
but then $I$ contains $\pm1$, which it does not. | |
\end{example} | |
A ring with the property that all its ideals | |
are principal is called a \vocab{principal ideal ring}. | |
We like this property because they effectively | |
let us take the ``greatest common factor'' | |
in a similar way as the GCD in $\ZZ$. | |
In practice, we actually usually care about | |
so-called \textbf{principal ideal domains (PID's)}. | |
But we haven't defined what a domain is yet. | |
Nonetheless, all the examples below are actually PID's, | |
so we will go ahead and use this word for now, | |
and tell you what the additional condition is in the next chapter. | |
\begin{example} | |
[Examples of PID's] | |
To reiterate, for now you should just verify | |
that these are principal ideal rings, | |
even though we are using the word PID. | |
\begin{enumerate}[(a)] | |
\ii As we saw, $\ZZ$ is a PID. | |
\ii As we also saw, $\ZZ[x]$ is not a PID, | |
since $I = (x,2015)$ for example is not principal. | |
\ii It turns out that for a field $k$ | |
the ring $k[x]$ is always a PID. | |
For example, $\QQ[x]$, $\RR[x]$, $\CC[x]$ are PID's. | |
If you want to try and prove this, | |
first prove an analog of B\'{e}zout's lemma, | |
which implies the result. | |
\ii $\CC[x,y]$ is not a PID, because $(x,y)$ | |
is not principal. | |
\end{enumerate} | |
\end{example} | |
\section{Noetherian rings} | |
\prototype{$\ZZ[x_1, x_2, \dots]$ is not Noetherian, | |
but most reasonable rings are. | |
In particular polynomial rings are. | |
(Equivalently, only weirdos care about non-Noetherian rings).} | |
If it's too much to ask that an ideal is generated by \emph{one} element, | |
perhaps we can at least ask that our ideals | |
are generated by \emph{finitely many} elements. | |
Unfortunately, in certain weird rings this is also not the case. | |
\begin{example} | |
[Non-Noetherian ring] | |
Consider the ring $R = \ZZ[x_1, x_2, x_3, \dots]$ | |
which has \emph{infinitely} many free variables. | |
Then the ideal $I = (x_1, x_2, \dots) \subseteq R$ | |
cannot be written with a finite generating set. | |
\end{example} | |
Nonetheless, most ``sane'' rings we work in | |
\emph{do} have the property that their ideals are finitely generated. | |
We now name such rings and give two equivalent definitions: | |
\begin{proposition}[The equivalent definitions of a Noetherian ring] | |
For a ring $R$, the following are equivalent: | |
\begin{enumerate}[(a)] | |
\ii Every ideal $I$ of $R$ is finitely generated | |
(i.e.\ can be written with a finite generating set). | |
\ii There does \emph{not} exist an | |
infinite ascending chain of ideals | |
\[ I_1 \subsetneq I_2 \subsetneq I_3 \subsetneq \dots. \] | |
The absence of such chains is often | |
called the \vocab{ascending chain condition}. | |
\end{enumerate} | |
Such rings are called \vocab{Noetherian}. | |
\end{proposition} | |
\begin{example} | |
[Non-Noetherian ring breaks ACC] | |
In the ring $R = \ZZ[x_1, x_2, x_3, \dots]$ we have | |
an infinite ascending chain | |
\[ (x_1) \subsetneq (x_1, x_2) \subsetneq (x_1,x_2,x_3) \subsetneq \dots. \] | |
\end{example} | |
From the example, you can kind of see why the proposition is true: | |
from an infinitely generated ideal you can extract an ascending chain | |
by throwing elements in one at a time. | |
I'll leave the proof to you if you want to | |
do it.\footnote{On the other hand, every undergraduate | |
class in this topic I've seen makes you do it as homework. | |
Admittedly I haven't gone to that many such classes.} | |
\begin{ques} | |
Why are fields Noetherian? | |
Why are PID's (such as $\ZZ$) Noetherian? | |
\end{ques} | |
This leaves the question: | |
is our prototypical non-example of a PID, | |
$\ZZ[x]$, a Noetherian ring? | |
The answer is a glorious yes, | |
according to the celebrated Hilbert basis theorem. | |
\begin{theorem}[Hilbert basis theorem] | |
Given a Noetherian ring $R$, | |
the ring $R[x]$ is also Noetherian. | |
Thus by induction, $R[x_1, x_2, \dots, x_n]$ is Noetherian | |
for any integer $n$. | |
\label{thm:hilbert_basis} | |
\end{theorem} | |
The proof of this theorem is really olympiad flavored, | |
so I couldn't possibly spoil it -- I've | |
left it as a problem at the end of this chapter. | |
Noetherian rings really shine in algebraic geometry, | |
and it's a bit hard for me to motivate them right now, | |
other than to say | |
``most rings you'll encounter are Noetherian''. | |
Please bear with me! | |
\section{\problemhead} | |
\begin{problem} | |
The ring $R = \RR[x] / (x^2+1)$ is one that you've seen before. | |
What is its name? | |
\begin{hint} | |
$R = \RR[i]$. | |
\end{hint} | |
\begin{sol} | |
This is just $\RR[i] = \CC$. | |
The isomorphism is given by $x \mapsto i$, | |
which has kernel $(x^2+1)$. | |
\end{sol} | |
\end{problem} | |
\begin{problem} | |
Show that $\CC[x] / (x^2-x) \cong \CC \times \CC$. | |
\begin{hint} | |
The isomorphism is given by $x \mapsto (1,0)$ | |
and $1-x \mapsto (0,1)$. | |
\end{hint} | |
\end{problem} | |
\begin{problem} | |
In the ring $\ZZ$, let $I = (2016)$ and $J = (30)$. | |
Show that $I \cap J$ is an ideal of $\ZZ$ and compute its elements. | |
\end{problem} | |
\begin{sproblem} | |
\label{prob:inclusion_preserving} | |
Let $R$ be a ring and $I$ an ideal. | |
Find an inclusion-preserving bijection between | |
\begin{itemize} | |
\ii ideals of $R/I$, and | |
\ii ideals of $R$ which contain $I$. | |
\end{itemize} | |
\end{sproblem} | |
\begin{problem} | |
Let $R$ be a ring. | |
\begin{enumerate}[(a)] | |
\ii Prove that there is exactly one ring homomorphism $\ZZ \to R$. | |
\ii Prove that the number of ring homomorphisms | |
$\ZZ[x] \to R$ is equal to the number of elements of $R$. | |
\end{enumerate} | |
\begin{hint} | |
For (b) homomorphism is uniquely determined by the choice of $\psi(x) \in R$ | |
\end{hint} | |
\end{problem} | |
%\begin{problem} | |
% [$\phi(1_R) = 1_S$ is really necessary] | |
% Find two rings $R$ and $S$ and a \emph{nonzero} function | |
% $f \colon R \to S$ such that | |
% \begin{itemize} | |
% \ii $f(x+y) = f(x) + f(y)$ for $x,y \in R$, | |
% \ii $f(xy) = f(x) f(y)$ for $x,y \in R$, | |
% \ii $f(1_R) \ne 1_S$ (i.e.\ $f$ is not a ring homomorphism). | |
% \end{itemize} | |
% \begin{hint} | |
% Take $S = R \times R$ with $R$ nonzero. | |
% \end{hint} | |
% \begin{sol} | |
% The map $R \to R \times R$ by $x \mapsto (x,0)$ will always work, | |
% with $R$ nonzero. | |
% \end{sol} | |
%\end{problem} | |
\begin{problem} | |
\gim | |
Prove the Hilbert basis theorem, \Cref{thm:hilbert_basis}. | |
\end{problem} | |
\begin{problem} | |
[USA Team Selection Test 2016] | |
Let $\FF_p$ denote the integers modulo a fixed prime number $p$. | |
Define $\Psi \colon \FF_p[x] \to \FF_p[x]$ by | |
\[ \Psi\left( \sum_{i=0}^n a_i x^i \right) = \sum_{i=0}^n a_i x^{p^i}. \] | |
Let $S$ denote the image of $\Psi$. | |
\begin{enumerate}[(a)] | |
\ii Show that $S$ is a ring with addition | |
given by polynomial addition, | |
and multiplication given by \emph{function composition}. | |
\ii Prove that $\Psi \colon \FF_p[x] \to S$ | |
is then a ring isomorphism. | |
\end{enumerate} | |
\end{problem} | |
\begin{problem} % from Brian Chen | |
\yod | |
Let $A \subseteq B \subseteq C$ be rings. | |
Suppose $C$ is a finitely generated $A$-module. | |
Does it follow that $B$ is a finitely generated $A$-module? | |
% Assume $A$ is Noetherian. Show that $B$ is finitely generated as an $A$-module. | |
% Find a counterexample where $A$ is not Noetherian. | |
\begin{hint} | |
I think the result is true if you add the assumption $A$ is Noetherian, | |
so look for trouble by picking $A$ not Noetherian. | |
\end{hint} | |
\begin{sol} | |
Nope! Pick | |
\begin{align*} | |
A &= \ZZ[x_1, x_2, \dots] \\ | |
B &= \ZZ[x_1, x_2, \dots, \eps x_1, \eps x_2, \dots] \\ | |
C &= \ZZ[x_1, x_2, \dots, \eps]. | |
\end{align*} | |
where $\eps \neq 0$ but $\eps^2 = 0$. | |
I think the result is true if you add the assumption $A$ is Noetherian. | |
\end{sol} | |
\end{problem} | |