Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Minkowski bound and class groups} | |
We now have a neat theory of unique factorization of ideals. | |
In the case of a PID, this in fact gives us a UFD. Sweet. | |
We'll define, in a moment, something called the \emph{class group} | |
which measures how far $\OO_K$ is from being a PID; | |
the bigger the class group, the farther $\OO_K$ is from being a PID. | |
In particular, the $\OO_K$ is a PID if it has trivial class group. | |
Then we will provide some inequalities which let us put restrictions on the class group; | |
for instance, this will let us show in some cases that the class group must be trivial. | |
Astonishingly, the proof will use Minkowski's theorem, a result from geometry. | |
\section{The class group} | |
\prototype{PID's have trivial class group.} | |
Let $K$ be a number field, | |
and let $J_K$ denote the multiplicative group of fractional ideals of $\OO_K$. | |
Let $P_K$ denote the multiplicative group of \vocab{principal fractional ideals}: | |
those of the form $(x) = x \OO_K$ for some $x \in K$. | |
\begin{ques} | |
Check that $P_K$ is also a multiplicative group. | |
(This is really easy: name $x\OO_K \cdot y\OO_K$ and $(x\OO_K)\inv$.) | |
\end{ques} | |
As $J_K$ is abelian, we can now define the \vocab{class group} to be the quotient | |
\[ \Cl_K \defeq J_K / P_K. \] | |
The elements of $\Cl_K$ are called \vocab{classes}. | |
Equivalently, | |
\begin{moral} | |
The class group $\Cl_K$ is the set of nonzero fractional ideals modulo | |
scaling by a constant in $K$. | |
\end{moral} | |
In particular, $\Cl_K$ is trivial if all ideals are principal, | |
since the nonzero principal ideals are the same up to scaling. | |
The size of the class group is called the \vocab{class number}. | |
It's a beautiful theorem that the class number is always finite, | |
and the bulk of this chapter will build up to this result. | |
It requires several ingredients. | |
\section{The discriminant of a number field} | |
\prototype{Quadratic fields.} | |
Let's say I have $K = \QQ(\sqrt 2)$. | |
As we've seen before, this means $\OO_K = \ZZ[\sqrt 2]$, meaning | |
\[ \OO_K = \left\{ a + b \sqrt 2 \mid a,b \in \ZZ \right\}. \] | |
The key insight now is that you might think of this as a \emph{lattice}: | |
geometrically, we want to think about this the same way we think about $\ZZ^2$. | |
Perversely, we might try to embed this into $\QQ^2$ by sending $a+b\sqrt 2$ to $(a, b)$. | |
But this is a little stupid, since we're rudely making $K$, | |
which somehow lives inside $\RR$ and is | |
``one-dimensional'' in that sense, into a two-dimensional space. | |
It also depends on a choice of basis, which we don't like. | |
A better way is to think about the fact that there are two embeddings | |
$\sigma_1 : K \to \CC$ and $\sigma_2 : K \to \CC$, namely the identity, and conjugation: | |
\begin{align*} | |
\sigma_1(a+b\sqrt2) &= a+b\sqrt 2 \\ | |
\sigma_2(a+b\sqrt2) &= a-b\sqrt 2. | |
\end{align*} | |
Fortunately for us, these embeddings both have real image. | |
This leads us to consider the set of points | |
\[ \left( \sigma_1(\alpha), \sigma_2(\alpha) \right) \in \RR^2 | |
\quad\text{as}\quad \alpha \in K. \] | |
This lets us visualize what $\OO_K$ looks like in $\RR^2$. | |
The points of $K$ are dense in $\RR^2$, but the points of $\OO_K$ cut out a lattice. | |
\begin{center} | |
\begin{asy} | |
size(8cm); | |
real T = 1.414; | |
int N = 10; | |
real M = 5; | |
for (int a = -N; a <= N; ++a) { | |
for (int b = -N; b <= N; ++b) { | |
if ((abs(a+T*b) <= M) && (abs(a-T*b) <= M)) | |
dot( (a+T*b, a-T*b) ); | |
} | |
} | |
draw( (-M,0)--(M,0), dotted); | |
draw( (0,-M)--(0,M), dotted); | |
label("$-2$", (-2,-2), dir(135)); | |
label("$-1$", (-1,-1), dir(135)); | |
label("$0$", (0,0), dir(135)); | |
label("$1$", (1,1), dir(135)); | |
label("$2$", (2,2), dir(135)); | |
label("$-\sqrt 2$", (-T,T), dir(135)); | |
label("$\sqrt 2$", (T,-T), dir(-45)); | |
label("$1+\sqrt 2$", (1+T,1-T), dir(-45)); | |
filldraw((0,0)--(1,1)--(1+T,1-T)--(T,-T)--cycle, heavycyan, blue); | |
\end{asy} | |
\end{center} | |
To see how big the lattice is, we look at how $\{1, \sqrt2\}$, the generators | |
of $\OO_K$, behave. | |
The point corresponding to $a+b\sqrt2$ in the lattice is | |
\[ a \cdot (1,1) + b \cdot (\sqrt 2, -\sqrt 2). \] | |
The \vocab{mesh} of the lattice\footnote{Most authors call this the volume, but I | |
think this is not the right word to use -- lattices have ``volume'' zero since they | |
are just a bunch of points! In contrast, the English word ``mesh'' really | |
does refer to the width of a ``gap''.} | |
is defined as the hypervolume of the ``fundamental parallelepiped'' I've colored blue above. | |
For this particular case, it ought to be equal to the | |
area of that parallelogram, which is | |
\[ | |
\det | |
\begin{bmatrix} | |
1 & -\sqrt 2 \\ | |
1 & \sqrt 2 | |
\end{bmatrix} | |
= 2\sqrt 2. | |
\] | |
The definition of the discriminant is precisely this, except with an extra square factor | |
(since permutation of rows could lead to changes in sign in the matrix above). | |
\Cref{prob:trace_discriminant} shows that the squaring makes $\Delta_K$ an integer. | |
To make the next definition, we invoke: | |
\begin{theorem}[The $n$ embeddings of a number field] | |
Let $K$ be a number field of degree $n$. | |
Then there are exactly $n$ field homomorphisms $K \injto \CC$, | |
say $\sigma_1, \dots, \sigma_n$, which fix $\QQ$. | |
\end{theorem} | |
\begin{proof} | |
Deferred to \Cref{thm:n_embeddings}, once we have the tools of Galois theory. | |
\end{proof} | |
In fact, in \Cref{thm:conj_distrb} we see that | |
for $\alpha \in K$, we have that $\sigma_i(\alpha)$ | |
runs over the conjugates of $\alpha$ as $i=1,\dots,n$. | |
It follows that | |
\[ | |
\Tr_{K/\QQ}(\alpha) = \sum_{i=1}^n \sigma_i(\alpha) | |
\quad\text{and}\quad | |
\Norm_{K/\QQ}(\alpha) = \prod_{i=1}^n \sigma_i(\alpha). | |
\] | |
This allows us to define: | |
\begin{definition} | |
Suppose $\alpha_1, \dots, \alpha_n$ is a $\ZZ$-basis of $\OO_K$. | |
The \vocab{discriminant} of the number field $K$ is defined by | |
\[ | |
\Delta_K \defeq \det | |
\begin{bmatrix} | |
\sigma_1(\alpha_1) & \dots & \sigma_n(\alpha_1) \\ | |
\vdots & \ddots & \vdots \\ | |
\sigma_1(\alpha_n) & \dots & \sigma_n(\alpha_n) \\ | |
\end{bmatrix}^2. | |
\] | |
\end{definition} | |
This does not depend on the choice of the $\{\alpha_i\}$; | |
we will not prove this here. | |
\begin{example}[Discriminant of $K = \QQ(\sqrt2)$] | |
We have $\OO_K = \ZZ[\sqrt 2]$ | |
and as discussed above the discriminant is | |
\[ | |
\Delta_K = | |
(-2 \sqrt 2)^2 = 8. | |
\] | |
\end{example} | |
\begin{example}[Discriminant of $\QQ(i)$] | |
Let $K = \QQ(i)$. | |
We have $\OO_K = \ZZ[i] = \ZZ \oplus i \ZZ$. | |
The embeddings are the identity and complex conjugation | |
which take $1$ to $(1,1)$ and $i$ to $(i, -i)$. | |
So | |
\[ | |
\Delta_K = | |
\det | |
\begin{bmatrix} | |
1 & 1 \\ | |
i & -i | |
\end{bmatrix}^2 | |
= | |
(-2i)^2 = -4. | |
\] | |
This example illustrates that the discriminant need not be positive | |
for number fields which wander into the complex plane | |
(the lattice picture is a less perfect analogy). | |
But again, as we'll prove in the problems the discriminant is always | |
an integer. | |
\end{example} | |
\begin{example} | |
[Discriminant of $\QQ(\sqrt 5)$] | |
Let $K = \QQ(\sqrt5)$. | |
This time, $\OO_K = \ZZ \oplus \frac{1+\sqrt5}{2} \ZZ$, and so the discriminant | |
is going to look a little bit different. | |
The embeddings are still $a+b\sqrt 5 \mapsto a+b\sqrt5, a-b\sqrt5$. | |
Applying this to the $\ZZ$-basis $\left\{ 1, \frac{1+\sqrt5}{2} \right\}$, we get | |
\[ | |
\Delta_K | |
= | |
\det | |
\begin{bmatrix} | |
1 & 1 \\ | |
\frac{1+\sqrt5}{2} & \frac{1-\sqrt5}{2} | |
\end{bmatrix}^2 | |
= (-\sqrt 5)^2 = 5. | |
\] | |
\end{example} | |
\begin{exercise} | |
Extend all this to show that | |
if $K = \QQ(\sqrt d)$ for $d \neq 1$ squarefree, we have | |
\[ | |
\Delta_K = | |
\begin{cases} | |
d & \text{if } d \equiv 1 \pmod 4 \\ | |
4d & \text{if } d \equiv 2, 3 \pmod 4. | |
\end{cases} | |
\] | |
\end{exercise} | |
Actually, let me point out something curious: recall that the polynomial discriminant of $Ax^2+Bx+C$ is $B^2-4AC$. Then: | |
\begin{itemize} | |
\ii In the $d \equiv 1 \pmod 4$ case, | |
$\Delta_K$ is the discriminant of $x^2 - x - \frac{d-1}{4}$, | |
which is the minimal polynomial of $\half(1+\sqrt d)$. | |
Of course, $\OO_K = \ZZ[\half(1+\sqrt d)]$. | |
\ii In the $d \equiv 2,3 \pmod 4$ case, | |
$\Delta_K$ is the discriminant of $x^2 - d$ | |
which is the minimal polynomial of $\sqrt d$. Once again, $\OO_K = \ZZ[\sqrt d]$. | |
\end{itemize} | |
This is not a coincidence! \Cref{prob:root_discriminant} asserts that this is true in general; | |
hence the name ``discriminant''. | |
\section{The signature of a number field} | |
\prototype{$\QQ(\sqrt[100]{2})$ has signature $(2, 49)$.} | |
In the example of $K = \QQ(i)$, | |
we more or less embedded $K$ into the space $\CC$. | |
However, $K$ is a degree two extension, | |
so what we'd really like to do is embed it into $\RR^2$. | |
To do so, we're going to take advantage of complex conjugation. | |
Let $K$ be a number field and $\sigma_1, \dots, \sigma_n$ be its embeddings. | |
We distinguish between the \vocab{real embeddings} | |
(which map all of $K$ into $\RR$) | |
and the \vocab{complex embeddings} | |
(which map some part of $K$ outside $\RR$). | |
Notice that if $\sigma$ is a complex embedding, | |
then so is the conjugate $\ol{\sigma} \ne \sigma$; | |
hence complex embeddings come in pairs. | |
\begin{definition} | |
Let $K$ be a number field of degree $n$, and set | |
\begin{align*} | |
r_1 &= \text{number of real embeddings} \\ | |
r_2 &= \text{number of pairs of complex embeddings}. | |
\end{align*} | |
The \vocab{signature} of $K$ is the pair $(r_1, r_2)$. | |
Observe that $r_1 + 2r_2 = n$. | |
\end{definition} | |
\begin{example}[Basic examples of signatures] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii $\QQ$ has signature $(1,0)$. | |
\ii $\QQ(\sqrt 2)$ has signature $(2,0)$. | |
\ii $\QQ(i)$ has signature $(0,1)$. | |
\ii Let $K = \QQ(\sqrt[3]{2})$, and let $\omega$ be a cube root of unity. | |
The elements of $K$ are | |
\[ K = \left\{ a + b\cbrt2 + c\cbrt4 \mid a,b,c \in \QQ \right\}. \] | |
Then the signature is $(1,1)$, because the three embeddings are | |
\[ \sigma_1 : \cbrt 2 \mapsto \cbrt 2, | |
\quad | |
\sigma_2 : \cbrt 2 \mapsto \cbrt 2 \omega, | |
\quad | |
\sigma_3 : \cbrt 2 \mapsto \cbrt 2 \omega^2. \] | |
The first of these is real and the latter two are conjugate pairs. | |
\end{enumerate} | |
\end{example} | |
\begin{example}[Even more signatures] | |
In the same vein $\QQ(\sqrt[99]{2})$ and $\QQ(\sqrt[100]{2})$ | |
have signatures $(1,49)$ and $(2,49)$. | |
\end{example} | |
\begin{ques} | |
Verify the signatures of the above two number fields. | |
\end{ques} | |
From now on, we will number the embeddings of $K$ in such a way that | |
\[ \sigma_1, \sigma_2, \dots, \sigma_{r_1} \]are the real embeddings, | |
while | |
\[ | |
\sigma_{r_1+1} = \ol{\sigma_{r_1+r_2+1}}, \quad | |
\sigma_{r_1+2} = \ol{\sigma_{r_1+r_2+2}}, \quad | |
\dots, \quad | |
\sigma_{r_1+r_2} = \ol{\sigma_{r_1+2r_2}}. | |
\] | |
are the $r_2$ pairs of complex embeddings. | |
We define the \vocab{canonical embedding} of $K$ as | |
\[ | |
K \overset\iota\injto \RR^{r_1} \times \CC^{r_2} \ | |
\quad\text{by}\quad | |
\alpha \mapsto \left( \sigma_1(\alpha), \dots, \sigma_{r_1}(\alpha), | |
\sigma_{r_1+1}(\alpha), \dots, \sigma_{r_1+r_2}(\alpha) \right). | |
\] | |
All we've done is omit, for the complex case, the second of the embeddings in each conjugate pair. | |
This is no big deal, since they are just conjugates; | |
the above tuple is all the information we need. | |
For reasons that will become obvious in a moment, I'll let $\tau$ denote the isomorphism | |
\[ | |
\tau : \RR^{r_1} \times \CC^{r_2} \taking{\sim} \RR^{r_1+2r_2} = \RR^n | |
\] | |
by breaking each complex number into its real and imaginary part, as | |
\begin{align*} | |
\alpha \mapsto \big(& \sigma_1(\alpha), \dots, \sigma_{r_1}(\alpha), \\ | |
& \Re \sigma_{r_1+1}(\alpha), \;\; \Im \sigma_{r_1+1}(\alpha), \\ | |
& \Re \sigma_{r_1+2}(\alpha), \;\; \Im \sigma_{r_1+2}(\alpha), \\ | |
& \dots, \\ | |
& \Re \sigma_{r_1+r_2}(\alpha), \;\; \Im \sigma_{r_1+r_2}(\alpha) \big). | |
\end{align*} | |
\begin{example}[Example of canonical embedding] | |
As before let $K = \QQ(\sqrt[3]{2})$ and set | |
\[ \sigma_1 : \cbrt 2 \mapsto \cbrt 2, | |
\quad | |
\sigma_2 : \cbrt 2 \mapsto \cbrt 2 \omega, | |
\quad | |
\sigma_3 : \cbrt 2 \mapsto \cbrt 2 \omega^2 \] | |
where $\omega = - \half + \frac{\sqrt3}{2} i$, | |
noting that we've already arranged indices so $\sigma_1 = \id$ is real | |
while $\sigma_2$ and $\sigma_3$ are a conjugate pair. | |
So the embeddings $K \overset\iota\injto \RR \times \CC \taking\sim \RR^3$ are given by | |
\[ | |
\alpha \overset\iota\longmapsto \left( \sigma_1(\alpha), \sigma_2(\alpha) \right) | |
\overset\tau\longmapsto \left( \sigma_1(\alpha), \; \Re\sigma_2(\alpha), \; \Im\sigma_2(\alpha) \right). \] | |
For concreteness, taking $\alpha = 9 + \cbrt 2$ gives | |
\begin{align*} | |
9 + \cbrt 2 &\overset\iota\mapsto | |
\left( 9 + \cbrt 2, \;\; 9 + \cbrt2\omega \right) \\ | |
&= \left( 9 + \cbrt 2, \;\; | |
9 - \half\cbrt2 + \frac{\sqrt[6]{108}}{2} i \right) \in \RR \times \CC \\ | |
&\overset\tau\mapsto \left( 9 + \cbrt 2, \;\; 9 - \half \cbrt 2, | |
\;\; \frac{\sqrt[6]{108}}{2} \right) \in \RR^3. | |
\end{align*} | |
\end{example} | |
Now, the whole point of this is that we want to consider the resulting lattice | |
when we take $\OO_K$. In fact, we have: | |
\begin{lemma} | |
\label{lem:vol_OK_mesh} | |
Consider the composition of the embeddings $K \injto \RR^{r_1} \times \CC^{r_2} \taking\sim \RR^n$. | |
Then as before, $\OO_K$ becomes a lattice $L$ in $\RR^n$, with mesh equal to | |
\[ \frac{1}{2^{r_2}} \sqrt{\left\lvert \Delta_K \right\rvert}. \] | |
\end{lemma} | |
\begin{proof} | |
Fun linear algebra problem (you just need to manipulate determinants). | |
Left as \Cref{prob:OK_linalg}. | |
\end{proof} | |
From this we can deduce: | |
\begin{lemma} | |
Consider the composition of the embeddings $K \injto \RR^{r_1} \times \CC^{r_2} \taking\sim \RR^n$. | |
Let $\ka$ be an ideal in $\OO_K$. | |
Then the image of $\ka$ is a lattice $L_\ka$ in $\RR^n$ with mesh equal to | |
\[ \frac{\Norm(\ka)}{2^{r_2}} \sqrt{\left\lvert \Delta_K \right\rvert}. \] | |
\end{lemma} | |
\begin{proof}[Sketch of Proof] | |
Let \[ d = \Norm(\ka) \defeq [\OO_K : \ka]. \] | |
Then in the lattice $L_\ka$, we somehow only take $\frac 1d$th of the points | |
which appear in the lattice $L$, which is why the area increases by a factor of $\Norm(\ka)$. | |
To make this all precise I would need to do a lot more with lattices and geometry | |
than I have space for in this chapter, so I will omit the details. | |
But I hope you can see why this is intuitively true. | |
\end{proof} | |
\section{Minkowski's theorem} | |
Now I can tell you why I insisted we move from $\RR^{r_1} \times \CC^{r_2}$ to $\RR^n$. | |
In geometry, there's a really cool theorem of Minkowski's that goes as follows. | |
\begin{theorem} | |
[Minkowski] | |
Let $S \subseteq \RR^n$ be a convex set containing $0$ | |
which is centrally symmetric (meaning that $x \in S \iff -x \in S$). | |
Let $L$ be a lattice with mesh $d$. | |
If either | |
\begin{enumerate}[(a)] | |
\ii The volume of $S$ exceeds $2^n d$, or | |
\ii The volume of $S$ equals $2^n d$ and $S$ is compact, | |
\end{enumerate} | |
then $S$ contains a nonzero lattice point of $L$. | |
\end{theorem} | |
\begin{ques} | |
Show that the condition $0 \in S$ is actually extraneous | |
in the sense that any nonempty, convex, centrally symmetric set contains the origin. | |
\end{ques} | |
\begin{proof}[Sketch of Proof] | |
Part (a) is surprisingly simple and has a very olympiad-esque solution: | |
it's basically Pigeonhole on areas. | |
We'll prove part (a) in the special case $n=2$, | |
$L = \ZZ^2$ for simplicity as the proof | |
can easily be generalized to any lattice and any $n$. | |
Thus we want to show that any such convex set $S$ | |
with area more than $4$ contains a lattice point. | |
Dissect the plane into $2 \times 2$ squares | |
\[ [2a-1, 2a+1] \times [2b-1, 2b+1] \] | |
and overlay all these squares on top of each other. | |
By the Pigeonhole Principle, we find there exist two points $p \neq q \in S$ which map to the same point. | |
Since $S$ is symmetric, $-q \in S$. Then $\half (p-q) \in S$ (convexity) and is a nonzero lattice point. | |
I'll briefly sketch part (b): the idea is to consider $(1+\eps) S$ for $\eps > 0$ | |
(this is ``$S$ magnified by a small factor $1+\eps$''). | |
This satisfies condition (a). So for each $\eps > 0$ the set of nonzero lattice points in $(1+\eps) S$, | |
say $S_\eps$, is a \emph{finite nonempty set} of (discrete) points | |
(the ``finite'' part follows from the fact that $(1+\eps)S$ is bounded). | |
So there has to be some point that's in $S_\eps$ for every $\eps > 0$ (why?), which implies it's in $S$. | |
\end{proof} | |
\section{The trap box} | |
The last ingredient we need is a set to apply Minkowski's theorem to. I propose: | |
\begin{definition} | |
Let $M$ be a positive real. | |
In $\RR^{r_1} \times \CC^{r_2}$, define the box $S$ to be the | |
set of points $(x_1, \dots, x_{r_1}, z_1, \dots, z_{r_2})$ such that | |
\[ | |
\sum_{i=1}^{r_1} \left\lvert x_i \right\rvert | |
+ 2 \sum_{j=1}^{r_2} \left\lvert z_j \right\rvert | |
\le M. | |
\] | |
Note that this depends on the value of $M$. | |
\end{definition} | |
Think of this box as a \emph{mousetrap}: anything that falls in it is going to have a small norm, | |
and our goal is to use Minkowski to lure some nonzero element into it. | |
\begin{center} | |
\begin{asy} | |
size(6cm); | |
real T = 1.414; | |
int N = 10; | |
real M = 3; | |
real K = 2.4; | |
filldraw( (K,0)--(0,K)--(-K,0)--(0,-K)--cycle, palered, darkred); | |
label("$S$", (2*K/3,K/3), dir(45)); | |
for (int a = -N; a <= N; ++a) { | |
for (int b = -N; b <= N; ++b) { | |
if ((abs(a+T*b) <= M) && (abs(a-T*b) <= M)) | |
dot( (a+T*b, a-T*b) ); | |
} | |
} | |
draw( (-M,0)--(M,0), dotted); | |
draw( (0,-M)--(0,M), dotted); | |
label("$-2$", (-2,-2), dir(135)); | |
label("$-1$", (-1,-1), dir(135)); | |
label("$0$", (0,0), dir(135)); | |
label("$1$", (1,1), dir(135)); | |
label("$2$", (2,2), dir(135)); | |
label("$-\sqrt 2$", (-T,T), dir(135)); | |
label("$\sqrt 2$", (T,-T), dir(-45)); | |
label("$1+\sqrt 2$", (1+T,1-T), dir(-45)); | |
\end{asy} | |
\end{center} | |
That is, suppose $\alpha \in \ka$ falls into the box I've defined above, which means | |
\[ | |
M \ge | |
\sum_{i=1}^{r_1} \left\lvert \sigma_i(\alpha) \right\rvert | |
+ 2 \sum_{i=r_1+1}^{r_1+r_2} \left\lvert \sigma_i(\alpha) \right\rvert | |
= \sum_{i=1}^{n} \left\lvert \sigma_i(\alpha) \right\rvert, | |
\] | |
where we are remembering that the last few $\sigma$'s come in conjugate pairs. | |
This looks like the trace, but the absolute values are in the way. | |
So instead, we apply AM-GM to obtain: | |
\begin{lemma}[Effect of the mousetrap] | |
Let $\alpha \in \OO_K$, and suppose $\iota(\alpha)$ is in $S$ | |
(where $\iota : K \injto \RR^{r_1} \times \CC^{r_2}$ as usual). | |
Then | |
\[ \Norm_{K/\QQ}(\alpha) | |
= \prod_{i=1}^n \left\lvert \sigma_i(\alpha) \right\rvert | |
\le \left( \frac Mn \right)^n. \] | |
\end{lemma} | |
The last step we need to do is compute the volume of the box. | |
This is again some geometry I won't do, but take my word for it: | |
\begin{lemma}[Size of the mousetrap] | |
Let $\tau : \RR^{r_1} \times \CC^{r_2} \taking\sim \RR^n$ as before. | |
Then the image of $S$ under $\tau$ is a convex, compact, centrally symmetric set with volume | |
\[ 2^{r_1} \cdot \left( \frac{\pi}{2} \right)^{r_2} \cdot \frac{M^n}{n!}. \] | |
\end{lemma} | |
\begin{ques} | |
(Sanity check) | |
Verify that the above is correct for the signatures $(r_1, r_2) = (2,0)$ and $(r_1,r_2) = (0,1)$, | |
which are the possible signatures when $n=2$. | |
\end{ques} | |
\section{The Minkowski bound} | |
We can now put everything we have together to obtain the great Minkowski bound. | |
\begin{theorem} | |
[Minkowski bound] | |
Let $\ka \subseteq \OO_K$ be any nonzero ideal. | |
Then there exists $0 \neq \alpha \in \ka$ such that | |
\[ \Norm_{K/\QQ}(\alpha) \le \left( \frac 4\pi \right)^{r_2} \frac{n!}{n^n} \sqrt{\left\lvert \Delta_K \right\rvert} | |
\cdot \Norm(\ka). \] | |
\end{theorem} | |
\begin{proof} | |
This is a matter of putting all our ingredients together. | |
Let's see what things we've defined already: | |
\begin{center} | |
\begin{tikzcd} | |
K \ar[r, "\iota", hook] | |
& \RR^{r_1} \times \CC^{r_2} \ar[r, "\tau"] | |
& \RR^n \\ | |
& \text{box } S \ar[r, mapsto] | |
& \tau\im(S) | |
& \quad\text{with volume } | |
2^{r_1} \left( \frac\pi2 \right)^{r_2} \frac{M^n}{n!} \\ | |
\OO_K \ar[rr, mapsto] | |
&& \text{Lattice } L | |
& \quad\text{with mesh } | |
2^{-r_2} \sqrt{\left\lvert \Delta_K \right\rvert} \\ | |
\ka \ar[rr, mapsto] | |
&& \text{Lattice } L_\ka | |
& \quad\text{with mesh } | |
2^{-r_2} \sqrt{\left\lvert \Delta_K \right\rvert} \Norm(\ka) | |
\end{tikzcd} | |
\end{center} | |
Pick a value of $M$ such that the mesh of $L_\ka$ | |
equals $2^{-n}$ of the volume of the box. | |
Then Minkowski's theorem gives that some $0 \neq \alpha \in \ka$ lands inside the box --- | |
the mousetrap is configured to force $\Norm_{K/\QQ}(\alpha) \le \frac{1}{n^n} M^n$. | |
The correct choice of $M$ is | |
\[ | |
M^n | |
= M^n \cdot 2^n \cdot \frac{\text{mesh}}{\text{vol box}} | |
= 2^n \cdot \frac{n!}{2^{r_1} \cdot \left( \frac \pi 2 \right)^{r_2}} | |
\cdot 2^{-r_2} \sqrt{\left\lvert \Delta_K \right\rvert} \Norm(\ka) | |
\] | |
which gives the bound after some arithmetic. | |
\end{proof} | |
\section{The class group is finite} | |
\begin{definition} | |
Let | |
$M_K = \left( \frac 4\pi \right)^{r_2} \frac{n!}{n^n} \sqrt{\left\lvert \Delta_K \right\rvert}$ | |
for brevity. | |
Note that it is a constant depending on $K$. | |
\end{definition} | |
So that's cool and all, but what we really wanted was to show that | |
the class group is finite. | |
How can the Minkowski bound help? | |
Well, you might notice that we can rewrite it to say | |
\[ \Norm\left( (\alpha) \cdot \ka\inv \right) \le M_K \] | |
where $M_K$ is some constant depending on $K$, and $\alpha \in \ka$. | |
\begin{ques} | |
Show that $(\alpha) \cdot \ka\inv$ is an integral ideal. | |
(Unwind definitions.) | |
\end{ques} | |
But in the class group we \emph{mod out} by principal ideals like $(\alpha)$. | |
If we shut our eyes for a moment and mod out, the above statement becomes ``$\Norm(\ka\inv)\le M_K$''. | |
The precise statement of this is | |
\begin{corollary} | |
\label{cor:class_rep_minkowski} | |
Let $K$ be a number field, and pick a fractional ideal $J$. | |
Then we can find $\alpha$ such that $\kb = (\alpha) \cdot J$ is integral | |
and $\Norm(\kb) \le M_K$. | |
\end{corollary} | |
\begin{proof} | |
For fractional ideals $I$ and $J$ write $I \sim J$ to mean | |
that $I = (\alpha)J$ for some $\alpha$; then $\Cl_K$ is just modding out by $\sim$. | |
Let $J$ be a fractional ideal. | |
Then $J\inv$ is some other fractional ideal. | |
By definition, for some $\alpha \in \OO_K$ we have that $\alpha J\inv$ is an integral ideal $\ka$. | |
The Minkowski bound tells us that for some $x \in \ka$, we have $\Norm(x\ka\inv) \le M_K$. | |
But $x\ka\inv \sim \ka\inv = (\alpha J\inv)\inv \sim J$. | |
\end{proof} | |
\begin{corollary}[Finiteness of class group] | |
Class groups are always finite. | |
\end{corollary} | |
\begin{proof} | |
For every class in $\Cl_K$, | |
we can identify an integral ideal $\ka$ with norm less than $M_K$. | |
We just have to show there are finitely many such integral ideals; | |
this will mean there are finitely many classes. | |
Suppose we want to build such an ideal $\ka = \kp_1^{e_1} \dots \kp_m^{e_m}$. | |
Recall that a prime ideal $\kp_i$ must have some rational prime $p$ inside it, | |
meaning $\kp_i$ divides $(p)$ and $p$ divides $\Norm(\kp_i)$. | |
So let's group all the $\kp_i$ we want to build $\ka$ with based on which $(p)$ they came from. | |
\begin{center} | |
\begin{asy} | |
size(4cm); | |
label("$(2)$", (1,0), dir(90)); | |
for (int i=1; i <= 4; ++i) dot( (1, -i/4) ); | |
label("$(3)$", (2,0), dir(90)); | |
for (int i=1; i <= 6; ++i) dot( (2, -i/4) ); | |
label("$(5)$", (3,0), dir(90)); | |
for (int i=1; i <= 2; ++i) dot( (3, -i/4) ); | |
label("$\dots$", (4,0), dir(90)); | |
\end{asy} | |
\end{center} | |
To be more dramatic: imagine you have a \emph{cherry tree}; | |
each branch corresponds to a prime $(p)$ | |
and contains as cherries (prime ideals) the factors of $(p)$ (finitely many). | |
Your bucket (the ideal $\ka$ you're building) can only hold a total weight | |
(norm) of $M_K$. So you can't even touch the branches higher than $M_K$. | |
You can repeat cherries (oops), | |
but the weight of a cherry on branch $(p)$ is definitely $\ge p$; | |
all this means that the number of ways to build $\ka$ is finite. | |
\end{proof} | |
\section{Computation of class numbers} | |
\begin{definition} | |
The order of $\Cl_K$ is called the \vocab{class number} of $K$. | |
\end{definition} | |
\begin{remark} | |
If $\Cl_K = 1$, then $\OO_K$ is a PID, hence a UFD. | |
\end{remark} | |
By computing the actual value of $M_K$, | |
we can quite literally build the entire ``cherry tree'' mentioned in the previous proof. | |
Let's give an example how! | |
\begin{proposition} | |
The field $\QQ(\sqrt{-67})$ has class number $1$. | |
\end{proposition} | |
\begin{proof} | |
Since $K = \QQ(\sqrt{-67})$ has signature $(0,1)$ | |
and discriminant $\Delta_K = -67$ (since $-67 \equiv 1 \pmod 4$) | |
we can compute | |
\[ M_K = \left( \frac 4\pi \right)^{1} \cdot \frac{2!}{2^2} \sqrt{67} \approx 5.2. \] | |
That means we can cut off the cherry tree after $(2)$, $(3)$, $(5)$, since any | |
cherries on these branches will necessarily have norm $\ge M_K$. | |
We now want to factor each of these in $\OO_K = \ZZ[\theta]$, where $\theta = \frac{1+\sqrt{-67}}{2}$ | |
has minimal polynomial $x^2 - x + 17$. | |
But something miraculous happens: | |
\begin{itemize} | |
\ii When we try to reduce $x^2-x+17 \pmod 2$, we get an irreducible polynomial $x^2-x+1$. | |
By the factoring algorithm (\Cref{thm:factor_alg}) this means $(2)$ is prime. | |
\ii Similarly, reducing mod $3$ gives $x^2-x+2$, which is irreducible. | |
This means $(3)$ is prime. | |
\ii Finally, for the same reason, $(5)$ is prime. | |
\end{itemize} | |
It's our lucky day; | |
all of the ideals $(2)$, $(3)$, $(5)$ are prime (already principal). | |
To put it another way, | |
each of the three branches has only one (large) cherry on it. | |
That means any time we put together an integral ideal with norm $\le M_K$, | |
it is actually principal. | |
In fact, these guys have norm $4$, $9$, $25$ respectively\dots | |
so we can't even touch $(3)$ and $(5)$, | |
and the only ideals we can get are $(1)$ and $(2)$ (with norms $1$ and $4$). | |
Now we claim that's all. | |
Pick a fractional ideal $J$. | |
By \Cref{cor:class_rep_minkowski}, | |
we can find an integral ideal $\kb \sim J$ with $\Norm(\kb) \le M_K$. | |
But by the above, either $\kb = (1)$ or $\kb = (2)$, both of which are principal, | |
and hence trivial in $\Cl_K$. | |
So $J$ is trivial in $\Cl_K$ too, as needed. | |
\end{proof} | |
Let's do a couple more. | |
\begin{theorem}[Gaussian integers {$\ZZ[i]$} form a UFD] | |
The field $\QQ(i)$ has class number $1$. | |
\end{theorem} | |
\begin{proof} | |
This is $\OO_K$ where $K = \QQ(i)$, so we just want $\Cl_K$ to be trivial. | |
We have $M_K = \frac{2}{\pi}\sqrt{4} < 2$. | |
So every class | |
has an integral ideal of norm $\kb$ satisfying | |
\[ | |
\Norm(\kb) | |
\le \left( \frac4\pi \right)^{1} \cdot \frac{2!}{2^2} \cdot \sqrt{4} | |
= \frac4\pi < 2. | |
\] | |
Well, that's silly: we don't have any branches to pick from at all. | |
In other words, we can only have $\kb = (1)$. | |
\end{proof} | |
Here's another example of something that still turns out to be unique factorization, | |
but this time our cherry tree will actually have cherries that can be picked. | |
\begin{proposition}[{$\ZZ[\sqrt7]$} is a UFD] | |
The field $\QQ(\sqrt7)$ has class number $1$. | |
\end{proposition} | |
\begin{proof} | |
First we compute the Minkowski bound. | |
\begin{ques} | |
Check that $M_K \approx 2.646$. | |
\end{ques} | |
So this time, the only branch is $(2)$. Let's factor $(2)$ as usual: the polynomial $x^2+7$ | |
reduces as $(x-1)(x+1) \pmod 2$, and hence | |
\[ (2) = \left( 2, \sqrt7-1 \right) \left( 2, \sqrt7+1 \right). \] | |
Oops! We now have two cherries, and they both seem reasonable. | |
But actually, I claim that | |
\[ \left( 2, \sqrt7-1 \right) = \left( 3 - \sqrt 7 \right). \] | |
\begin{ques} | |
Prove this. | |
\end{ques} | |
So both the cherries are principal ideals, and as before we conclude that $\Cl_K$ is trivial. | |
But note that this time, the prime ideal $(2)$ actually splits; we got lucky | |
that the two cherries were principal but this won't always work. | |
\end{proof} | |
How about some nontrivial class groups? | |
First, we use a lemma that will help us with | |
narrowing down the work in our cherry tree. | |
\begin{lemma}[Ideals divide their norms] | |
Let $\kb$ be an integral ideal with $\Norm(\kb) = n$. | |
Then $\kb$ divides the ideal $(n)$. | |
\end{lemma} | |
\begin{proof} | |
By definition, $n = \left\lvert \OO_K / \kb \right\rvert$. | |
Treating $\OO_K/\kb$ as an (additive) abelian group and using Lagrange's theorem, we find | |
\[ 0 \equiv | |
\underbrace{\alpha + \dots + \alpha}_{n\text{ times}} = n\alpha | |
\pmod \kb \qquad\text{for all } \alpha \in \OO_K. \] | |
Thus $(n) \subseteq \kb$, done. | |
\end{proof} | |
Now we can give such an example. | |
\begin{proposition}[Class group of $\QQ(\sqrt{-17})$] | |
The number field $K = \QQ(\sqrt{-17})$ has class group $\Zc 4$. | |
\end{proposition} | |
You are not obliged to read the entire proof in detail, | |
as it is somewhat gory. | |
The idea is just that there are some cherries which are not trivial in the class group. | |
\begin{proof} | |
Since $\Delta_K = -68$, we compute | |
the Minkowski bound | |
\[ M_K = \frac{4}{\pi} \sqrt{17} < 6. \] | |
Now, it suffices to factor with $(2)$, $(3)$, $(5)$. | |
The minimal polynomial of $\sqrt{-17}$ is $x^2+17$, so as usual | |
\begin{align*} | |
(2) &= (2, \sqrt{-17}+1)^2 \\ | |
(3) &= (3, \sqrt{-17}-1)(3,\sqrt{-17}+1) \\ | |
(5) &= (5) | |
\end{align*} | |
corresponding to the factorizations of $x^2+17$ modulo each of $2$, $3$, $5$. | |
Set $\kp = (2, \sqrt{-17}+1)$ and $\kq_1 = (3, \sqrt{-17}-1)$, $\kq_2 = (3, \sqrt{-17}+1)$. | |
We can compute | |
\[ \Norm(\kp) = 2 \quad\text{and}\quad \Norm(\kq_1) = \Norm(\kq_2) = 3. \] | |
In particular, they are not principal. | |
The ideal $(5)$ is out the window; it has norm $25$. | |
Hence, the three cherries are $\kp$, $\kq_1$, $\kq_2$. | |
The possible ways to arrange these cherries into ideals with norm $\le 5$ are | |
\[ \left\{ (1), \kp, \kq_1, \kq_2, \kp^2 \right\}. \] | |
However, you can compute \[ \kp^2 = (2) \] so $\kp^2$ and $(1)$ are in the same class group; | |
that is, they are trivial. | |
In particular, the class group has order at most $4$. | |
From now on, let $[\ka]$ denote the class (member of the class group) that $\ka$ is in. | |
Since $\kp$ isn't principal (so $[\kp] \neq [(1)]$), it follows that $\kp$ has order two. | |
So Lagrange's theorem says that $\Cl_K$ has order either $2$ or $4$. | |
Now we claim $[\kq_1]^2 \neq [(1)]$, which implies that $\kq_1$ has order greater than $2$. | |
If not, $\kq_1^2$ is principal. | |
We know $\Norm(\kq_1) = 3$, | |
so this can only occur if $\kq_1^2 = (3)$; | |
this would force $\kq_1 = \kq_2$. | |
This is impossible since $\kq_1 + \kq_2 = (1)$. | |
Thus, $\kq_1$ has even order greater than $2$. | |
So it has to have order $4$. | |
From this we deduce \[ \Cl_K \cong \Zc 4. \qedhere \] | |
\end{proof} | |
\begin{remark} | |
When we did this at Harvard during Math 129, | |
there was a five-minute interruption in which students (jokingly) complained | |
about the difficulty of evaluating $\frac{4}{\pi} \sqrt{17}$. Excerpt: | |
\begin{quote} | |
``Will we be allowed to bring a small calculator on the exam?'' -- Student 1 \\ | |
``What does the size have to do with anything? You could have an Apple Watch'' -- Professor \\ | |
``Just use the fact that $\pi \ge 3$'' -- me \\ | |
``Even [other professor] doesn't know that, how are we supposed to?'' -- Student 2 \\ | |
``You have to do this yourself!'' -- Professor \\ | |
``This is an outrage.'' -- Student 1 | |
\end{quote} | |
\end{remark} | |
\section\problemhead | |
\begin{problem} | |
Show that $K = \QQ(\sqrt{-163})$ has trivial class group, | |
and hence $\OO_K = \ZZ\left[ \frac{1+\sqrt{-163}}{2} \right]$ | |
is a UFD.\footnote{In fact, | |
$n = 163$ is the largest number | |
for which $\QQ(\sqrt{-n})$ has trivial class group. | |
The complete list is $1, 2, 3, 7, 11, 19, 43, 67, 163$, | |
the \vocab{Heegner numbers}. | |
You might notice Euler's prime-generating polynomial $t^2+t+41$ | |
when doing the above problem. Not a coincidence!} | |
\begin{hint} | |
Repeat the previous procedure. | |
\end{hint} | |
\end{problem} | |
\begin{problem} | |
Determine the class group of $\QQ(\sqrt{-31})$. | |
\begin{hint} | |
You should get a group of order three. | |
\end{hint} | |
\end{problem} | |
\begin{problem}[China TST 1998] | |
Let $n$ be a positive integer. | |
A polygon in the plane (not necessarily convex) has area greater than $n$. | |
Prove that one can translate it so that it contains at least $n+1$ lattice points. | |
\begin{hint} | |
Mimic the proof of part (a) of Minkowski's theorem. | |
\end{hint} | |
\end{problem} | |
\begin{problem} | |
[\Cref{lem:vol_OK_mesh}] | |
\label{prob:OK_linalg} | |
Consider the composition of the embeddings $K \injto \RR^{r_1} \times \CC^{r_2} \taking\sim \RR^n$. | |
Show that the image of $\OO_K \subseteq K$ has mesh equal to | |
\[ \frac{1}{2^{r_2}} \sqrt{\left\lvert \Delta_K \right\rvert}. \] | |
\begin{hint} | |
Linear algebra. | |
\end{hint} | |
\end{problem} | |
\begin{problem} | |
Let $p \equiv 1 \pmod 4$ be a prime. | |
Show that there are unique integers $a > b > 0$ such that $a^2+b^2=p$. | |
\begin{hint} | |
Factor in $\QQ(i)$. | |
\end{hint} | |
% \begin{sol} | |
% Let's factor $(p)$ in $\QQ(i)$, which has ring of integers $\ZZ[i]$. | |
% Using \Cref{thm:factor_alg}, we get $\ZZ[x] / (p,x^2+1) \cong \FF_p[x] / (x^2+1)$. | |
% Since $p \equiv 1 \pmod 4$, $x^2+1 \equiv (x+t)(x-t) \pmod p$ for some $t$. | |
% Thus $(p) = (p, t+i)(p, t-i)$ is the full decomposition into prime factors. | |
% \end{sol} | |
\end{problem} | |
\begin{problem} | |
[Korea national olympiad 2014] | |
Let $p$ be an odd prime and $k$ a positive integer such that $p \mid k^2+5$. | |
Prove that there exist positive integers $m$, $n$ such that $p^2 = m^2+5n^2$. | |
\begin{hint} | |
Factor $p$, show that the class group of $\QQ(\sqrt{-5})$ has order two. | |
\end{hint} | |
\begin{sol} | |
Let $K = \QQ(\sqrt{-5})$. Check that $\Cl_K$ has order two using the Minkowski bound; | |
moreover $\Delta_K = 20$. | |
Now note that $\OO_K = \ZZ[\sqrt{-5}]$, and $x^2+5$ factors mod $p$ as $(x+k)(x-k)$; | |
hence in $\OO_K$ we have $(p) = (p, \sqrt{-5}+k)(p, \sqrt{-5}-k) = \kp_1 \kp_2$, say. | |
For $p > 5$ the prime $p$ does not ramify and we have $\kp_1 \neq \kp_2$, since $\Delta_K = 20$. | |
Then $(p^2) = \kp_1^2 \cdot \kp_2^2$. Because the class group has order two, | |
both $\kp_1^2$ and $\kp_2^2$ are principal, and because $\kp_1 \neq \kp_2$ they are distinct. | |
Thus $p^2$ is a nontrivial product of two elements of $\OO_K$; from this we can extract the desired factorization. | |
\end{sol} | |
\end{problem} | |