Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Integrality and valuation rings} | |
\label{intchapter} | |
The notion of integrality is familiar from number theory: it is similar to | |
``algebraic'' but with the polynomials involved are required to be monic. In algebraic geometry, integral | |
extensions of rings correspond to correspondingly nice morphisms on the | |
$\spec$'s---when the extension is finitely generated, it turns out that the | |
fibers are finite. That is, there are only finitely many ways to lift a prime | |
ideal to the extension: if $A \to B$ is integral and finitely generated, then | |
$\spec B \to \spec A$ has finite fibers. | |
Integral domains that are \emph{integrally closed} in their quotient field will play an | |
important role for us. Such ``normal domains'' are, for example, regular in | |
codimension one, which means that the theory of Weil divisors | |
(see \cref{weildivsec}) applies | |
to them. It is particularly nice because Weil divisors are sufficient to | |
determine whether a function is regular on a normal variety. | |
A canonical example of an integrally closed ring is a valuation ring; we shall | |
see in this chapter that any integrally closed ring is an intersection of such. | |
\section{Integrality} | |
\subsection{Fundamentals} | |
As stated in the introduction to the chapter, integrality is a condition on | |
rings parallel to that of algebraicity for field extensions. | |
\begin{definition} \label{intdefn} | |
Let $R$ be a ring, and $R'$ an $R$-algebra. An element $x \in R'$ | |
is said to be \textbf{integral} over $R$ if $x$ satisfies a monic polynomial | |
equation in $R[X]$, say | |
\[ x^n + r_1 x^{n-1} + \dots + r_n = 0, \quad r_1, \dots, r_n \in R. \] | |
We can say that $R'$ is \textbf{integral} over $R$ if every $x \in R'$ is | |
integral over $R$. | |
\end{definition} | |
Note that in the definition, we are not requiring $R$ to be a \emph{subring} of | |
$R'$. | |
\begin{example} \label{sixthroot} | |
$\frac{1+\sqrt{-3}}{2}$ is integral over $\mathbb{Z}$; it is in fact a sixth | |
root of unity, thus satisfying the equation $X^6 -1 = 0$. | |
However, $\frac{1+\sqrt{5}}{2}$ is not integral over $\mathbb{Z}$. To explain this, | |
however, we will | |
need to work a bit more (see \cref{onegeneratorintegral} below). | |
\end{example} | |
\begin{example} | |
Let $L/K$ be a field extension. Then $L/K$ is integral if and only if it is | |
algebraic, since $K$ is a field and we can divide polynomial equations by the | |
leading coefficient to make them monic. | |
\end{example} | |
\begin{example} | |
Let $R$ be a graded ring. Then the subring $R^{(d)} \subset R$ was defined in | |
\cref{dpowerofring}; recall that this consists of elements of $R$ all of whose | |
nonzero homogeneous components live in degrees that are multiples of $d$. | |
Then the $d$th power of any homogeneous | |
element in $R$ is in $R^{(d)}$. As a result, every homogeneous element of $R$ | |
is integral over $R^{(d)}$. | |
\end{example} | |
We shall now interpret the condition of integrality in terms of finite | |
generation of certain modules. | |
Suppose $R$ is a ring, and $R'$ an $R$-algebra. Let $x \in R'$. | |
\begin{proposition} \label{onegeneratorintegral} | |
$x \in R'$ is integral over $R$ if and only if the subalgebra $R[x] \subset R'$ | |
(generated by $R, x$) is a finitely generated | |
$R$-module. | |
\end{proposition} | |
This notation is an abuse of notation (usually $R[x]$ refers to a polynomial | |
ring), but it should not cause confusion. | |
This result for instance lets us show that $\frac{1+\sqrt{-5}}{2}$ is not integral | |
over $\mathbb{Z}$, because when you keep taking powers, you get arbitrarily | |
large denominators: the arbitrarily large denominators imply that it cannot be | |
integral. | |
\begin{proof} | |
If $x \in R'$ is integral, then $x$ satisfies | |
\[ x^n + r_1 x^{n-1}+\dots+r_n = 0, \quad r_i \in R. \] | |
Then $R[x]$ is generated as an $R$-module by $1, x, \dots, x^{n-1}$. This is | |
because the submodule of $R'$ generated by $1, x ,\dots, x^{n-1}$ is closed under | |
multiplication by $R$ and by multiplication by $x$ (by the above equation). | |
Now suppose $x$ generates a subalgebra $R[x] \subset R'$ which is a finitely | |
generated $R$-module. Then the increasing sequence | |
of $R$-modules generated by $\{1\}, \left\{1, x\right\}, \left\{1, x, | |
x^2\right\} | |
, \dots$ must stabilize, since the union is $R[x]$.\footnote{As an easy | |
exercise, one may see that if a finitely generated module $M$ is the union of | |
an increasing sequence of submodules $M_1 \subset M_2 \subset M_3 \subset | |
\dots$, then $M = M_n$ for some $n$; we just need to take $n$ large enough | |
such that $M_n$ contains each of the finitely many generators of $M$.} It follows that some $x^n$ | |
can be expressed as a linear combination of smaller powers of $x$. Thus $x$ is | |
integral over $R$. | |
\end{proof} | |
So, if $R'$ is an $R$-module, we can say that | |
an element $x \in R'$ is \textbf{integral} over $R$ if either of the following | |
equivalent conditions are satisfied: | |
\begin{enumerate} | |
\item There is a monic polynomial in $R[X]$ which vanishes on $x$. | |
\item $R[x] \subset R'$ is a finitely generated $R$-module. | |
\end{enumerate} | |
\begin{example} | |
Let $F$ be a field, $V$ a finite-dimensional $F$-vector space, $T: V \to V$ a | |
linear transformation. Then the ring generated by $T$ and $F$ inside | |
$\mathrm{End}_F(V)$ (which is a noncommutative ring) is finite-dimensional | |
over $F$. | |
Thus, by similar reasoning, $T$ must satisfy a polynomial equation with | |
coefficients in $F$ (e.g. the characteristic polynomial). | |
\end{example} | |
Of course, if $R'$ is integral over $R$, $R'$ may not be a finitely generated | |
$R$-module. For instance, $\overline{\mathbb{Q}}$ is not a finitely generated | |
$\mathbb{Q}$-module, although the extension is integral. As we shall see in | |
the next section, this is always the case if $R'$ is a finitely | |
generated $R$-\emph{algebra}. | |
We now will add a third equivalent condition to this idea of ``integrality,'' | |
at least in the case where the structure map is an injection. | |
\begin{proposition} \label{thirdintegralitycriterion} | |
Let $R$ be a ring, and suppose $R$ is a subring of $R'$. | |
$x \in R'$ is integral if and only if there exists a | |
finitely generated faithful $R$-module $M \subset R'$ such that $R \subset M$ and | |
$xM \subset M$. | |
\end{proposition} | |
A module $M$ is \emph{faithful} if $x M = 0$ implies $x=0$. That is, the map | |
from $R$ into the $\mathbb{Z}$-endomorphisms of $M$ is injective. | |
If $R$ is a \emph{subring} of $R'$ (i.e. the structure map $R \to R'$ is | |
injective), then $R'$ for instance is a faithful $R$-module. | |
\begin{proof} | |
It's obvious that the second condition above (equivalent to integrality) | |
implies the condition of this | |
proposition. Indeed, one could just take $M = R[x]$. | |
Now let us prove that if there exists such an $M$ which is finitely generated, | |
then $x$ is integral. Just because $M$ is finitely generated, the | |
submodule $R[x]$ is not obviously finitely generated. In particular, this | |
implication requires a bit of proof. | |
We shall prove that the condition of this proposition implies integrality. | |
Suppose $y_1, \dots, y_k \in M$ generate $M$ as $R$-module. Then multiplication | |
by $x$ gives an $R$-module map $M \to M$. In particular, we can write | |
\[ xy_i = \sum a_{ij} y_j \] | |
where each $a_{ij} \in R$. | |
These $\left\{a_{ij}\right\}$ may not be unique, but let us make some choices; | |
we get a $k$-by-$k$ matrix $A \in M_k(R)$. The claim is that $x$ satisfies the | |
characteristic polynomial of $A$. | |
Consider the matrix | |
\[ (x 1 - A) \in M_n(R'). \] | |
Note that $(x1-A)$ annihilates each $y_i$, by the choice of $A$. | |
We can consider the adjoint $B = (x1 -A)^{adj}$. Then | |
\[ B(x1 - A) = \det(x1 - A) 1. \] | |
This product of matrices obviously annihilates each vector $y_i$. It follows | |
that | |
\[ (\det(x1 - A) y_i = 0, \quad \forall i, \] | |
which implies that $\det (x1-A)$ kills $M$. This implies that $\det (x1 - | |
A)=0$ since $M$ is faithful. | |
As a result, $x$ satisfies the characteristic polynomial. | |
\end{proof} | |
\begin{exercise} | |
Let $R$ be a noetherian | |
local domain with maximal ideal $\mathfrak{m}$. As we will define shortly, $R$ | |
is \emph{integrally closed} if every element of the quotient field $K=K(R)$ | |
integral over $R$ belongs to $R$ itself. Then if $x \in K$ and $x \mathfrak{m} | |
\subset \mathfrak{m}$, we have $x \in R$. | |
\end{exercise} | |
\begin{exercise} Let us say that an $A$-module is \emph{$n$-generated} | |
if it is generated by at most $n$ elements. | |
Let $A$ and $B$ be two rings such that $A\subset B$, so that $B$ is an | |
$A$-module. | |
Let | |
$n\in\mathbb{N}$. Let $u\in B$. Then, the following four assertions | |
are equivalent: | |
\begin{enumerate} | |
\item There exists a monic polynomial | |
$P\in A\left[ X\right] $ with $\deg P=n$ and $P\left( u\right) =0$. | |
\item There exist a $B$-module $C$ and an | |
$n$-generated $A$-submodule $U$ of $C$ such that $uU\subset U$ and such that | |
every $v\in B$ satisfying $vU=0$ satisfies $v=0$. (Here, $C$ is an $A$-module, | |
since $C$ is a $B$-module and $A\subset B$.) | |
\item There exists an $n$-generated | |
$A$-submodule $U$ of $B$ such that $1\in U$ and $uU\subset U$. | |
\item As an $A$-module, $A[u]$ is spanned by $1, u, \dots, | |
u^{n-1}$.\end{enumerate} | |
\end{exercise} | |
We proved this to show that the set of integral elements is well behaved. | |
\begin{proposition} | |
Let $R \subset R'$. Let $S = \left\{x \in R': x \text{ is integral over } | |
R\right\}$. Then $S$ is a subring of $R'$. In particular, it is closed under | |
addition and multiplication. | |
\end{proposition} | |
\begin{proof} | |
Suppose $x,y \in S$. | |
We can consider the finitely generated modules $R[x], R[y] \subset R'$ | |
generated (as algebras) by $x$ over $R$. By assumption, these are finitely | |
generated $R$-modules. In particular, the tensor product | |
\[ R[x] \otimes_R R[y] \] | |
is a finitely generated $R$-module (by \cref{fingentensor}). | |
We have a ring-homomorphism $R[x]\otimes_R R[y] \to R'$ | |
which comes from the inclusions $R[x], R[y] \rightarrowtail R'$. | |
Let $M$ be the image of $R[x] \otimes_R R[y]$ in $R'$. Then $M$ is an | |
$R$-submodule of $R'$, indeed an $R$-subalgebra containing $x,y$. Also, $M$ is | |
finitely generated. Since $x+y, xy\in M$ and $M$ is a subalgebra, it | |
follows that | |
\[ (x+y) M \subset M, \quad xy M \subset M. \] | |
Thus $x+y, xy$ are integral over $R$. | |
\end{proof} | |
Let us consider the ring $\mathbb{Z}[\sqrt{-5}]$; this is the canonical | |
example of a ring where unique factorization fails. This is because \( 6 = 2 | |
\times 3 = (1+\sqrt{-5})(1-\sqrt{-5}). \) | |
One might ask: what about $\mathbb{Z}[\sqrt{-3}]$? It turns out that | |
$\mathbb{Z}[\sqrt{-3}]$ lacks unique factorization as well. Indeed, here we have | |
\[ (1 - \sqrt{-3})(1+\sqrt{-3}) = 4 = 2 \times 2. \] | |
These elements can be factored no more, and $1 - \sqrt{-3}$ and $2$ do not | |
differ by units. | |
So in this ring, we have a failure of unique factorization. Nonetheless, the | |
failure of unique factorization in $\mathbb{Z}[\sqrt{-3}]$ is less | |
noteworthy, because $\mathbb{Z}[\sqrt{-3}]$ is not \emph{integrally closed}. Indeed, it turns out that $\mathbb{Z}[\sqrt{-3}]$ is | |
contained in the larger ring | |
\( \mathbb{Z}\left[ \frac{1 + \sqrt{-3}}{2}\right], \) | |
which does have unique factorization, and this larger ring is finite over | |
$\mathbb{Z}[\sqrt{-3}]$.\footnote{In fact, $\mathbb{Z}[\sqrt{-3}]$ is an index two subgroup of $\mathbb{Z}\left[ | |
\frac{1 + \sqrt{-3}}{2}\right]$, as the ring $\mathbb{Z}[ \frac{1 + \sqrt{-3}}{2}]$ | |
can be described as the set of elements $a + | |
b\sqrt{-3}$ where $a,b$ are either both integers or both integers plus | |
$\frac{1}{2}$, as is easily seen: this set is closed under addition and | |
multiplication.} | |
Since being integrally closed is a | |
prerequisite for having unique factorization (see \cref{} below), the failure in | |
$\mathbb{Z}[\sqrt{-3}]$ is not particularly surprising. | |
Note that, by contrast, $\mathbb{Z}[ \frac{1 + \sqrt{-5}}{2}]$ does not | |
contain $\mathbb{Z}[\sqrt{-5}]$ as a finite index subgroup---it cannot be | |
slightly enlarged in the same sense. When one enlarges $\mathbb{Z}[\sqrt{-5}]$, | |
one has to add a lot of stuff. | |
We will see more formally that $\mathbb{Z}[\sqrt{-5}]$ is \emph{integrally | |
closed} in its quotient field, while $\mathbb{Z}[\sqrt{-3}]$ is not. Since | |
unique factorization domains are automatically integrally closed, the failure | |
of $\mathbb{Z}[\sqrt{-5}]$ to be a UFD is much more significant than that of | |
$\mathbb{Z}[\sqrt{-3}]$. | |
\subsection{Le sorite for integral extensions} | |
In commutative algebra and algebraic geometry, there are a lot of standard | |
properties that a \emph{morphism} of rings $\phi: R \to S$ can have: it could | |
be of \emph{finite type} (that is, $S$ is finitely generated over $\phi(R)$), | |
it could be \emph{finite} (that is, $S$ is a finite $R$-module), or it could | |
be \emph{integral} (which we have defined in \cref{intdefn}). There are many more examples | |
that we will encounter as we dive deeper into commutative algebra. | |
In algebraic geometry, there are corresponding properties of morphisms of | |
\emph{schemes,} and there are many more interesting ones here. | |
In these cases, there is usually---for any reasonable property---a standard | |
and familiar list of | |
properties that one proves about them. We will refer to such lists as | |
``sorites,'' and prove our first one now. | |
\begin{proposition}[Le sorite for integral morphisms] | |
\begin{enumerate} | |
\item For any ring $R$ and any ideal $I \subset R$, the map $R \to R/I$ is | |
integral. | |
\item If $\phi: R \to S$ and $\psi: S \to T$ are integral morphisms, then so | |
is $\psi \circ \phi: R \to T$. | |
\item If $\phi: R \to S$ is an integral morphism and $R'$ is an $R$-algebra, | |
then the base-change | |
$R' \to R' \otimes_R S$ is integral. | |
\end{enumerate} | |
\end{proposition} | |
\begin{proof} | |
The first property is obvious. For the second, the condition of | |
integrality in a morphism of rings depends on the inclusion of the image | |
in the codomain. So we can suppose that $R \subset S \subset T$. Suppose $t | |
\in T$. By assumption, there is a monic polynomial equation | |
\[ t^n + s_1 t^{n-1} + \dots + s_n = 0 \] | |
that $t$ satisfies, where each $s_i \in S$. | |
In particular, we find that $t$ is integral over $R[s_1, \dots, s_n]$. | |
As a result, the module $R[s_1, \dots, s_n, t]$ is finitely generated over the | |
ring $R'=R[s_1, \dots, s_n]$. | |
By the following \cref{finitelygeneratedintegral}, $R'$ is a finitely generated | |
$R$-module. In | |
particular, $R[s_1, \dots, s_n,t]$ is a finitely generated $R$-module (not | |
just a | |
finitely generated $R'$-module). | |
Thus the $R$-module $R[s_1, \dots, s_n,t]$ is a faithful | |
$R'$ module, finitely generated over $R$, which is preserved under | |
multiplication by $t$. | |
\end{proof} | |
We now prove a result that can equivalently be phrased as ``finite type plus | |
integral implies finite'' for a map of rings. | |
\begin{proposition} \label{finitelygeneratedintegral} | |
Let $R'$ be a finitely generated, integral $R$-algebra. Then $R'$ is a | |
finitely generated $R$-module: that is, the map $R \to R'$ is finite. | |
\end{proposition} | |
\begin{proof} | |
Induction on the number of generators of $R'$ as $R$-algebra. For one | |
generator, this follows from \rref{onegeneratorintegral}. | |
In general, we will have $R' = R[\alpha_1 ,\dots, \alpha_n]$ for some | |
$\alpha_i \in R'$. | |
By the inductive hypothesis, $R[\alpha_1 , \dots, \alpha_{n-1}]$ is a finite | |
$R$-module; by the case of one generator, $R'$ is a finite $R[\alpha_1, \dots, | |
\alpha_{n-1}]$-module. This establishes the result by the next exercise. | |
\end{proof} | |
\begin{exercise} | |
Let $R \to S, S \to T$ be morphisms of rings. Suppose $S$ is a finite | |
$R$-module and $T$ a finite $T$-module. Then $T$ is a finite $R$-module. | |
\end{exercise} | |
\subsection{Integral closure} | |
Let $R, R'$ be rings. | |
\begin{definition} | |
If $R \subset R'$, then the set $S = \left\{x \in R': x \ \mathrm{is \ | |
integral }\right\}$ is called the \textbf{integral closure} of $R$ in $R'$. We | |
say that $R$ is \textbf{integrally closed in $R'$} if $S = R'$. | |
When $R$ is a domain, and $K$ is the quotient field, we shall simply | |
say that $R$ is \textbf{integrally closed} if it is integrally closed in | |
$K$. | |
Alternatively, some people say that $R$ is \textbf{normal} in this case. | |
\end{definition} | |
Integral closure (in, say, the latter sense) is thus an operation that maps | |
integral domains to integral domains. It is easy to see that the operation is | |
\emph{idempotent:} the integral closure of the integral closure is the integral | |
closure. | |
\begin{example} | |
The integers $\mathbb{Z} \subset \mathbb{C}$ have as integral closure (in | |
$\mathbb{C}$) the set | |
of complex numbers satisfying a monic polynomial with integral | |
coefficients. This set is called the set of \textbf{algebraic integers}. | |
For instance, $i$ is an algebraic integer because it satisfies the equation $X^2 +1 = 0$. | |
$\frac{1 - \sqrt{-3}}{2}$ is an algebraic integer, as we talked about last | |
time; it is a sixth root of unity. On the other hand, $\frac{1+\sqrt{-5}}{2}$ | |
is not an algebraic integer. | |
\end{example} | |
\begin{example} | |
Take $\mathbb{Z} \subset \mathbb{Q}$. The claim is that $\mathbb{Z}$ is | |
integrally closed in its quotient field $\mathbb{Q}$, or simply---integrally | |
closed. | |
\end{example} | |
\begin{proof} | |
We will build on this proof later. Here is the point. Suppose $\frac{a}{b} | |
\in \mathbb{Q}$ satisfying an equation | |
\[ P(a/b) = 0, \quad P(t) = t^n + c_1 t^{n-1} + \dots + c_0 , \ \forall c_i \in | |
\mathbb{Z}.\] | |
Assume that $a,b$ have no common factors; we must prove that $b$ has no prime | |
factors, so is $\pm 1$. | |
If $b$ had a prime factor, say $q$, then we must obtain a contradiction. | |
We interrupt with a definition. | |
\begin{definition} | |
The \textbf{valuation at $q$} (or \textbf{$q$-adic valuation}) is the map | |
$v_q: \mathbb{Q}^* \to \mathbb{Z}$ is the | |
function sending $q^k (a/b)$ to $k$ if $q \nmid a,b$. We extend this to all | |
rational numbers via $v(0) = \infty$. | |
\end{definition} | |
In general, this just counts the number of factors of $q$ in the expression. | |
Note the general property that | |
\begin{equation} \label{val1} v_q(x+y) \geq \min( v_q(x), v_q(y)) . | |
\end{equation} | |
If $x,y$ are both divisible by some power of $q$, so is $x+y$; this is the | |
statement above. We also have the useful property | |
\begin{equation} \label{val2} v_q(xy) = v_q(x) + v_q(y). \end{equation} | |
Now return to the proof that $\mathbb{Z}$ is normal. We would like to show that | |
\( v_q(a/b) \geq 0. \) | |
This will prove that $b$ is not divisible by $q$. When we show this for all | |
$q$, it will follow that $a/b \in \mathbb{Z}$. | |
We are assuming that $P(a/b) = 0$. In particular, | |
\[ \left( \frac{a}{b} \right)^n = -c_1 \left( \frac{a}{b} \right)^{n-1} - | |
\dots - c_0. \] | |
Apply $v_q$ to both sides: | |
\[ n v_q ( a/b) \geq \min_{i>0} v_q( c_i (a/b)^{n-i}). \] | |
Since the $c_i \in \mathbb{Z}$, their valuations are nonnegative. In | |
particular, the right hand side is at least | |
\[ \min_{i>0} (n-i) v_q(a/b). \] | |
This cannot happen if $v_q(a/b)<0$, because $n-i < n$ for each $i>0$. | |
\end{proof} | |
This argument applies more generally. If $K$ is a field, and $R \subset K$ is | |
a subring ``defined | |
by valuations,'' such as the $v_q$, then $R$ is integrally closed in its | |
quotient field. | |
More precisely, | |
note the reasoning of the previous example: the key idea was that $\mathbb{Z} | |
\subset \mathbb{Q}$ was characterized by the rational numbers $x$ such that | |
$v_q(x) \geq 0$ for all primes $q$. | |
We can abstract this idea as follows. If there exists a family of functions $\mathcal{V}$ from $K^* | |
\to \mathbb{Z}$ (such as | |
$\left\{v_q: \mathbb{Q}^* \to \mathbb{Z}\right\}$) satisfying \eqref{val1} and | |
\eqref{val2} above such that $R$ is | |
the set of elements such that $v(x) \geq 0, v \in \mathcal{V}$ (along with | |
$0$), then $R$ is integrally closed in $K$. | |
We will talk more about | |
this, and about valuation rings, below. | |
\begin{example} | |
We saw earlier (\cref{sixthroot}) that $\mathbb{Z}[\sqrt{-3}]$ is not | |
integrally closed, as $\frac{1 + \sqrt{-3}}{2}$ is integral over this ring and | |
in the quotient field, but not in the ring. | |
\end{example} | |
We shall give more examples in the next subsection. | |
\subsection{Geometric examples} | |
Let us now describe the geometry of a non-integrally closed ring. | |
Recall that finitely generated (reduced) $\mathbb{C}$-algebras are supposed to | |
correspond to affine algebraic varieties. A \emph{smooth} variety (i.e., one | |
that is a complex manifold) will always correspond to an integrally closed | |
ring (though this relies on a deep result that a regular local ring is a | |
factorization domain, and consequently integrally closed): non-normality is a sign of singularities. | |
\begin{example} | |
Here is a ring which is not integrally closed. Take $\mathbb{C}[x, y]/(x^2 | |
- y^3)$. Algebraically, this is the subring of the polynomial ring | |
$\mathbb{C}[t]$ generated by $t^2$ and $t^3$. | |
In the complex plane, $\mathbb{C}^2$, this corresponds to the subvariety $C | |
\subset \mathbb{C}^2$ defined by $x^2 = | |
y^3$. In $\mathbb{R}^2$, this can be drawn: it has a singularity at $(x,y)=0$. | |
Note that $x^2 = y^3$ if and only if there is a complex number $z$ such that $x | |
= z^3, y = z^2$. This complex number $z$ can be recovered via $x/y$ when $x,y | |
\neq 0$. In particular, there is a map $\mathbb{C} \to C$ which sends $z \to | |
(z^3, z^2)$. At every point other than the origin, the inverse can be | |
recovered using rational functions. But this does not work at the origin. | |
We can think of $\mathbb{C}[x,y]/(x^2 - y^3)$ as the subring $R'$ of | |
$\mathbb{C}[z]$ | |
generated by $\left\{z^n, n \neq 1\right\}$. There is a map from | |
$\mathbb{C}[x,y]/(x^2 - y^3)$ sending $x \to z^3, y \to z^2$. Since these two | |
domains are isomorphic, and $R'$ is not integrally closed, it follows that | |
$\mathbb{C}[x,y]/(x^2 - y^3)$ is not integrally closed. | |
The element $z$ can be thought of as an element of the fraction field of $R'$ | |
or of $\mathbb{C}[x,y]/(x^2 - y^3)$. | |
It is integral, though. | |
The failure of the ring to be integrally closed has to do with the singularity | |
at the | |
origin. | |
\end{example} | |
We now give a generalization of the above example. | |
\begin{example} | |
This example is outside the scope of the present course. Say that $X \subset | |
\mathbb{C}^n$ is given as the zero locus of some holomorphic functions | |
$\left\{f_i: \mathbb{C}^{n} \to \mathbb{C}\right\}$. We just gave an example | |
when $n=2$. | |
Assume that $0 \in X$, i.e. each $f_i$ vanishes at the origin. | |
Let $R$ be the ring of germs of holomorphic functions $0$, in other words | |
holomorphic functions from small open neighborhoods of zero. Each of these | |
$f_i$ becomes an element of $R$. The ring | |
\( R/(\left\{f_i\right\} ) \) | |
is called the ring of germs of holomorphic functions on $X$ at zero. | |
Assume that $R$ is a domain. This assumption, geometrically, means that near | |
the point zero in $X$, $X$ can't be broken into two smaller closed analytic | |
pieces. The fraction field of $R$ is to be thought of as the ring of | |
germs of meromorphic functions on $X$ at zero. | |
We state the following without proof: | |
\begin{theorem} | |
Let $g/g'$ be an element of the fraction field, i.e. $g, g' \in R$. Then $g/g'$ | |
is integral over $R$ if and only if $g/g'$ is bounded near zero. | |
\end{theorem} | |
In the previous example of $X$ defined by $x^2 = y^3$, the function $x/y$ | |
(defined near the origin on the curve) is | |
bounded near the origin, so it is integral over the ring of germs of regular | |
functions. The reason it is not defined near the origin is \emph{not} that it | |
blows up. In fact, it extends continuously, but not holomorphically, to the | |
rest of the variety $X$. | |
\end{example} | |
\section{Lying over and going up} | |
We now interpret integrality in terms of the geometry of $\spec$. | |
In general, for $R \to S$ a ring-homomorphism, the induced map $\spec S \to | |
\spec R$ need not be topologically nice; for instance, even if $S$ is a | |
finitely generated $R$-algebra, the image of $\spec S$ in $\spec R$ need not | |
be either open or closed.\footnote{It is, however, true that if $R$ is | |
\emph{noetherian} (see \rref{noetherian}) and $S$ finitely generated over | |
$R$, then the image of $\spec S$ is \emph{constructible,} that is, a finite | |
union of locally closed subsets. \add{this result should be added sometime}.} | |
We shall see that under conditions of integrality, more can be said. | |
\subsection{Lying over} | |
In general, given a morphism of algebraic varieties $f: X \to Y$, the image of | |
a closed subset $Z \subset X$ is far from closed. For instance, a regular function $f: X \to \mathbb{C}$ | |
that is a closed map would have to be either surjective or constant (if $X$ is | |
connected, say). | |
Nonetheless, under integrality hypotheses, we can say more. | |
\begin{proposition}[Lying over] \label{lyingover} | |
If $\phi: R \to R'$ is an integral morphism, then the induced map | |
\[ \spec R' \to \spec R \] | |
is a closed map; it is surjective if $\phi$ is injective. | |
\end{proposition} | |
Another way to state the last claim, without mentioning $\spec R'$, is the | |
following. Assume $\phi$ is injective and integral. Then if | |
$\mathfrak{p} \subset R$ is prime, then there exists $\mathfrak{q} \subset R'$ | |
such that $\mathfrak{p}$ is the inverse image $\phi^{-1}(\mathfrak{q})$. | |
\begin{proof} First suppose $\phi$ injective, in which case we must prove the | |
map $\spec R' \to \spec R$ surjective. | |
Let us reduce to the case of a local ring. | |
For a prime $\mathfrak{p} \in \spec R$, we must show that $\mathfrak{p}$ | |
arises as the inverse image of an element of $\spec R'$. | |
So we replace $R$ with $R_{\mathfrak{p}}$. We get a map | |
\[ \phi_{\mathfrak{p}}: R_{\mathfrak{p}} \to (R- \mathfrak{p})^{-1} R' \] | |
which is injective if $\phi$ is, since localization is an exact functor. Here | |
we | |
have localized both $R, R'$ at the multiplicative subset $R - \mathfrak{p}$. | |
Note that $\phi_{\mathfrak{p}}$ is an integral extension too. This follows | |
because integrality is preserved by base-change. | |
We will now prove the result for $\phi_{\mathfrak{p}}$; in particular, we | |
will show | |
that there is a prime ideal of $(R- \mathfrak{p})^{-1} R'$ that pulls back to | |
$\mathfrak{p}R_{\mathfrak{p}}$. These will imply that if we pull this prime | |
ideal back to $R'$, it will pull back to $\mathfrak{p}$ in $R$. In detail, we | |
can consider the diagram | |
\[ \xymatrix{ | |
\spec (R-\mathfrak{p})^{-1} R'\ar[d] \ar[r] & \spec R_{\mathfrak{p}} \ar[d] \\ | |
\spec R' \ar[r] & \spec R | |
}\] | |
which shows that if $\mathfrak{p} R_{\mathfrak{p}}$ appears in the image of the top map, then | |
$\mathfrak{p}$ | |
arises as the image of something in $\spec R'$. | |
So it is sufficient for the proposition (that is, the case of $\phi$ | |
injective) to handle the case of $R$ local, and | |
$\mathfrak{p}$ the maximal ideal. | |
In other words, we need to show that: | |
\begin{quote} | |
If $R$ is a \emph{local} ring, $\phi: R \hookrightarrow R'$ an injective | |
integral morphism, then the maximal ideal of $R$ is the inverse image of | |
something in $\spec R'$. | |
\end{quote} | |
Assume $R$ is local with maximal ideal $\mathfrak{p}$. We want to find a prime ideal $\mathfrak{q} \subset R'$ such that | |
$\mathfrak{p} = \phi^{-1}(\mathfrak{q})$. Since $\mathfrak{p}$ is already | |
maximal, it will suffice to show that $\mathfrak{p} \subset | |
\phi^{-1}(\mathfrak{q})$. In particular, we need to show that there is a prime | |
ideal $\mathfrak{q}$ such that | |
\( \mathfrak{p} R' \subset \mathfrak{q}. \) | |
The pull-back of this will be $\mathfrak{p}$. | |
If $\mathfrak{p}R' \neq R'$, then | |
$\mathfrak{q}$ exists, since every proper ideal of a ring is contained in a | |
maximal ideal. We will in fact show | |
\begin{equation} \label{thingdoesn'tgenerate} \mathfrak{p} R' \neq R', | |
\end{equation} | |
or that $\mathfrak{p}$ does not generate the unit ideal in $R'$. | |
If we prove \eqref{thingdoesn'tgenerate}, we will thus be able to find our | |
$\mathfrak{q}$, and we will be done. | |
Suppose the | |
contrary, i.e. $\mathfrak{p}R' = R'$. We will derive a contradiction using | |
Nakayama's lemma (\cref{nakayama}). | |
Right now, we cannot apply Nakayama's lemma directly because $R'$ is not a | |
finite $R$-module. The idea is that we will ``descend'' the ``evidence'' that | |
\eqref{thingdoesn'tgenerate} fails to a small subalgebra of $R'$, and then | |
obtain a contradiction. | |
To do this, note that $1 \in \mathfrak{p}R'$, and we can write | |
\[ 1 = \sum x_i \phi(y_i) \] | |
where $x_i \in R', y_i \in \mathfrak{p}$. | |
This is the ``evidence'' that \eqref{thingdoesn'tgenerate} fails, and it | |
involves only a finite amount of data. | |
Let $R''$ be the subalgebra of $R'$ generated by $\phi(R)$ and the $x_i$. Then | |
$R'' \subset R'$ and is finitely generated as an $R$-algebra, because it is generated by | |
the $x_i$. However, $R''$ is integral over $R$ and thus finitely generated as an | |
$R$-module, by \cref{finitelygeneratedintegral}. This | |
is where integrality comes in. | |
So $R''$ is a finitely generated $R$-module. Also, the expression | |
$1 = \sum x_i \phi(y_i)$ shows that $\mathfrak{p}R'' = R''$. However, this | |
contradicts Nakayama's lemma. That brings the contradiction, showing that | |
$\mathfrak{p}$ cannot generate $(1)$ in $R'$, and proving the surjectivity | |
part of lying over theorem. | |
Finally, we need to show that if $\phi: R \to R'$ is \emph{any} integral | |
morphism, then $\spec R' \to \spec R$ is a closed map. Let $X = V(I) $ be a | |
closed subset of $\spec R'$. Then the image of $X$ in $\spec R$ is the image | |
of the map | |
\[ \spec R'/I \to \spec R \] | |
obtained from the morphism $R \to R' \to R'/I$, which is integral; thus we are | |
reduced to showing that any integral morphism $\phi$ has closed image on the | |
$\spec$. | |
Thus we are reduced to $X = \spec R'$, if we throw out $R'$ and replace it by | |
$R'/I$. | |
In other words, we must prove the following statement. Let $\phi: R \to R'$ be | |
an integral morphism; then the image of $\spec R'$ in $\spec R$ is closed. | |
But, quotienting by $\ker \phi$ and taking the map $R/\ker \phi \to R'$, we | |
may reduce to the case of $\phi$ injective; however, then this follows from | |
the surjectivity result already proved. | |
\end{proof} | |
In general, there will be \emph{many} lifts of a given prime ideal. | |
Consider for instance the inclusion $\mathbb{Z} \subset \mathbb{Z}[i]$. | |
Then the prime ideal $(5) \in \spec \mathbb{Z}$ can be lifted either to $(2+i) | |
\in \spec \mathbb{Z}[i]$ | |
or $(2-i) \in \spec \mathbb{Z}[i]$. | |
These are distinct prime ideals: $\frac{2+i}{2-i} \notin \mathbb{Z}[i]$. But | |
note that any element of $\mathbb{Z}$ divisible by $2+i$ is automatically | |
divisible by its conjugate $2-i$, and consequently by their product $5$ | |
(because $\mathbb{Z}[i]$ is a UFD, being a euclidean domain). | |
Nonetheless, the different lifts are incomparable. | |
\begin{proposition} \label{incomparableintegral} | |
Let $\phi: R \to R'$ be an integral morphism. Then given $\mathfrak{p} \in | |
\spec R$, there are no inclusions among the elements $\mathfrak{q} \in \spec | |
R'$ lifting $\mathfrak{p}$. | |
\end{proposition} | |
In other words, if $\mathfrak{q}, \mathfrak{q}' \in \spec R'$ lift | |
$\mathfrak{p}$, then $\mathfrak{q} \not\subset \mathfrak{q}'$. | |
\begin{proof} | |
We will give a ``slick'' proof by various reductions. | |
Note that the operations of localization and quotienting only shrink the | |
$\spec$'s: they do not ``merge'' heretofore distinct prime ideals into one. | |
Thus, by quotienting $R$ by $\mathfrak{p}$, we may assume $R$ is a | |
\emph{domain} and that $\mathfrak{p} = 0$. | |
Suppose we had two primes $\mathfrak{q} \subsetneq \mathfrak{q}'$ of $R'$ | |
lifting $(0) \in \spec R$. | |
Quotienting $R'$ by $\mathfrak{q}$, we may assume that $\mathfrak{q} = 0$. | |
We could even assume $R \subset R'$, by quotienting by the kernel of $\phi$. | |
The next lemma thus completes the proof, because it shows that $\mathfrak{q}' | |
= 0$, contradiction. | |
\end{proof} | |
\begin{lemma} | |
Let $R \subset R'$ be an inclusion of integral domains, which is an integral | |
morphism. If $\mathfrak{q} \in \spec R'$ is a nonzero prime ideal, then | |
$\mathfrak{q} \cap R$ is nonzero. | |
\end{lemma} | |
\begin{proof} | |
Let $x \in \mathfrak{q}'$ be nonzero. There is an equation | |
\[ x^n + r_1 x^{n-1} + \dots + r_n = 0, \quad r_i \in R, \] | |
that $x$ satisfies, by assumption. | |
Here we can assume $r_n \neq 0$; then $r_n \in \mathfrak{q}' \cap R$ by | |
inspection, though. So this intersection is nonzero. | |
\end{proof} | |
\begin{corollary} | |
Let $R \subset R'$ be an inclusion of integral domains, such that $R'$ is | |
integral over $R$. Then if one of $R, R'$ is a field, so is the other. | |
\end{corollary} | |
\begin{proof} | |
Indeed, $\spec R' \to \spec R$ is surjective by \cref{lyingover}: so if $\spec | |
R'$ has one element (i.e., $R'$ is a field), the same holds for $\spec R$ | |
(i.e., $R$ is a field). Conversely, suppose $R$ a field. | |
Then any two prime ideals in $\spec R'$ pull back to the same element of $\spec | |
R$. So, by \cref{incomparableintegral}, there can be no inclusions among the prime ideals of $\spec | |
R'$. But $R'$ is a domain, so it must then be a field. | |
\end{proof} | |
\begin{exercise} Let $k$ be a field. | |
Show that $k[\mathbb{Q}_{\geq 0}]$ is integral over the polynomial ring $k[T]$. | |
Although this is a \emph{huge} extension, the prime ideal $(T)$ lifts in only | |
one way to $\spec k[\mathbb{Q}_{\geq 0}]$. | |
\end{exercise} | |
\begin{exercise} | |
Suppose $A \subset B$ is an inclusion of rings over a field of characteristic | |
$p$. Suppose $B^p \subset A$, so that $B/A$ is integral in a very strong sense. | |
Show that the map $\spec B \to \spec A$ is a \emph{homeomorphism.} | |
\end{exercise} | |
\subsection{Going up} | |
Let $R \subset R'$ be an inclusion of rings with $R'$ integral over $R$. We saw in the | |
lying over theorem (\cref{lyingover}) that any prime $\mathfrak{p} \in \spec R$ | |
has a prime $\mathfrak{q} \in \spec R'$ ``lying over'' $\mathfrak{p}$, i.e. | |
such that $R \cap \mathfrak{q} = \mathfrak{p}$. | |
We now want to show that we can lift finite \emph{inclusions} of primes to $R'$. | |
\begin{proposition}[Going up] | |
Let $R \subset R'$ be an integral inclusion of rings. | |
Suppose $\mathfrak{p}_1 \subset \mathfrak{p}_2 \subset \dots \subset | |
\mathfrak{p}_n \subset R$ is a | |
finite ascending chain of prime ideals in $R$. | |
Then there is an ascending chain $\mathfrak{q}_1 \subset \mathfrak{q}_2 \subset | |
\dots \subset \mathfrak{q}_n$ in $\spec R'$ lifting this chain. | |
Moreover, $\mathfrak{q}_1$ can be chosen arbitrarily so as to lift | |
$\mathfrak{p}_1$. | |
\end{proposition} | |
\begin{proof} | |
By induction and lying over (\cref{lyingover}), it suffices to show: | |
\begin{quote} | |
Let $\mathfrak{p}_1 \subset \mathfrak{p}_2$ be an inclusion of primes in $\spec | |
R$. Let $\mathfrak{q}_1 \in \spec R'$ lift $\mathfrak{p}_1$. Then there is | |
$\mathfrak{q}_2 \in \spec R'$, which satisfies the dual conditions of lifting | |
$\mathfrak{p}_2$ and containing $\mathfrak{q}_1$. | |
\end{quote} | |
To show that this is true, we apply \cref{lyingover} to the inclusion | |
$R/\mathfrak{p}_1 \hookrightarrow R'/\mathfrak{q}_1$. There is an element of | |
$\spec R'/\mathfrak{q}_1$ lifting $\mathfrak{p}_2/\mathfrak{p}_1$; the | |
corresponding element of $\spec R'$ will do for $\mathfrak{q}_2$. | |
\end{proof} | |
\section{Valuation rings} | |
A valuation ring is a special type of local ring. Its distinguishing | |
characteristic is that divisibility is a ``total preorder.'' That is, two | |
elements of the quotient field are never incompatible under divisibility. | |
We shall see in this section that integrality can be detected using | |
valuation rings only. | |
Geometrically, the valuation ring is something like a local piece of a smooth | |
curve. In fact, in algebraic geometry, a more compelling reason to study | |
valuation rings is provided by the valuative criteria for separatedness and | |
properness (cf. \cite{EGA} or \cite{Ha77}). One key observation about | |
valuation rings that leads the last results is that any local domain can be | |
``dominated'' by a valuation ring with the same quotient field (i.e. mapped | |
into a valuation ring via local | |
homomorphism), but valuation rings are the maximal elements in this relation | |
of domination. | |
\subsection{Definition} | |
\begin{definition} | |
A \textbf{valuation ring} is a domain $R$ such that for every pair of elements | |
$a,b \in R$, either $a \mid b$ or $b \mid a$. | |
\end{definition} | |
\begin{example} | |
$\mathbb{Z}$ is not a valuation ring. It is neither true that 2 divides 3 | |
nor that 3 divides 2. | |
\end{example} | |
\begin{example} | |
$\mathbb{Z}_{(p)}$, which is the set of all fractions of the form $a/b \in | |
\mathbb{Q}$ where $p \nmid b$, is a valuation ring. To check whether $a/b$ | |
divides $a'/b'$ or vice versa, one just has to check which is divisible by | |
the larger power of $p$. | |
\end{example} | |
\begin{proposition} | |
Let $R$ be a domain with quotient field $K$. Then $R$ is a valuation ring if | |
and only if for every $x \in K$, either $x$ or $x^{-1}$ lies in $R$. | |
\end{proposition} | |
\begin{proof} Indeed, if $x=a/b , \ a,b \in R$, then either $a \mid | |
b$ or $b \mid a$, so either $x$ or $x^{-1} \in R$. This condition is equivalent | |
to $R$'s being a valuation ring. | |
\end{proof} | |
\subsection{Valuations} | |
The reason for the name ``valuation ring'' is provided by the next definition. | |
As we shall see, any valuation ring comes from a ``valuation.'' | |
By definition, an \emph{ordered abelian group} is an abelian group $A$ | |
together with a set of \emph{positive elements} $A_+ \subset A$. This set is | |
required to be closed under addition and satisfy the property that if $x \in | |
A$, then precisely one of the following is true: $x \in A_+$, $-x \in A_+$, | |
and $x = 0$. This allows one to define an ordering $<$ on $A$ by writing $x<y$ | |
if $y-x \in A_+$. | |
Given $A$, we often formally adjoin an element $\infty$ which is bigger than | |
every element in $A$. | |
\begin{definition} | |
Let $K$ be a field. A \textbf{valuation} on $K$ is a map $v: K \to A \cup | |
\left\{\infty\right\}$ for some | |
ordered abelian group $A$ satisfying: | |
\begin{enumerate} | |
\item $v(0) = \infty$ and $v(K^*) \subset A$. | |
\item For $x,y \in K^*$, $v(xy) = v(x) + v(y)$. That is, $v|_{K^*}$ is | |
a homomorphism. | |
\item For $x,y \in K$, $v(x+y) \geq \min (v(x), v(y))$. | |
\end{enumerate} | |
\end{definition} | |
Suppose that $K$ is a field and $v: K \to A \cup \left\{\infty\right\}$ is a | |
valuation (i.e. $v(0) = \infty$). Define $R = \left\{x \in K: v(x) \geq | |
0\right\}$. | |
\begin{proposition} | |
$R$ as just defined is a valuation ring. | |
\end{proposition} | |
\begin{proof} First, we prove that $R$ is a ring. | |
$R$ is closed under addition and multiplication by the two conditions | |
\[ v(xy) = v(x) + v(y) \] | |
and | |
\[ v(x+y) \geq \min v(x), v(y) , \] | |
so if $x,y \in R$, then $x+y, xy$ have nonnegative valuations. | |
Note that $0 \in R$ because $v(0) = \infty$. Also $v(1) = 0$ since $v: K^* | |
\to A$ | |
is a homomorphism. So $1 \in R$ too. | |
Finally, $-1 \in R$ because $v(-1) =0$ since $A$ is totally ordered. It | |
follows that $R$ is also a group. | |
Let us now show that $R$ is a valuation ring. If $x \in K^*$, either $v(x) \geq | |
0$ or $v(x^{-1}) \geq 0$ since $A$ is totally ordered.\footnote{Otherwise $0 | |
=v(x)+v(x^{-1}) < 0$, contradiction.} So either $x, x^{-1} \in R$. | |
\end{proof} | |
In particular, the set of elements with nonnegative valuation is a valuation | |
ring. | |
The converse also holds. Whenever you have a valuation ring, it comes about in | |
this manner. | |
\begin{proposition} | |
Let $R$ be a valuation ring with quotient field $K$. There is an ordered | |
abelian group | |
$A$ and a valuation $v: K^* \to A$ such that $R$ is the set of elements with | |
nonnegative valuation. | |
\end{proposition} | |
\begin{proof} | |
First, we construct $A$. In fact, it is the quotient of $K^*$ by the subgroup | |
of units $R^*$ of $R$. | |
We define an ordering by saying that $x \leq y$ if $y/x \in R$---this doesn't | |
depend on the representatives in $K^*$ chosen. Note that either $x \leq y$ or | |
$y \leq x$ must hold, since $R$ is a valuation ring. | |
The combination of $ x \leq y$ and $y \leq x$ implies that $x,y$ are equivalent | |
classes. | |
The nonnegative elements in this group are those whose representatives in $K^*$ | |
belong to $R$. | |
It is easy to see that $K^*/R^*$ in this way is a totally ordered abelian | |
group with | |
the image of 1 as the unit. The | |
reduction map $K^* \to K^*/R^*$ defines a valuation whose corresponding ring | |
is just $R$. We have omitted some details; for instance, it should be checked | |
that the valuation of $x+y$ is at least the minimum of $v(x), v(y)$. | |
\end{proof} | |
To summarize: | |
\begin{quote} | |
Every valuation ring $R$ determines a valuation $v$ from the fraction field of | |
$R$ into $A \cup \left\{\infty\right\}$ for $A$ a totally ordered abelian group | |
such that $R$ is just the set of elements of $K$ with nonnegative valuation. As | |
long as we require that $v: K^* \to A$ is surjective, then $A$ is uniquely | |
determined as well. | |
\end{quote} | |
\begin{definition} | |
A valuation ring $R$ is \textbf{discrete} if we can choose $A$ to be | |
$\mathbb{Z}$. | |
\end{definition} | |
\begin{example} | |
$\mathbb{Z}_{(p)}$ is a discrete valuation ring. | |
\end{example} | |
The notion of a valuation ring is a useful one. | |
\subsection{General remarks} | |
Let $R$ be a commutative ring. Then $\spec R$ is the set of primes of $R$, | |
equipped | |
with a certain topology. The space $\spec R$ is almost never Hausdorff. It is | |
almost always a bad idea to apply the familiar ideas from elementary topology | |
(e.g. the fundamental group) to $\spec R$. Nonetheless, it has some other nice | |
features that substitute for its non-Hausdorffness. | |
For instance, if $R = \mathbb{C}[x,y]$, then $\spec R$ corresponds to | |
$\mathbb{C}^2$ with some additional nonclosed points. The injection of | |
$\mathbb{C}^2$ with its usual topology into $\spec R$ is continuous. While in | |
$\spec R$ you don't want to think of continuous paths, you can in | |
$\mathbb{C}^2$. | |
Suppose you had two points $x,y \in \mathbb{C}^2$ and their images in $\spec | |
R$. Algebraically, you can still think about algebraic curves passing | |
through $x,y$. | |
This is a subset of $x,y$ defined by a single polynomial equation. | |
This curve will have what's called a ``generic point,'' since the ideal | |
generated by this curve will be a prime ideal. | |
The closure of this generic point will be precisely this algebraic | |
curve---including $x,y$. | |
\begin{remark} | |
If $ \mathfrak{p}, \mathfrak{p}' \in \spec R$, then | |
\[ \mathfrak{p}' \in \overline{\left\{\mathfrak{p}\right\}} \] | |
iff | |
\[ \mathfrak{p}' \supset \mathfrak{p}. \] | |
Why is this? Well, the closure of $\left\{\mathfrak{p}\right\}$ is just | |
$V(\mathfrak{p})$, since this is the smallest closed subset of $\spec R$ | |
containing $\mathfrak{p}$. | |
\end{remark} | |
The point of this discussion is that instead of paths, one can transmit | |
information from point to point in $\spec R$ by having one point be in a | |
closure of another. | |
However, we will show that this relation is contained by the theory of | |
valuation rings. | |
\begin{theorem} | |
Let $R$ be a domain containing a prime ideal $\mathfrak{p}$. Let $K$ be the | |
fraction field of $R$. | |
Then there is a valuation $v$ on $K$ defining a valuation ring $R' \subset | |
K$ such that | |
\begin{enumerate} | |
\item $R \subset R'$. | |
\item $\mathfrak{p} = \left\{x \in R: v(x) > 0\right\}$. | |
\end{enumerate} | |
\end{theorem} | |
Let us motivate this by the remark: | |
\begin{remark} | |
A valuation ring is automatically a local ring. A local ring is a ring where | |
either $x, 1-x$ is invertible for all $x$ in the ring. Let us show that this is | |
true for a valuation ring. | |
If $x $ belongs to a valuation ring $R$ with valuation $v$, it is invertible if | |
$v(x)=0$. So if $x, 1-x$ were both noninvertible, then both would have | |
positive valuation. However, that would imply that $v(1) \geq \min v(x), | |
v(1-x)$ is positive, contradiction. | |
\end{remark} | |
\begin{quote} | |
If $R'$ is any valuation ring (say defined by a valuation $v$), then $R'$ is | |
local with maximal ideal consisting of elements with positive valuation. | |
\end{quote} | |
The theorem above says that there's a good supply of valuation rings. | |
In particular, if $R$ is any domain, $\mathfrak{p} \subset R$ a prime ideal, | |
then we can choose a valuation ring $R' \supset R$ such that $\mathfrak{p}$ is | |
the intersection of the maximal ideal of $R'$ intersected with $R$. | |
So the map $\spec R' \to \spec R$ contains $\mathfrak{p}$. | |
\begin{proof} | |
Without loss of generality, replace $R$ by $R_{\mathfrak{p}}$, which is a local | |
ring with maximal ideal $\mathfrak{p}R_{\mathfrak{p}}$. The maximal ideal | |
intersects $R$ only in $\mathfrak{p}$. | |
So, we can assume without loss of generality that | |
\begin{enumerate} | |
\item $R$ is local. | |
\item $\mathfrak{p}$ is maximal. | |
\end{enumerate} | |
Let $P$ be the collection of all subrings $R' \subset K$ such that $R' \supset | |
R$ but $\mathfrak{p}R' \neq R'$. Then $P$ is a poset under inclusion. The | |
poset is nonempty, since $R \in P$. Every totally ordered chain in $P$ has an | |
upper bound. If you have a totally ordered subring of elements in $P$, then | |
you can take the union. | |
We invoke: | |
\begin{lemma} | |
Let $R_\alpha$ be a chain in $P$ and $R' = \bigcup R_\alpha$. Then $R' \in P$. | |
\end{lemma} | |
\begin{proof} | |
Indeed, it is easy to see that this is a subalgebra of $K$ containing $R$. The | |
thing to observe is that | |
\[ \mathfrak{p}R' = \bigcup_\alpha \mathfrak{p} R_\alpha ;\] | |
since by assumption, $1 \notin \mathfrak{p}R_\alpha$ (because each $R_\alpha | |
\in P$), $1 \notin \mathfrak{p}R'$. In particular, $R' \notin P$. | |
\end{proof} | |
By the lemma, Zorn's lemma to the poset $P$. In particular, $P$ has a maximal | |
element $R'$. By construction, $R'$ is some subalgebra of $K$ and | |
$\mathfrak{p}R' \neq R'$. Also, $R'$ is maximal with respect to these | |
properties. | |
We show first that $R'$ is local, with maximal ideal $\mathfrak{m}$ satisfying | |
\[ \mathfrak{m} \cap R = \mathfrak{p}. \] | |
The second part is evident from locality of $R'$, since $\mathfrak{m} $ | |
must contain | |
the proper ideal $\mathfrak{p}R'$, and $\mathfrak{p} \subset R$ is a maximal | |
ideal. | |
Suppose that $x \in R'$; we show that either $x, 1-x$ belongs to $R'^*$ (i.e. | |
is invertible). Take the ring $R'[x^{-1}]$. If $x$ is noninvertible, this | |
properly contains $R'$. By maximality, it follows that $\mathfrak{p}R'[x^{-1}] | |
= R'[x^{-1}]$. | |
And we're out of time. We'll pick this up on Monday. | |
\end{proof} | |
Let us set a goal. | |
First, recall the notion introduced last time. A \textbf{valuation ring} is a | |
domain $R$ where for all $x$ in the fraction field of $R$, either $x$ or | |
$x^{-1}$ lies in $R$. We saw that if $R$ is a valuation ring, then $R$ is | |
local. That is, there is a unique maximal ideal $\mathfrak{m} \subset R$, | |
automatically prime. Moreover, the zero ideal $(0)$ is prime, as $R$ is a | |
domain. So if you look at the spectrum $\spec R$ of a valuation ring $R$, there | |
is a unique closed point $\mathfrak{m}$, and a unique generic point | |
$(0)$. There might be some other prime ideals in $\spec R$; this depends on | |
where the additional valuation lives. | |
\begin{example} | |
Suppose the valuation defining the valuation ring $R$ takes values in | |
$\mathbb{R}$. Then the only primes are $\mathfrak{m}$ and zero. | |
\end{example} | |
Let $R$ now be any ring, with $\spec R$ containing prime ideals | |
$\mathfrak{p} \subset \mathfrak{q}$. In particular, $\mathfrak{q}$ lies in | |
the closure of $\mathfrak{p}$. | |
As we will see, this implies that there is a map | |
\[ \phi: R \to R' \] | |
such that $\mathfrak{p} = \phi^{-1}(0)$ and $\mathfrak{q} = | |
\phi^{-1}(\mathfrak{m})$, where $\mathfrak{m}$ is the maximal ideal of $R'$. | |
This statement says that the relation of closure in $\spec R$ is always | |
controlled by valuation rings. | |
In yet another phrasing, in the map | |
\[ \spec R' \to \spec R \] | |
the closed point goes to $\mathfrak{q}$ and the generic point to | |
$\mathfrak{p}$. This is our eventual goal. | |
To carry out this goal, we need some more elementary facts. Let us discuss | |
things that don't have any obvious relation to it. | |
\subsection{Back to the goal} Now we return to the goal of the lecture. Again, | |
$R$ | |
was any ring, and we had primes $\mathfrak{p} \subset \mathfrak{q} \subset | |
R$. We | |
wanted a valuation ring $R'$ and a map $\phi: R \to R'$ such that zero pulled | |
back to $\mathfrak{p}$ and the maximal ideal pulled back to $\mathfrak{q}$. | |
What does it mean for $\mathfrak{p}$ to be the inverse image of $(0) \subset | |
R'$? This means that $\mathfrak{p} = \ker \phi$. So we get an injection | |
\[ R/\mathfrak{p} \rightarrowtail R'. \] | |
We will let $R'$ be a subring of the quotient field $K$ of the domain | |
$R/\mathfrak{p}$. Of course, this subring will contain $R/\mathfrak{p}$. | |
In this case, we will get a map $R \to R'$ such that the pull-back of zero is | |
$\mathfrak{p}$. What we want, further, to be true is that $R'$ is a valuation | |
ring and the pull-back of the maximal ideal is $\mathfrak{q}$. | |
This is starting to look at the problem we discussed last time. | |
Namely, let's throw out $R$, and replace it with $R/\mathfrak{p}$. | |
Moreover, we can replace $R$ with $R_{\mathfrak{q}}$ and assume that $R$ is | |
local with maximal ideal $\mathfrak{q}$. | |
What we need to show is that a valuation ring $R' $ contained in the fraction | |
field of $R$, containing $R$, such that the intersection of the maximal | |
ideal of | |
$R'$ with $R$ is equal to $\mathfrak{q} \subset R$. | |
If we do this, then we will have accomplished our goal. | |
\begin{lemma} | |
Let $R$ be a local domain. Then there is a valuation subring $R'$ of the | |
quotient | |
field of $R$ that \emph{dominates} $R$, i.e .the map $R \to R'$ is a | |
\emph{local} homomorphism. | |
\end{lemma} | |
Let's find $R'$ now. | |
Choose $R'$ maximal such that $\mathfrak{q} R' \neq R'$. Such a ring exists, by | |
Zorn's lemma. We gave this argument at the end last time. | |
\begin{lemma} | |
$R'$ as described is local. | |
\end{lemma} | |
\begin{proof} | |
Look at $\mathfrak{q}R' \subset R'$; it is a proper subset, too, by assumption. | |
In particular, $\mathfrak{q}R'$ is contained in some maximal ideal | |
$\mathfrak{m}\subset R'$. Replace $R'$ by $R'' = R'_{\mathfrak{m}}$. | |
Note that | |
\[ R' \subset R'' \] | |
and | |
\[ \mathfrak{q}R'' \neq R'' \] | |
because $\mathfrak{m}R'' \neq R''$. But $R'$ is maximal, so $R' = R''$, and | |
$R''$ is a local ring. So $R'$ is a local ring. | |
\end{proof} | |
Let $\mathfrak{m}$ be the maximal ideal of $R'$. Then $\mathfrak{m} \supset | |
\mathfrak{q}R$, so $\mathfrak{m} \cap R = \mathfrak{q}$. | |
All that is left to prove now is that $R'$ is a valuation ring. | |
\begin{lemma} | |
$R'$ is integrally closed. | |
\end{lemma} | |
\begin{proof} | |
Let $R''$ be its integral closure. Then $\mathfrak{m} R'' \neq R''$ by lying | |
over, since $\mathfrak{m}$ (the maximal ideal of $R'$) lifts up to $R''$. So | |
$R''$ satisfies | |
\[ \mathfrak{q}R'' \neq R'' \] | |
and by maximality, we have $R'' = R'$. | |
\end{proof} | |
To summarize, we know that $R'$ is a local, integrally closed subring of the | |
quotient field of $R$, such that the maximal ideal of $R'$ pulls back to | |
$\mathfrak{q}$ in $R$. | |
All we now need is: | |
\begin{lemma} | |
$R'$ is a valuation ring. | |
\end{lemma} | |
\begin{proof} | |
Let $x$ lie in the fraction field. We must show that either $x$ or $x^{-1} \in | |
R'$. Say $x \notin R'$. This means by maximality of $R'$ that $R'' = | |
R'[x]$ satisfies | |
\[ \mathfrak{q}R'' = R''. \] | |
In particular, we can write | |
\[ 1 = \sum q_i x^i, \quad q_i \in \mathfrak{q}R' \subset R'. \] | |
This implies that | |
\[ (1-q_0) + \sum_{i > 0} -q_i x^i = 0. \] | |
But $1-q_0$ is invertible in $R'$, since $R'$ is local. We can divide by the | |
highest power of $x$: | |
\[ x^{-N} + \sum_{i>0} \frac{-q_i}{1-q_0} x^{-N+i} = 0. \] | |
In particular, $1/x$ is integral over $R'$; this implies that $1/x \in | |
R'$ since | |
$R'$ is integrally closed and $q_0$ is a nonunit. So | |
$R'$ is a valuation ring. | |
\end{proof} | |
We can state the result formally. | |
\begin{theorem} | |
Let $R$ be a ring, $\mathfrak{p} \subset \mathfrak{q}$ prime ideals. Then there | |
is a homomorphism $\phi: R \to R'$ into a valuation ring $R'$ with maximal | |
ideal | |
$\mathfrak{m}$ such that | |
\[ \phi^{-1}(0) = \mathfrak{p} \] | |
and | |
\[ \phi^{-1}(\mathfrak{m} ) = \mathfrak{q} .\] | |
\end{theorem} | |
There is a related fact which we now state. | |
\begin{theorem} | |
Let $R$ be any domain. Then the integral closure of $R$ in the quotient field | |
$K$ is the intersection | |
\[ \bigcap R_{\alpha} \] | |
of all valuation rings $R_{\alpha} \subset K$ containing $R$. | |
\end{theorem} | |
So an element of the quotient field is integral over $R$ if and only if its | |
valuation is nonnegative at every valuation which is nonnegative on $R$. | |
\begin{proof} | |
The $\subset$ argument is easy, because one can check that a valuation ring is | |
integrally closed. (Exercise.) | |
The interesting direction is to assume that $v(x) \geq 0$ for all $v$ | |
nonnegative | |
on $R$. | |
Let us suppose $x$ is nonintegral. Suppose $R' = R[1/x]$ and $I$ be the ideal | |
$(x^{-1}) \subset R'$. There are two cases: | |
\begin{enumerate} | |
\item $I = R'$. Then in the ring $R'$, $x^{-1} $ is invertible. In particular, | |
$x^{-1}P(x^{-1}) = 1$. Multiplying by a high power of $x$ shows that $x$ is | |
integral over $R$. Contradiction. | |
\item Suppose $I \subsetneq R'$. Then $I$ is contained in a maximal ideal | |
$\mathfrak{q} \subset R'$. There is a valuation subring $R'' \subset K$ , | |
containing $R'$, such that the corresponding valuation is positive on | |
$\mathfrak{q}$. In particular, this valuation is positive on $x^{-1}$, | |
so it is | |
negative on $x$, contradiction. | |
\end{enumerate} | |
\end{proof} | |
So the integral closure has this nice characterization via valuation rings. In | |
some sense, the proof that $\mathbb{Z}$ is integrally closed has the property | |
that every integrally closed ring is integrally closed for that reason: | |
it's the | |
common nonnegative locus for some valuations. | |
\section{The Hilbert Nullstellensatz} | |
The Nullstellensatz is the basic algebraic fact, which we have invoked in the | |
past to justify various examples, that connects the idea of | |
the $\spec$ of a ring to classical algebraic geometry. | |
\subsection{Statement and initial proof of the Nullstellensatz} | |
There are several ways in which the Nullstellensatz can be stated. Let us | |
start with the following very concrete version. | |
\begin{theorem} \label{nullstellensatzoverC} | |
All maximal ideals in the polynomial ring $R=\mathbb{C}[x_1, \dots, x_n]$ come | |
from points in $\mathbb{C}^n$. In other words, if $\mathfrak{m} \subset R$ is | |
maximal, then there exist $a_1, \dots, a_n \in \mathbb{C}$ such that | |
$\mathfrak{m} = (x_1 - a_1, \dots, x_n - a_n)$. | |
\end{theorem} | |
The maximal spectrum of $R=\mathbb{C}[x_1, \dots, x_n]$ is thus identified with | |
$\mathbb{C}^n$. | |
We shall now reduce \rref{nullstellensatzoverC} to an easier claim. | |
Let | |
$\mathfrak{m}\subset R$ be a maximal ideal. Then there is a map | |
\[ \mathbb{C} \to R \to R/\mathfrak{m} \] | |
where $R/\mathfrak{m}$ is thus a finitely generated $\mathbb{C}$-algebra, as | |
$R$ is. The ring $R/\mathfrak{m}$ is also a field by maximality. | |
We would like to show that $R/\mathfrak{m}$ is a finitely generated | |
$\mathbb{C}$-vector | |
space. This would imply that $R/\mathfrak{m}$ is integral over $\mathbb{C}$, | |
and there are no proper algebraic extensions of $\mathbb{C}$. Thus, if we | |
prove this, it will follow that the map $\mathbb{C} \to R/\mathfrak{m}$ is an | |
isomorphism. If $a_i \in \mathbb{C}$ ($1 \leq i \leq n$) is the image of | |
$x_i $ in $R/\mathfrak{m} = | |
\mathbb{C}$, it will follow that $(x_1 - a_1, \dots, x_n - a_n) \subset | |
\mathfrak{m}$, so $(x_1 - a_1, \dots, x_n - a_n)= | |
\mathfrak{m}$. | |
Consequently, the Nullstellensatz in this form would follow from the next | |
claim: | |
\begin{proposition} | |
Let $k$ be a field, $L/k$ an extension of fields. Suppose $L$ is a finitely | |
generated $k$-algebra. Then $L$ is a finite $k$-vector space. | |
\end{proposition} | |
This is what we will prove. | |
We start with an easy proof in the special case: | |
\begin{lemma} | |
Assume $k$ is uncountable (e.g. $\mathbb{C}$, the original case of interest). | |
Then the above proposition is true. | |
\end{lemma} | |
\begin{proof} | |
Since $L$ is a finitely generated $k$-algebra, it suffices to show that $L/k$ | |
is algebraic. | |
If not, there exists $x \in L$ which isn't algebraic over $k$. So $x$ satisfies | |
no nontrivial polynomials. | |
I claim now that the uncountably many elements $\frac{1}{x-\lambda}, \lambda | |
\in K$ are linearly | |
independent over $K$. This will be a contradiction as $L$ is a finitely | |
generated $k$-algebra, hence at most countably dimensional over $k$. (Note that | |
the polynomial ring is countably dimensional over $k$, and $L$ is a quotient.) | |
So let's prove this. Suppose not. Then there is a nontrivial linear dependence | |
\[ \sum \frac{c_i}{x - \lambda_i} = 0, \quad c_i, \lambda_i \in K. \] | |
Here the $\lambda_j$ are all distinct to make this nontrivial. Clearing | |
denominators, we find | |
\[ \sum_i c_i \prod_{j \neq i } (x- \lambda_j) = 0. \] | |
Without loss of generality, $c_1 \neq 0$. | |
This equality was in the field $L$. But $x$ is transcendental over $k$. So we | |
can think of this as a polynomial ring relation. | |
Since we can think of this as a relation in the polynomial ring, we see that | |
doing so, all but the $i =1$ term in the sum is divisible by $x - \lambda_1$ | |
as a polynomial. | |
It follows that, as polynomials in the indeterminate $x$, | |
\[ x - \lambda_1 \mid c_1 \prod_{j \neq 1} (x - \lambda_j). \] | |
This is a contradiction since all the $\lambda_i$ are distinct. | |
\end{proof} | |
This is kind of a strange proof, as it exploits the fact that $\mathbb{C}$ is | |
uncountable. | |
This shouldn't be relevant. | |
\subsection{The normalization lemma} | |
Let's now give a more algebraic proof. | |
We shall exploit the following highly useful fact in commutative algebra: | |
\begin{theorem}[Noether normalization lemma] Let $k$ be a field, and $R = | |
k[x_1, \dots, x_n]/\mathfrak{p}$ be a finitely generated domain over $k$ (where | |
$\mathfrak{p}$ is a prime ideal in the polynomial ring). | |
Then there exists a polynomial subalgebra $k[y_1, \dots, y_m] \subset R$ such | |
that $R$ is integral over $k[y_1, \dots, y_m]$. | |
\end{theorem} | |
Later we will see that $m$ is the \emph{dimension} of $R$. | |
There is a geometric picture here. Then $\spec R$ is some irreducible algebraic | |
variety in $k^n$ (plus some additional points), with a smaller dimension than | |
$n$ if $\mathfrak{p} \neq 0$. Then there exists a \emph{finite map} to $k^m$. | |
In particular, we can map surjectively $\spec R \to k^m$ which is integral. | |
The fibers are in fact finite, because integrality implies finite fibers. (We | |
have not actually proved this yet.) | |
How do we actually find such a finite projection? In fact, in characteristic | |
zero, we just take a | |
vector space projection $\mathbb{C}^n \to \mathbb{C}^m$. For a ``generic'' | |
projection onto a subspace of the appropriate dimension, the projection will | |
will do as our finite map. In characteristic $p$, this may not work. | |
\begin{proof} | |
First, note that $m$ is uniquely determined as the transcendence degree of the | |
quotient field of $R$ over $k$. | |
Among the variables $x_1, \dots, x_n \in R$ (which we think of as in $R$ by an | |
abuse of notation), choose a maximal subset which is algebraically independent. | |
This subset has no nontrivial polynomial relations. In particular, the ring | |
generated by that subset is just the polynomial ring on that subset. | |
We can permute these variables and assume that | |
$$\left\{x_1,\dots, x_m\right\}$$ is the maximal subset. In particular, $R$ | |
contains the \emph{polynomial ring} $k[x_1, \dots, x_m]$ and is generated by | |
the rest of the variables. The rest of the variables are not adjoined freely | |
though. | |
The strategy is as follows. We will implement finitely many changes of | |
variable so that $R$ becomes integral over $k[x_1, \dots, x_m]$. | |
The essential case is where $m=n-1$. Let us handle this. So we have | |
\[ R_0 = k[x_1, \dots, x_m] \subset R = R_0[x_n]/\mathfrak{p}. \] | |
Since $x_n$ is not algebraically independent, there is a nonzero polynomial | |
$f(x_1, \dots, x_m, x_n) \in \mathfrak{p}$. | |
We want $f$ to be monic in $x_n$. This will buy us integrality. A priori, this | |
might not be true. We will modify the coordinate system to arrange that, | |
though. Choose $N \gg 0$. Define for $1 \leq i \leq m$, | |
\[ x_i' = x_i + x_n^{N^i}. \] | |
Then the equation becomes: | |
\[ 0 = f(x_1, \dots, x_m, x_n) = f( \left\{x_i' - x_n^{N^i}\right\} , x_n). \] | |
Now $f(x_1, \dots, x_n, x_{n+1})$ looks like some sum | |
\[ \sum \lambda_{a_1 \dots b} x_1^{a_1} \dots x_m^{a_m} x_n^{b} , \quad | |
\lambda_{a_1 \dots b} \in k. \] | |
But $N$ is really really big. Let us expand this expression in the $x_i'$ and | |
pay attention to the largest power of $x_n$ we see. | |
We find that | |
\[ f(\left\{x_i' - x_n^{N_i}\right\},x_n) | |
\] | |
has the largest power of $x_n$ precisely where, in the expression for $f$, | |
$a_m$ is maximized first, then | |
$a_{m-1}, $ and so on. The largest exponent would have the form | |
\[ x_n^{a_m N^m + a_{m-1}N^{m-1} + \dots + b}. \] | |
We can't, however, get any exponents of $x_n$ in the expression | |
\( f(\left\{x_i' - x_n^{N_i}\right\},x_n)\) other than these. If $N$ is super | |
large, then all these exponents will be different from each other. | |
In particular, each power of $x_n$ appears precisely once in the expansion of | |
$f$. We see in particular that $x_n$ is integral over $x_1', \dots, x_n'$. | |
Thus each $x_i$ is as well. | |
So we find | |
\begin{quote} | |
$R$ is integral over $k[x_1', \dots, x_m']$. | |
\end{quote} | |
We have thus proved the normalization lemma in the codimension one case. What | |
about the general case? We repeat this. | |
Say we have | |
\[ k[x_1, \dots, x_m] \subset R. \] | |
Let $R'$ be the subring of $R$ generated by $x_{1}, \dots,x_m, x_{m+1}$. The | |
argument we just gave implies that we can choose $x_1', \dots, x_m'$ such that | |
$R'$ is integral over $k[x_1', \dots, x_m']$, and the $x_i'$ are | |
algebraically independent. | |
We know in fact that $R' = k[x_1', \dots, x_m', x_{m+1}]$. | |
Let us try repeating the argument while thinking about $x_{m+2}$. Let $R'' = | |
k[x_1', \dots, x_m', x_{m+2}]$ modulo whatever relations that $x_{m+2}$ has to | |
satisfy. So this is a subring of $R$. The same argument shows that we can | |
change variables such that $x_1'', \dots, x_m''$ are algebraically independent | |
and $R''$ is integral over $k[x_1'', \dots, x_m'']$. We have furthermore that | |
$k[x_1'', \dots, x_m'', x_{m+2}] = R''$. | |
Having done this, let us give the argument where $m=n-2$. You will then see | |
how to do the general case. Then I claim that: | |
\begin{quote} | |
$R$ is integral over $k[x_1'', \dots, x_m'']$. | |
\end{quote} | |
For this, we need to check that $x_{m+1}, x_{m+2}$ are integral (because these | |
together with the $x''_i$ generate $R''[x_{m+2}][x_{m+2}]=R$. | |
But $x_{m+2}$ is integral over this by construction. The integral closure of | |
$k[x_1'', \dots, x_{m}'']$ in $R$ thus contains | |
\[ k[x_1'', \dots, x''_m, x_{m+2}] = R''. \] | |
However, $R''$ contains the elements $x_1', \dots, x_m'$. But by construction, | |
$x_{m+1}$ is integral over the $x_1', \dots, x_m'$. The integral closure of | |
$k[x_1'', \dots, x_{m}'']$ must contain $x_{m+2}$. This completes the proof in | |
the case $m=n-2$. The general case is similar; we just make several changes of | |
variables, successively. | |
\end{proof} | |
\subsection{Back to the Nullstellensatz} | |
Consider a finitely generated $k$-algebra $R$ which is a field. We need to | |
show that $R$ is a | |
finite $k$-module. This will prove the proposition. | |
Well, note that $R$ is integral over a polynomial ring $k[x_1, \dots, x_m]$ for | |
some $m$. | |
If $m > 0$, then this polynomial ring has more than one prime. | |
For instance, $(0)$ and $(x_1, \dots, x_m)$. But these must lift to primes in | |
$R$. Indeed, we have seen that whenever you have an integral extension, the | |
induced map on spectra is surjective. So | |
\[ \spec R \to \spec k[x_1, \dots, x_m] \] | |
is surjective. If $R$ is a field, this means $\spec k[x_1, \dots, x_m]$ has one | |
point and $m=0$. So $R$ is integral over $k$, thus algebraic. This implies that | |
$R$ is finite as it is finitely generated. This proves one version of the | |
Nullstellensatz. | |
Another version of the Nullstellensatz, which is | |
more precise, says: | |
\begin{theorem} \label{gennullstellensatz} | |
Let $I \subset \mathbb{C}[x_1, \dots, x_n]$. Let $V \subset \mathbb{C}^n$ be | |
the subset of $\mathbb{C}^n$ defined by the ideal $I$ (i.e. the zero locus of | |
$I$). | |
Then $\rad(I)$ is precisely the collection of $f$ such that $f|_V = 0$. In | |
particular, | |
\[ \rad(I) = \bigcap_{\mathfrak{m} \supset I, \mathfrak{m} \ | |
\mathrm{maximal}} \mathfrak{m}. \] | |
\end{theorem} | |
In particular, there is a bijection between radical ideals and algebraic | |
subsets of $\mathbb{C}^n$. | |
The last form of the theorem, which follows from the expression of maximal | |
ideals in the polynomial ring, is very similar to the result | |
\[ \rad(I) = \bigcap_{\mathfrak{p} \supset I, \mathfrak{p} \ | |
\mathrm{prime}} \mathfrak{p}, \] | |
true in any commutative ring. However, this general result is not necessarily | |
true. | |
\begin{example} | |
The intersection of all primes in a DVR is zero, but the intersection of all | |
maximals is nonzero. | |
\end{example} | |
\begin{proof}[Proof of \cref{gennullstellensatz}] | |
It now suffices to show that for every $\mathfrak{p} \subset \mathbb{C}[x_1, | |
\dots, x_n]$ prime, we have | |
\[ \mathfrak{p} = \bigcap_{\mathfrak{m} \supset I \ \mathrm{maximal}} | |
\mathfrak{m} \] | |
since every radical ideal is an intersection of primes. | |
Let $R = \mathbb{C}[x_1, \dots, | |
x_n]/\mathfrak{p}$. This is a domain finitely generated over $\mathbb{C}$. We | |
want to show that the intersection of maximal ideals in $R$ is zero. This is | |
equivalent to the above displayed equality. | |
So fix $f \in R - \left\{0\right\}$. Let $R'$ be the localization $R'= R_f$. Then $R'$ is also an | |
integral domain, finitely generated over $\mathbb{C}$. $R'$ has a maximal | |
ideal $\mathfrak{m}$ (which a priori could be zero). If | |
we look at the map $R' \to R'/\mathfrak{m}$, we get a map into a field | |
finitely generated | |
over $\mathbb{C}$, which is thus $\mathbb{C}$. | |
The composite map | |
\[ R \to R' \to R'/\mathfrak{m} \] | |
is just given by an $n$-tuple of complex numbers, i.e. to a point in | |
$\mathbb{C}^n$ which is even in $V$ as it is a map out of $R$. This corresponds | |
to a maximal ideal in $R$. | |
This maximal ideal does not contain $f$ by construction. | |
\end{proof} | |
\begin{exercise} | |
Prove the following result, known as ``Zariski's lemma'' (which easily implies | |
the Nullstellensatz): if $k$ is a field, $k'$ a field extension of $k$ which | |
is a finitely generated $k$-\emph{algebra}, then $k'$ is finite algebraic over | |
$k$. Use the following argument of McCabe (in \cite{Mc76}): | |
\begin{enumerate} | |
\item $k'$ contains a subring $S$ of the form $S= k[ x_1, \dots, x_t]$ where | |
the $x_1, \dots, x_t$ are algebraically independent over $k$, and $k'$ is | |
algebraic over the quotient field of $S$ (which is a polynomial ring). | |
\item If $k'$ is not algebraic over $k$, then $S \neq k$ is not a field. | |
\item Show that there is $y \in S$ such that $k'$ is integral over $S_y$. | |
Deduce that $S_y$ is a field. | |
\item Since $\spec (S_y) = \left\{0\right\}$, argue that $y$ lies in every | |
non-zero prime ideal of $\spec S$. Conclude that $1+y \in k$, and $S$ is a | |
field---contradiction. | |
\end{enumerate} | |
\end{exercise} | |
\subsection{A little affine algebraic geometry} | |
In what follows, let $k$ be algebraically closed, and let $A$ be a finitely generated $k$-algebra. Recall that $\Specm A$ denotes the set of maximal ideals in $A$. Consider the natural $k$-algebra structure on $\mathrm{Funct}(\Specm A, k)$. We have a map | |
$$A \rightarrow \mathrm{Funct}(\Specm A, k)$$ | |
which comes from the Weak Nullstellensatz as follows. Maximal ideals $\mathfrak{m}\subset A$ are in bijection with maps $\varphi_\mathfrak{m}:A\rightarrow k$ where $\ker(\varphi_\mathfrak{m})=\mathfrak{m}$, so we define $a\longmapsto [\mathfrak{m}\longmapsto \varphi_\mathfrak{m}(a)]$. If $A$ is reduced, then this map is injective because if $a\in A$ maps to the zero function, then $a\in \cap\, \mathfrak{m}$ $\rightarrow$ $a$ is nilpotent $\rightarrow$ $a=0$.\\ | |
\begin{definition} A function $f\in \mathrm{Funct}(\Specm A,k)$ is called {\bf algebraic} if it is in the image of $A$ under the above map. (Alternate words for this are {\bf polynomial} and {\bf regular}.) \end{definition} | |
Let $A$ and $B$ be finitely generated $k$-algebras and $\phi:A\rightarrow B$ a homomorphism. This yields a map $\Phi:\Specm B\rightarrow \Specm A$ given by taking pre-images. | |
\begin{definition} A map $\Phi:\Specm B\rightarrow \Specm A$ is called {\bf algebraic} if it comes from a homomorphism $\phi$ as above.\end{definition} | |
To demonstrate how these definitions relate to one another we have the following proposition. | |
\begin{proposition} A map $\Phi:\Specm B\rightarrow \Specm A$ is algebraic if and only if for any algebraic function $f\in \mathrm{Funct}(\Specm A,k)$, the pullback $f\circ \Phi\in \mathrm{Funct}(\Specm B,k)$ is algebraic.\end{proposition} | |
\begin{proof} | |
Suppose that $\Phi$ is algebraic. It suffices to check that the following diagram is commutative: | |
\[ | |
\xymatrix{ | |
\mathrm{Funct}(\Specm A,k) \ar[r]^{-\circ\Phi} & \mathrm{Funct}(\Specm B,k) \\ | |
A \ar[u] \ar[r]_{\phi} & B \ar[u]\\ | |
} | |
\] | |
where $\phi:A\rightarrow B$ is the map that gives rise to $\Phi$. | |
[$\Leftarrow$] Suppose that for all algebraic functions $f\in \mathrm{Funct}(\Specm A,k)$, the pull-back $f\circ\Phi$ is algebraic. Then we have an induced map, obtained by chasing the diagram counter-clockwise: | |
$$ | |
\xymatrix{ | |
\mathrm{Funct}(\Specm A,k) \ar[r]^{-\circ\Phi} & \mathrm{Funct}(\Specm B,k) \\ | |
A \ar[u] \ar@{-->}[r]_{\phi} & B \ar[u]\\ | |
} | |
$$ | |
From $\phi$, we can construct the map $\Phi':\Specm B \rightarrow \Specm A$ given by $\Phi'(\mathfrak{m})=\phi^{-1}(\mathfrak{m})$. I claim that $\Phi=\Phi'$. If not, then for some $\mathfrak{m}\in \Specm B$ we have $\Phi(\mathfrak{m})\neq \Phi'(\mathfrak{m})$. By definition, for all algebraic functions $f\in \mathrm{Funct}(\Specm A,k)$, $f\circ\Phi=f\circ\Phi'$ so to arrive at a contradiction we show the following lemma:\\ | |
Given any two distinct points in $\Specm A=V(I)\subset k^n$, there exists some | |
algebraic $f$ that separates them. This is trivial when we realize that any | |
polynomial function is algebraic, and such polynomials separate points. | |
\end{proof} | |
\section{Serre's criterion and its variants} | |
We are going to now prove a useful criterion for a noetherian ring to be a | |
product of | |
normal domains, due to Serre: it states that a (noetherian) ring is normal if | |
and only if most of the localizations at prime ideals are discrete valuation | |
rings (this corresponds to the ring being \emph{regular} in codimension one, | |
though we have not defined regularity yet) and a more technical condition that | |
we will later interpret in terms of \emph{depth.} One advantage of this | |
criterion is that it does \emph{not} require the ring to be a product of | |
domains a priori. | |
\subsection{Reducedness} | |
There is a ``baby'' version of Serre's criterion for testing whether a ring is | |
reduced, which we star with. | |
Recall: | |
\begin{definition} | |
A ring $R$ is \textbf{reduced} if it has no nonzero nilpotents. | |
\end{definition} | |
\begin{proposition} \label{reducedcrit} | |
If $R$ is noetherian, then $R$ is reduced if and only if it satisfies the | |
following conditions: | |
\begin{enumerate} | |
\item Every associated prime of $R$ is minimal (no embedded primes). | |
\item If $\mathfrak{p}$ is minimal, then $R_{\mathfrak{p}}$ is a field. | |
\end{enumerate} | |
\end{proposition} | |
\begin{proof} | |
First, assume $R$ reduced. What can we say? Say $\mathfrak{p}$ is a minimal | |
prime; then $R_{\mathfrak{p}}$ has precisely one prime ideal (namely, | |
$\mathfrak{m}=\mathfrak{p}R_{\mathfrak{p}}$). It is in fact a local artinian | |
ring, though we | |
don't need that fact. The radical of $R_{\mathfrak{p}}$ is just $\mathfrak{m}$. | |
But $R$ was reduced, so $R_{\mathfrak{p}}$ was reduced; it's an easy argument | |
that localization preserves reducedness. So $\mathfrak{m}=0$. The fact that 0 | |
is a maximal ideal in $R_{\mathfrak{p}}$ says that it is a field. | |
On the other hand, we still have to do part 1. $R$ is reduced, so $\rad(R) = | |
\bigcap_{\mathfrak{p} \in \spec R} \mathfrak{p} = 0$. In particular, | |
\[ \bigcap_{\mathfrak{p} \ \mathrm{minimal}}\mathfrak{p} = 0. \] | |
The map | |
\[ R \to \prod_{\mathfrak{p} \ \mathrm{minimal}}R/\mathfrak{p} \] | |
is injective. The associated primes of the product, however, are just the | |
minimal primes. So $\ass(R)$ can contain only minimal primes. | |
That's one direction of the proposition. Let us prove the converse now. Assume | |
$R$ satisfies the two conditions listed. In other words, $\ass(R)$ consists of | |
minimal primes, and each $R_{\mathfrak{p}}$ for $\mathfrak{p} \in \ass(R)$ is a | |
field. We would like to show that $R$ is reduced. | |
Primary decomposition tells us that there is an injection | |
\[ R \hookrightarrow \prod_{\mathfrak{p}_i \ \mathrm{minimal}} M_i, \quad M_i | |
\ \ \mathfrak{p}_i-\mathrm{primary}. \] | |
In this case, each $M_i$ is primary with respect to a minimal prime. We have a | |
map | |
\[ R \hookrightarrow \prod M_i \to \prod (M_i)_{\mathfrak{p}_i}, \] | |
which is injective, because when you localize a primary module at its | |
associated prime, you don't kill anything by definition of primariness. Since | |
we can draw a diagram | |
\[ | |
\xymatrix{ | |
R \ar[r] \ar[d] & \prod M_i \ar[d] \\ | |
\prod R_{\mathfrak{p}_i} \ar[r] & \prod (M_i)_{\mathfrak{p}_i} | |
} | |
\] | |
and the map $R \to \prod (M_i)_{\mathfrak{p}_i}$ is injective, the downward | |
arrow on the right injective. Thus $R$ can be embedded in | |
a product of the fields $\prod R_{\mathfrak{p}_i}$, so is reduced. | |
\end{proof} | |
This proof actually shows: | |
\begin{proposition}[Scholism] A noetherian ring $R$ is reduced iff it injects | |
into a product of fields. We can take the fields to be the localizations at the | |
minimal primes. | |
\end{proposition} | |
\begin{example} | |
Let $R = k[X]$ be the coordinate ring of a variety $X$ in | |
$\mathbb{C}^n$. Assume $X$ is | |
reduced. Then $\mathrm{MaxSpec} R$ is a union of irreducible components | |
$X_i$, which | |
are the closures of the minimal primes of $R$. The fields you get by localizing | |
at minimal primes depend only on the irreducible components, and in fact are | |
the rings of meromorphic functions on $X_i$. | |
Indeed, we have a map | |
\[ k[X] \to \prod k[X_i] \to \prod k(X_i). \] | |
If we don't assume that $R$ is radical, this is \textbf{not} true. | |
\end{example} | |
There is a stronger condition than being reduced we could impose. We could say: | |
\begin{proposition} | |
If $R$ is a noetherian ring, then $R$ is a domain iff | |
\begin{enumerate} | |
\item $R$ is reduced. | |
\item $R$ has a unique minimal prime. | |
\end{enumerate} | |
\end{proposition} | |
\begin{proof} | |
One direction is obvious. A domain is reduced and $(0)$ is the minimal prime. | |
The other direction is proved as follows. Assume 1 and 2. Let $\mathfrak{p}$ be | |
the unique minimal prime of $R$. Then $\rad (R) = 0 = \mathfrak{p}$ as every | |
prime ideal contains $\mathfrak{p}$. As $(0)$ is a prime ideal, $R$ is | |
a domain. | |
\end{proof} | |
We close by making some remarks about this embedding of $R$ into a product of | |
fields. | |
\begin{definition} | |
Let $R$ be any ring, not necessarily a domain. Let $K(R)$ be the localized ring | |
$S^{-1}R$ where $S$ is the multiplicatively closed set of nonzerodivisors in | |
$R$. $K(R)$ is called the \textbf{total ring of fractions} of $R$. | |
When $R$ is a field, this is the quotient field. | |
\end{definition} | |
First, to get a feeling for this, we show: | |
\begin{proposition} Let $R$ be noetherian. The set of nonzerodivisors $S$ | |
can be described by | |
$S = R- \bigcup_{\mathfrak{p} \in \ass(R)} \mathfrak{p}$. | |
\end{proposition} | |
\begin{proof} | |
If $x \in\mathfrak{p} \in \ass(R)$, then $x$ must kill something in $R$ as it | |
is in an associated prime. So $x$ is a zerodivisor. | |
Conversely, suppose $x$ is a zerodivisor, say $xy = 0$ for some $y \in R - | |
\left\{0\right\}$. In | |
particular, $x \in \ann(y)$. We have an injection $R/\ann(y) \hookrightarrow R$ | |
sending 1 to $y$. But $R/\ann(y)$ is nonzero, so it has an associated prime | |
$\mathfrak{p}$ of $R/\ann(y)$, which contains $\ann(y)$ and thus $x$. But | |
$\ass(R/\ann(y)) \subset \ass(R)$. | |
So $x$ is contained in a prime in $\ass(R)$. | |
\end{proof} | |
Assume now that $R$ is reduced. Then $K(R) = S^{-1}R$ where $S$ is the | |
complement of the union of the minimal primes. | |
At least, we can claim: | |
\begin{proposition} Let $R$ be reduced and noetherian. Then | |
$K(R) = \prod_{\mathfrak{p}_i \ \mathrm{minimal}} R_{\mathfrak{p}_i}$. | |
\end{proposition} | |
So $K(R)$ is the product of fields into which $R$ embeds. | |
We now continue the discussion begun last time. Let $R$ be noetherian and $M$ a | |
finitely generated $R$-module. We would like to understand very rough features | |
of $M$. | |
We can embed $M$ into a larger $R$-module. | |
Here are two possible approaches. | |
\begin{enumerate} | |
\item $S^{-1}M$, where $S$ is a large multiplicatively closed subset of $M$. | |
Let us take $S $ to be the set of all $a \in R$ such that $M | |
\stackrel{a}{\to}M$ is injective, i.e. $a$ is not a zerodivisor on $M$. Then | |
the map | |
\[ M \to S^{-1}M \] | |
is an injection. Note that $S$ is the complement of the union of $\ass(R)$. | |
\item Another approach would be to use a \emph{primary decomposition} | |
\[ M \hookrightarrow \prod M_i, \] | |
where each $M_i$ is $\mathfrak{p}_i$-primary for some prime $\mathfrak{p}_i$ | |
(and these primes range over $\ass(M)$). In this case, it is clear that | |
anything not in each $\mathfrak{p}_i$ acts injectively. So we can draw a | |
commutative diagram | |
\[ | |
\xymatrix{ | |
M \ar[d] \ar[r] & \prod M_i \ar[d] \\ | |
\prod M_{\mathfrak{p}_i} \ar[r] & \prod (M_i)_{\mathfrak{p}_i} | |
}. | |
\] | |
The map going right and down is injective. | |
It follows that $M$ injects into the product of its localizations at associated | |
primes. | |
\end{enumerate} | |
The claim is that these constructions agree if $M$ has no embedded primes. | |
I.e., if there are no nontrivial containments among the associated primes of | |
$M$, then $S^{-1}M$ (for $S = R - \bigcup_{\mathfrak{p} \in \ass(M)} | |
\mathfrak{p}$) | |
is just $\prod M_{\mathfrak{p}}$. | |
To see this, note that any element of $S$ must act invertibly on $\prod | |
M_{\mathfrak{p}}$. We thus see that there is always a map | |
\[ S^{-1}M \to \prod_{\mathfrak{p} \in \ass(M)} M_{\mathfrak{p}} . \] | |
\begin{proposition} | |
This is an isomorphism if $M$ has no embedded primes. | |
\end{proposition} | |
\begin{proof} | |
Let us go through a series of reductions. Let $I = \ann(M) = \left\{a: aM | |
= 0\right\}$. Without loss of generality, we can replace $R $ by $R/I$. This plays nice with the | |
associated primes. | |
The assumption is now that $\ass(M)$ consists of the minimal | |
primes of $R$. | |
Without loss of generality, we can next replace $R$ by $S^{-1}R$ and $M$ by | |
$S^{-1}M$, because that doesn't affect the conclusion; localization plays nice | |
with associated primes. | |
Now, however, $R$ is artinian: i.e., all primes of $R$ are minimal (or | |
maximal). Why is this? | |
Let $R$ be \emph{any} noetherian ring and $S = R - \bigcup_{\mathfrak{p} \ | |
\mathrm{minimal}} \mathfrak{p}$. Then I claim that $S^{-1}R$ is artinian. We'll | |
prove this in a moment. | |
So $R$ is artinian, hence a product $\prod R_i$ where each $R_i$ is local | |
artinian. Without loss of generality, we can replace $R$ by $R_i$ by taking | |
products. The condition we are trying to prove is now that | |
\[ S^{-1}M \to M_{\mathfrak{m}} \] | |
for $\mathfrak{m} \subset R$ the maximal ideal. But $S$ is the complement of | |
the union of the minimal primes, so it is $R - \mathfrak{m}$ as $R$ has one | |
minimal (and maximal) ideal. This is obviously an isomorphism: indeed, both | |
are $M$. | |
\end{proof} | |
\add{proof of artianness} | |
\begin{corollary} | |
Let $R$ be a noetherian ring with no embedded primes (i.e. $\ass(R)$ consists | |
of minimal primes). | |
Then $K(R) = \prod_{\mathfrak{p}_i \ \mathrm{minimal}} R_{\mathfrak{p_i}}$. | |
\end{corollary} | |
If $R$ is reduced, we get the statement made last time: there are no | |
embedded primes, and $K(R)$ is a product of | |
fields. | |
\subsection{The image of $M \to S^{-1}M$} | |
Let's ask now the following question. Let $R$ be a noetherian ring, $M$ | |
a finitely generated | |
$R$-module, and $S$ the set of nonzerodivisors on $M$, i.e. $R - | |
\bigcup_{\mathfrak{p} \in \ass(M)} \mathfrak{p}$. We have seen that there is an | |
imbedding | |
\[ \phi: M \hookrightarrow S^{-1}M. \] | |
What is the image? Given $x \in S^{-1}M$, when does it belong to the imbedding | |
above. | |
To answer such a question, it suffices to check locally. In particular: | |
\begin{proposition} | |
$x$ belongs to the image of $M $ in $S^{-1}M$ iff for every $\mathfrak{p} \in | |
\spec R$, the image of $x$ in $(S^{-1}M)_{\mathfrak{p}}$ lies inside | |
$M_{\mathfrak{p}}$. | |
\end{proposition} | |
This isn't all that interesting. However, it turns out that you can check this | |
at a smaller set of primes. | |
\begin{proposition} | |
In fact, it suffices to show that $x$ is in the image of $\phi_{\mathfrak{p}}$ | |
for every $\mathfrak{p} \in \ass(M/sM)$ where $s \in S$. | |
\end{proposition} | |
This is a little opaque; soon we'll see what it actually means. | |
The proof is very simple. | |
\begin{proof} | |
Remember that $ x \in S^{-1}M$. In particular, we can write $x = y/s$ where $y | |
\in M, s \in S$. What we'd like to prove that $x \in M$, or equivalently that | |
$y \in sM$.\footnote{In general, this would be equivalent to $ty \in tsM$ for | |
some $t \in S$; but $S$ consists of nonzerodivisors on $M$.} | |
In particular, we want to know that $y$ maps to zero in $M/sM$. If not, there | |
exists an associated prime $\mathfrak{p} \in \ass(M/sM)$ such that $y$ does not | |
get | |
killed in $(M/sM)_{\mathfrak{p}}$. | |
We have assumed, however, for every associated prime $\mathfrak{p}\in \ass(M)$, | |
$x \in ( S^{-1}M)_{\mathfrak{p}}$ lies in the image of $M_{\mathfrak{p}}$. This | |
states that the image of $y$ in this quotient $(M/sM)_{\mathfrak{p}}$ is zero, | |
or that $y$ is divisible by $s$ in this localization. | |
\end{proof} | |
The case we actually care about is the following: | |
Take $R$ as a noetherian domain and $M = R$. Then $S = R - \left\{0\right\}$ | |
and $S^{-1}M $ is just the fraction field $K(R)$. The goal is to describe $R$ | |
as a subset of $K(R)$. What we have proven is that $R$ is the intersection in | |
the fraction field | |
\[ \boxed{ R = \bigcap_{\mathfrak{p} \in \ass(R/s), s \in R - 0} | |
R_{\mathfrak{p}} . }\] | |
So to check that something belongs to $R$, we just have to check that in a | |
\emph{certain set of localizations}. | |
Let us state this as a result: | |
\begin{theorem} \label{notreallykrullthm} | |
If $R$ is a noetherian domain | |
\[ R = \bigcap_{\mathfrak{p} \in \ass(R/s), s \in R - 0} | |
R_{\mathfrak{p}} \] | |
\end{theorem} | |
\subsection{Serre's criterion} | |
We can now state a result. | |
\begin{theorem}[Serre] | |
\label{serrecrit1} | |
Let $R$ be a noetherian domain. Then $R $ is integrally | |
closed iff it satisfies | |
\begin{enumerate} | |
\item For any $\mathfrak{p} \subset R$ of height one, $R_{\mathfrak{p}}$ is a | |
DVR. | |
\item For any $s \neq 0$, $R/s$ has no embedded primes (i.e. all the | |
associated primes of $R/s$ are height one). | |
\end{enumerate} | |
\end{theorem} | |
Here is the non-preliminary version of the Krull theorem. | |
\begin{theorem}[Algebraic Hartogs] | |
Let $R$ be a noetherian integrally closed ring. Then | |
\[ R = \bigcap_{\mathfrak{p} \ \mathrm{height \ one}} R_{\mathfrak{p}}, \] | |
where each $R_{\mathfrak{p}}$ is a DVR. | |
\end{theorem} | |
\begin{proof} | |
Now evident from the earlier result \cref{notreallykrullthm} and Serre's criterion. | |
\end{proof} | |
Earlier in the class, we proved that a domain was integrally closed if and only | |
if it could be described as an intersection of valuation rings. We have now | |
shown that when $R$ is noetherian, we can take \emph{discrete} valuation rings. | |
\begin{remark} | |
In algebraic geometry, say $R = \mathbb{C}[x_1, \dots, x_n]/I$. Its maximal | |
spectrum is a subset of $\mathbb{C}^n$. If $I$ is prime, and $R$ a domain, | |
this variety is | |
irreducible. We are trying to describe $R$ inside its field of fractions. | |
The field of fractions are like the ``meromorphic functions''; $R$ is like the | |
holomorphic functions. Geometrically, this states to check that a meromorphic | |
function is holomorphic, you can just check this by computing the ``poleness'' | |
along each codimension one subvariety. If the function doesn't blow up on each | |
of the codimension one subvarieties, and $R$ is normal, then you can extend it | |
globally. | |
This is an algebraic version of Hartog's theorem: this states that a | |
holomorphic function on $\mathbb{C}^2 - (0,0)$ extends over the origin, because | |
this has codimension $>1$. | |
All the obstructions of extending a function to all of $\spec R$ are in | |
codimension one. | |
\end{remark} | |
Now, we prove Serre's criterion. | |
\begin{proof} | |
Let us first prove that $R$ is integrally closed if 1 and 2 occur. We know that | |
\[ R = \bigcap_{\mathfrak{p} \in \ass(R/x), x \neq 0} R_{\mathfrak{p}} ; \] | |
by condition 1, each such $\mathfrak{p}$ is of height one, and | |
$R_{\mathfrak{p}}$ is a DVR. So $R$ is the intersection of DVRs and thus | |
integrally closed. | |
The hard part is going in the other direction. Assume $R$ is integrally closed. | |
We want to prove the two conditions. In $R$, consider the following conditions | |
on a prime ideal $\mathfrak{p}$: | |
\begin{enumerate} | |
\item $\mathfrak{p}$ is an associated prime of $R/x$ for some $x \neq 0$. | |
\item $\mathfrak{p} $ is height one. | |
\item $\mathfrak{p}_{\mathfrak{p}}$ is principal in $R_{\mathfrak{p}}$. | |
\end{enumerate} | |
First, 3 implies 2 implies 1. 3 implies that $\mathfrak{p}$ contains an element | |
$x$ which | |
generates $\mathfrak{p}$ after localizing. | |
It follows that there can be no prime between $(x)$ and $\mathfrak{p}$ because | |
that would be preserved under localization. Similarly, 2 implies 1 is easy. If | |
$\mathfrak{p}$ is minimal over $(x)$, then $\mathfrak{p} \in \ass R/(x)$ since | |
the minimal primes in the support are always associated. | |
We are trying to prove the inverse implications. In that case, the claims | |
of the theorem will be proved. We have to show that 1 implies 3. | |
This is an argument we really saw last time, but let's see it again. Say | |
$\mathfrak{p} \in \ass(R/x)$. We can replace $R$ by $R_{\mathfrak{p}}$ so that | |
we can assume that $\mathfrak{p}$ is maximal. We want to show that | |
$\mathfrak{p}$ is generated by one | |
element. | |
What does the condition $\mathfrak{p} \in \ass(R/x)$ buy us? It tells us that | |
there is $\overline{y} \in R/x$ such that $\ann(\overline{y}) = \mathfrak{p}$. | |
In particular, there is $y \in R$ such that $\mathfrak{p}y \subset (x)$ and $y | |
\notin (x)$. | |
We have the element $y/x \in K(R)$ which sends $\mathfrak{p}$ into $R$. That | |
is, | |
\[ (y/x) \mathfrak{p} \subset R. \] | |
There are two cases to consider, as in last time: | |
\begin{enumerate} | |
\item $(y/x) \mathfrak{p} = R$. Then $\mathfrak{p} = R (x/y)$ so $\mathfrak{p} | |
$ is principal. | |
\item $(y/x) \mathfrak{p} \neq R$. In particular, $(y/x)\mathfrak{p} \subset | |
\mathfrak{p}$. Then since $\mathfrak{p}$ is finitely generated, we find that | |
$y/x $ is | |
integral over $R$, hence in $R$. This is a contradiction as $y \notin (x)$. | |
\end{enumerate} | |
Only the first case is now possible. So $\mathfrak{p}$ is in fact principal. | |
\end{proof} | |