s21-algebra-1/inputs/projective_spaces.tex
2023-07-22 21:52:48 +02:00

677 lines
25 KiB
TeX

Let $\mathfrak{l}$ be any field.
\begin{definition}
For a $\mathfrak{l}$-vector space $V$, let $\mathbb{P}(V)$ be
the set of
one-dimensional subspaces of $V$.
Let $\mathbb{P}^n(\mathfrak{l}) \coloneqq
\mathbb{P}(\mathfrak{l}^{n+1})$, the
\vocab[Projective space]{$n$-dimensional projective space over $\mathfrak{l}$}.
If $\mathfrak{l}$ is kept fixed, we will often write
$\mathbb{P}^n$ for
$\mathbb{P}^n(\mathfrak{l})$.
When dealing with $\mathbb{P}^n$, the usual convention is to use
$0$ as the
index of the first coordinate.
We denote the one-dimensional subspace generated by $(x_0,\ldots,x_n) \in
\mathfrak{k}^{n+1} \setminus \{0\}$ by
$[x_0,\ldots,x_n] \in \mathbb{P}^n$.
If $x = [x_0,\ldots,x_n] \in \mathbb{P}^n$, the
$(x_{i})_{i=0}^n$ are
called
\vocab{homogeneous coordinates} of $x$.
At least one of the $x_{i}$ must be $\neq 0$.
\end{definition}
\begin{remark}
There are points $[1,0],
[0,1] \in \mathbb{P}^1$ but there
is no point $[0,0]
\in \mathbb{P}^1$.
\end{remark}
\begin{definition}[Infinite hyperplane]
For $0 \le i \le n$ let $U_i \subseteq
\mathbb{P}^n$ denote the set of $[x_0,\ldots,x_{n}]$ with
$x_{i}\neq 0$.
This is a correct definition since two different sets $[x_0,\ldots,x_{n}]$ and
$[\xi_0,\ldots,\xi_n]$ of homogeneous coordinates for the
same point $x \in
\mathbb{P}^n$ differ by scaling with a $\lambda \in
\mathfrak{l}^{\times}$,
$x_i = \lambda \xi_i$.
Since not all $x_i$ may be $0$, $\mathbb{P}^n =
\bigcup_{i=0}^n U_i$.
We identify $\mathbb{A}^n =
\mathbb{A}^n(\mathfrak{l}) = \mathfrak{l}^n$
with
$U_0$ by identifying $(x_1,\ldots,x_n) \in \mathbb{A}^n$ with
$[1,x_1,\ldots,x_n] \in \mathbb{P}^n$.
Then $\mathbb{P}^1 = \mathbb{A}^1 \cup \{\infty\} $
where $\infty=[0,1]$.
More generally, when $n > 0$ $\mathbb{P}^n \setminus
\mathbb{A}^n$ can be
identified with $\mathbb{P}^{n-1}$ identifying
$[0,x_1,\ldots,x_n] \in
\mathbb{P}^n \setminus \mathbb{A}^n$ with
$[x_1,\ldots,x_n] \in
\mathbb{P}^{n-1}$.
Thus $\mathbb{P}^n$ is $\mathbb{A}^n \cong
\mathfrak{l}^n$ with a copy of
$\mathbb{P}^{n-1}$ added as an \vocab{infinite hyperplane} .
\end{definition}
\subsubsection{Graded rings and homogeneous ideals}
\begin{notation}
Let $\mathbb{I} = \N$ or $\mathbb{I} = \Z$.
\end{notation}
\begin{definition}
By an \vocab[Graded ring]{$\mathbb{I}$-graded ring} $A_\bullet$ we understand a
ring $A$ with a collection $(A_d)_{d \in \mathbb{I}}$ of
subgroups of the
additive group $(A, +)$ such that $A_a \cdot A_b \subseteq
A_{a + b}$ for $a,b
\in \mathbb{I}$ and such that $A = \bigoplus_{d \in \mathbb{I}} A_d$ in
the
sense that every $r \in A$ has a unique decomposition $r =
\sum_{d \in
\mathbb{I}} r_d$ with $r_d \in A_d$ and but finitely many $r_d
\neq 0$.
We call the $r_d$ the \vocab{homogeneous components} of $r$.
An ideal $I \subseteq A$ is called \vocab{homogeneous} if $r \in I
\implies
\forall d \in \mathbb{I} ~ r_d \in I_d$ where $I_d \coloneqq I
\cap A_d$.
By a \vocab{graded ring} we understand an $\N$-graded ring.
Tin this case, $A_{+} \coloneqq \bigoplus_{d=1}^{\infty}
A_d = \{r \in A | r_0
= 0\} $ is called the \vocab{augmentation ideal} of $A$.
\end{definition}
\begin{remark}[Decomposition of $1$]
If $1 = \sum_{d \in \mathbb{I}} \varepsilon_d$ is the
decomposition into homogeneous components, then $\varepsilon_a = 1 \cdot
\varepsilon_a = \sum_{b \in \mathbb{I}} \varepsilon_a\varepsilon_b$ with
$\varepsilon_a\varepsilon_b \in A_{a+b}$.
By the uniqueness of the decomposition into homogeneous components,
$\varepsilon_a \varepsilon_0 = \varepsilon_a$ and $b \neq 0 \implies
\varepsilon_a \varepsilon_b = 0$.
Applying the last equation with $a = 0$ gives $b\neq 0 \implies \varepsilon_b =
\varepsilon_0 \varepsilon _b = 0$.
Thus $1 = \varepsilon_0 \in A_0$.
\end{remark}
\begin{remark}
The augmentation ideal of a graded ring is a homogeneous ideal.
\end{remark}
% Graded rings and homogeneous ideals (2)
\begin{proposition}
\footnote{This holds for both $\Z$-graded and $\N$-graded rings.}
\begin{itemize}
\item
A principal ideal generated by a homogeneous element is homogeneous.
\item
The operations $\sum, \bigcap, \sqrt{}$ preserve homogeneity.
\item
An ideal is homogeneous iff it can be generated by a family of homogeneous
elements.
\end{itemize}
\end{proposition}
\begin{proof}
Most assertions are trivial.
We only show that $J$ homogeneous $\implies \sqrt{J} $
homogeneous.
Let $A$ be $\mathbb{I}$-graded, $f \in \sqrt{J} $ and
$f = \sum_{d \in
\mathbb{I}} f_d$ the decomposition.
To show that all $f_d \in \sqrt{J} $, we use induction on $N_f
\coloneqq \# \{d
\in \mathbb{I} | f_d \neq 0\}$.
$N_f = 0$ is trivial.
Suppose $N_f > 0$ and $e \in \mathbb{I}$ is maximal with $f_e \neq
0$.
For $l \in \N$, the $le$-th homogeneous component of $f^l$ is $f_e^l$.
Choosing $l$ large enough such that $f^l \in J$ and using the homogeneity of
$J$, we find $f_e \in \sqrt{J}$.
As $\sqrt{J} $ is an ideal, $\tilde f \coloneqq f - f_e \in
\sqrt{J} $.
As $N_{\tilde f} = N_f -1$, the induction assumption may be
applied to $\tilde
f$ and shows $f_d \in \sqrt{J} $ for $d \neq e$.
\end{proof}
\begin{fact}
A homogeneous ideal is finitely generated iff it can be generated by finitely
many of its homogeneous elements.
In particular, this is always the case when $A$ is a Noetherian ring.
\end{fact}
\subsubsection{The Zariski topology on $\mathbb{P}^n$}
\begin{notation}
Recall that for $\alpha \in \N^{n+1}$ $|\alpha| =
\sum_{i=0}^{n} \alpha_i$ and
$x^\alpha = x_0^{\alpha_0} \cdot \ldots \cdot x_n^{\alpha_n}$.
\end{notation}
\begin{definition}[Homogeneous polynomials]
Let $R$ be any ring and $f = \sum_{\alpha \in \N^{n+1}}
f_\alpha X^{\alpha}\in R[X_0,\ldots,X_n]$.
We say that $f$ is \vocab{homogeneous of degree $d$} if $|\alpha| \neq d
\implies f_\alpha = 0$ .
We denote the subset of homogeneous polynomials of degree $d$ by
$R[X_0,\ldots,X_n]_d \subseteq R[X_0,\ldots,X_n]$.
\end{definition}
\begin{remark}
This definition gives $R$ the structure of a graded ring.
\end{remark}
\begin{definition}[Zariski topology on $\mathbb{P}^n(\mathfrak{k})$]
\label{ztoppn}
Let $A = \mathfrak{k}[X_0,\ldots,X_n]$.
\footnote{As always, $\mathfrak{k}$ is algebraically closed}
For $f \in A_d = \mathfrak{k}[X_0,\ldots,X_n]_d$, the validity of the equation
$f(x_0,\ldots,x_{n}) = 0$ does not depend on the choice of homogeneous
coordinates, as
\[
f(\lambda x_0,\ldots, \lambda x_n) 0 \lambda^d
f(x_0,\ldots,x_n)
\]
Let $\Vp(f) \coloneqq \{x \in \mathbb{P}^n | f(x)
= 0\}$.
We call a subset $X \subseteq \mathbb{P}^n$ Zariski-closed if it
can be
represented as
\[
X = \bigcap_{i=1}^k \Vp(f_i)
\]
where the $f_i \in A_{d_i}$
are homogeneous polynomials.
\end{definition}
\pagebreak
\begin{fact}
If $X = \bigcap_{i = 1}^k \Vp(f_i) \subseteq
\mathbb{P}^n$ is closed, then $Y
= X \cap \mathbb{A}^n$ can be identified with the closed subset
\[
\{(x_1,\ldots,x_n) \in \mathfrak{k}^n | f_i(1,x_1,\ldots,x_n) = 0, 1
\le i \le
k\} \subseteq \mathfrak{k}^n
\]
Conversely, if $Y \subseteq \mathfrak{k}^n$ is
closed it has the form
\[
\{(x_1,\ldots,x_n) \in \mathfrak{k}^n |
g_i(x_1,\ldots,x_n) = 0, 1 \le i \le k\}
\]
and can thus be identified with $X
\cap \mathbb{A}^n$ where $X \coloneqq \bigcap_{i=1}^k
\Vp(f_i)$ is given by
\[
f_i(X_0,\ldots,X_n) \coloneqq X_0^{d_i} g_i(X_1 / X_0,\ldots, X_n / X_0), d_i
\ge \deg(g_i)
\]
Thus, the Zariski topology on $\mathfrak{k}^n$ can be
identified with the topology induced by the Zariski topology on
$\mathbb{A}^n =
U_0$, and the same holds for $U_i$ with $0 \le i \le n$.
In this sense, the Zariski topology on $\mathbb{P}^n$ can be
thought of as
gluing the Zariski topologies on the $U_i \cong \mathfrak{k}^n$.
\end{fact}
% The Zariski topology on P^n (2)
\begin{definition}
Let $I \subseteq A = \mathfrak{k}[X_0,\ldots,X_n]$ be a homogeneous ideal.
Let $\Vp(I) \coloneqq
\{[x_0,\ldots,_n] \in \mathbb{P}^n |
\forall f \in I ~
f(x_0,\ldots,x_n) = 0\}$ As $I$ is homogeneous, it is sufficient to impose this
condition for the homogeneous elements $f \in I$.
Because $A$ is Noetherian, $I$ can finitely generated by homogeneous elements
$(f_i)_{i=1}^k$ and
$\Vp(I)=\bigcap_{i=1}^k \Vp(f_i)$ as
in
\ref{ztoppn}.
Conversely, if the homogeneous $f_i$ are given, then $I = \langle
f_1,\ldots,f_k \rangle_A$ is homogeneous.
\end{definition}
\begin{remark}
Note that $V(A) = V(A_+) = \emptyset$.
\end{remark}
\begin{fact}
For homogeneous ideals in $A$ and $m \in \N$, we have:
\begin{itemize}
\item
$\Vp(\sum_{\lambda \in \Lambda} I_\lambda) = \bigcap_{\lambda \in \Lambda}
\Vp(I_\lambda)$
\item
$\Vp(\bigcap_{k=1}^m I_k) = \Vp(\prod_{k=1}^{m} I_k) =
\bigcup_{k=1}^m \Vp(I_k)$
\item
$\Vp(\sqrt{I}) = \Vp(I)$
\end{itemize}
\end{fact}
\begin{fact}
If $X = \bigcup_{\lambda \in \Lambda} U_\lambda$ is an
open covering of a topological space then $X$ is Noetherian iff there is a
finite subcovering and all $U_\lambda$ are Noetherian.
\end{fact}
\begin{proof}
By definition, a topological space is Noetherian $\iff$ all open subsets are
quasi-compact.
\end{proof}
\begin{corollary}
The Zariski topology on $\mathbb{P}^n$ is indeed a topology.
The induced topology on the open set $\mathbb{A}^n =
\mathbb{P}^n \setminus
\Vp(X_0) \cong \mathfrak{k}^n$ is the Zariski
topology on $\mathfrak{k}^n$.
The same holds for all $U_i = \mathbb{P}^n \setminus
\Vp(X_i) \cong
\mathfrak{k}^n$.
Moreover, the topological space $\mathbb{P}^n$ is Noetherian.
\end{corollary}
\subsection{Noetherianness of graded rings}
\begin{proposition}
For a graded ring $R_{\bullet}$, the following conditions
are equivalent:
\begin{enumerate}[A]
\item
$R$ is Noetherian.
\item
Every homogeneous ideal of $R_{\bullet}$ is finitely
generated.
\item
Every chain $I_0\subseteq I_1 \subseteq \ldots$ of homogeneous ideals
terminates.
\item
Every set $\mathfrak{M} \neq \emptyset$ of homogeneous ideals has a
$\subseteq$-maximal element.
\item
$R_0$ is Noetherian and the ideal $R_+$ is finitely generated.
\item
$R_0$ is Noetherian and $R / R_0$ is of finite type.
\end{enumerate}
\end{proposition}
\begin{proof}
\noindent\textbf{A $\implies$ B,C,D} trivial.
\noindent\textbf{B $\iff$ C $\iff$ D} similar to the proof about Noetherianness.
\noindent\textbf{B $\land$ C $\implies $E}
B implies that $R_+$ is finitely generated.
Since $I \oplus R_+$ is homogeneous for any homogeneous ideal $I \subseteq
R_0$, C implies the Noetherianness of $R_0$.
\noindent\textbf{E $\implies$ F}
Let $R_+$ be generated by $f_i \in R_{d_i}, d_i > 0$ as
an ideal.
\begin{claim}
The $R_0$-subalgebra $\tilde R$ of $R$ generated by the $f_i$ equals $R$.
\end{claim}
\begin{subproof}
It is sufficient to show that every homogeneous $f \in R_d$ belongs to $\tilde
R$.
We use induction on $d$.
The case of $d = 0$ is trivial.
Let $d > 0$ and $R_e \subseteq \tilde R$ for all $e < d$.
as $f \in R_+$, $f = \sum_{i=1}^{k} g_if_i$.
Let $f_a = \sum_{i=1}^{k} g_{i, a-d_i} f_i$, where
$g_i = \sum_{b=0}^{\infty}
g_{i,b}$ is the decomposition into homogeneous
components.
Then $f = \sum_{a=0}^{\infty} f_a$ is the decomposition of $f$ into
homogeneous
components, hence $a \neq d \implies f_a = 0 $.
Thus we may assume $g_i \in R_{d-d_i}$.
As $d_i > 0$, the induction assumption may now be applied to $g_i$, hence $g_i
\in \tilde R$, hence $f \in \tilde R$.
\end{subproof}
\noindent\textbf{F $\implies$ A}
Hilbert's Basissatz (
\ref{basissatz})
\end{proof}
\subsection{The projective form of the Nullstellensatz and the closed subsets
of $\mathbb{P}^n$} Let $A = \mathfrak{k}[X_0,\ldots,X_n]$.
% Lecture 12
\begin{proposition}[Projective form of the Nullstellensatz]
\label{hnsp}
If $I \subseteq A$ is a homogeneous ideal and $f \in A_d$ with $d>0$, then
$\Vp(I) \subseteq \Vp(f) \iff f \in
\sqrt{I}$.
\end{proposition}
\begin{proof}
$\impliedby$ is clear.
Let $\Vp(I) \subseteq \Vp(f)$.
If $x = (x_0,\ldots,x_n) \in \Va(I)$, then either $x = 0$ in
which case $f(x) =
0$ since $d > 0$ or the point $[x_0,\ldots,x_n] \in
\mathbb{P}^n$ is
well-defined and belongs to $\Vp(I) \subseteq
\Vp(f)$, hence $f(x) = 0$.
Thus $\Va(I) \subseteq \Va(f)$ and $f \in
\sqrt{I}$ be the Nullstellensatz
(
\ref{hns3}).
\end{proof}
\begin{definition}
\footnote{This definition is not too important, the characterization in the following remark suffices.}.
For a graded ring $R_\bullet$, let $\Proj(R_\bullet)$ be the set of
$\fp \in
\Spec R$ such that $\fp$ is a homogeneous ideal and $\fp \not\supseteq R_+$.
\end{definition}
\begin{remark}
\label{proja}
As the elements of $A_0 \setminus \{0\}$ are units in $A$ it follows that for
every homogeneous ideal $I$ we have $I \subseteq A_+$ or $I = A$.
In particular, $\Proj(A_\bullet) = \{\fp \in \Spec A \setminus A_+ |
\fp
\text{ is homogeneous}\} $.
\end{remark}
\begin{proposition}
\label{bijproj}
There is a bijection
\begin{align*}
f: \{I \subseteq A_+ | I \text{ homogeneous
ideal}, I = \sqrt{I}\} & \longrightarrow \{X \subseteq \mathbb{P}^n | X \text{
closed}\} \\ I & \longmapsto \Vp(I)\\ \langle \{f \in A_d | d > 0, X
\subseteq \Vp(f)\} \rangle & \longmapsfrom X
\end{align*}
Under this bijection,
the irreducible subsets correspond to the elements of
$\Proj(A_\bullet)$.
\end{proposition}
\begin{proof}
From the projective form of the Nullstellensatz it follows that $f$ is
injective and that $f^{-1}(\Vp\left( I \right))
= \sqrt{I} = I$.
If $X \subseteq \mathbb{P}^n$ is closed, then $X =
\Vp(J)$ for some homogeneous
ideal $J \subseteq A$.
Without loss of generality loss of generality $J = \sqrt{J}$.
If $J \not\subseteq A_+$, then $J = A$ (
\ref{proja}), hence $X =
\Vp(J) =
\emptyset = \Vp(A_+)$.
Thus we may assume $J \subseteq A_+$, and $f$ is surjective.
Suppose $\fp \in \Proj(A_\bullet)$.
Then $\fp \neq A_+$ hence $X = \Vp(\fp) \neq \emptyset$ by the
proven part of
the proposition.
Assume $X = X_1 \cup X_2$ is a decomposition into proper closed subsets, where
$X_k = \Vp(I_k)$ for some $I_k \subseteq A_+, I_k =
\sqrt{I_k}$.
Since $X_k$ is a proper subset of $X$, there is $f_k \in I_k \setminus \fp$.
We have $\Vp(f_1f_2) \supseteq \Vp(f_k) \supseteq
\Vp(I_k)$ hence $\Vp(f_1f_2)
\supseteq \Vp(I_1) \cup \Vp(I_2) = X =
\Vp(\fp)$ and it follows that $f_1f_2\in
\sqrt{\fp} = \fp \lightning$.
Assume $X = \Vp(\fp)$ is irreducible, where $\fp =
\sqrt{\fp} \in A_+$ is
homogeneous.
The $\fp \neq A_+$ as $X = \emptyset$ otherwise.
Assume that $f_1f_2 \in \fp$ but $f_i \not\in A_{d_i}
\setminus \fp$.
Then $X \not \subseteq \Vp(f_i)$ by the projective
Nullstellensatz when $d_i >
0$ and because $\Vp(1) = \emptyset$ when $d_i = 0$.
Thus $X = (X \cap \Vp\left( f_1 \right)) \cup (X \cap \Vp(f_2))$ is a
proper
decomposition $\lightning$.
By lemma
\ref{homprime}, $\fp$ is a prime ideal.
\end{proof}
\begin{remark}
It is important that $I \subseteq A_{\color{red} +}$, since
$\Vp(A) = \Vp(A_+)
= \emptyset$ would be a counterexample.
\end{remark}
\begin{corollary}
$\mathbb{P}^n$ is irreducible.
\end{corollary}
\begin{proof}
Apply
\ref{bijproj} to $\{0\} \in \Proj(A_\bullet)$.
\end{proof}
\subsection{Some remarks on homogeneous prime ideals}
\begin{lemma}
\label{homprime}
Let $R_\bullet$ be an $\mathbb{I}$ graded ring
($\mathbb{I} = \N$ or
$\mathbb{I} = \Z$).
A homogeneous ideal $I \subseteq R$ is a prime ideal iff $1 \not\in I$ and for
homogeneous elements $f, g \in R , fg \in I \implies f \in I \lor g \in I$.
\end{lemma}
\begin{proof}
$\implies$ is trivial.
It suffices to show that for arbitrary $f,g \in R fg \in I \implies f \in I
\lor g \in I$.
Let $f = \sum_{d \in \mathbb{I}} f_d, g = \sum_{d \in \mathbb{I}} g_d $ be the
decompositions into homogeneous components.
If $f \not\in I$ and $g \not\in I$ there are $d,e \in I$ with $f_d \in I, g_e
\in I$, and they may assumed to be maximal with this property.
As $I$ is homogeneous and $fg \in I$, we have
$(fg)_{d+e} \in I$ but
\[
(fg)_{d+e} = f_dg_e + \sum_{\delta = 1}^{\infty} (f_{d + \delta} g_{e - \delta}
+ f_{d - \delta} g_{e + \delta})
\]
where $f_dg_e \not\in I$ by our assumption
on $I$ and all other summands on the right hand side are $\in I$ (as
$f_{d+
\delta} \in I$ and $g_{e + \delta} \in
I$ by the maximality of $d$ and $e$), a
contradiction.
\end{proof}
\begin{remark}
If $R_\bullet$ is $\N$-graded and $\fp \in \Spec R_0$, then $\fp \oplus R_+ =
\{r \in R | r_0 \in \fp\} $ is a homogeneous prime ideal of $R$.
\[
\{\fp \in \Spec R | \fp \text{ is a homogeneous ideal of }
R_\bullet\} = \Proj(R_\bullet) \sqcup \{\fp \oplus R_+ | \fp \in \Spec
R_0\}
\]
\end{remark}
\subsection{Dimension of $\mathbb{P}^n$}
\begin{proposition}
\begin{itemize}
\item
$\mathbb{P}^n$ is catenary.
\item
$\dim(\mathbb{P}^n) = n$.
Moreover, $\codim(\{x\} ,\mathbb{P}^n) = n$ for every $x \in
\mathbb{P}^n$.
\item
If $X \subseteq \mathbb{P}^n$ is irreducible and $x \in X$, then
$\codim(\{x\}, X) = \dim(X) = n -
\codim(X, \mathbb{P}^n)$.
\item
If $X \subseteq Y \subseteq \mathbb{P}^n$ are irreducible subsets,
then $\codim(X,Y) = \dim(Y) -
\dim(X)$.
\end{itemize}
\end{proposition}
\begin{proof}
Let $X \subseteq \mathbb{P}^n$ be irreducible.
If $x \in X$, there is an integer $0 \le i \le n$ and $X \in U_i =
\mathbb{P}^n \setminus \Vp(X_i)$.
Without loss of generality loss of generality $i = 0$.
Then $\codim(X, \mathbb{P}^n) = \codim(X \cap \mathbb{A}^n, \mathbb{A}^n)$ by
the locality of Krull codimension (
\ref{lockrullcodim}).
Applying this with $X = \{x\}$ and our results about the affine case gives the
second assertion.
If $Y$ and $Z$ are also irreducible with $X \subseteq Y \subseteq Z$, then
$\codim(X,Y) = \codim(X \cap \mathbb{A}^n, Y \cap \mathbb{A}^n)$, $\codim(X,Z)
= \codim(X \cap \mathbb{A}^n, Z \cap \mathbb{A}^n)$ and $\codim(Y,Z) =
\codim(Y
\cap \mathbb{A}^n, Z \cap \mathbb{A}^n)$.
Thus
\begin{align*}
\codim(X,Y) + \codim(Y,Z) & = \codim(X \cap \mathbb{A}^n, Y
\cap \mathbb{A}^n) + \codim(Y \cap \mathbb{A}^n, Z \cap \mathbb{A}^n) \\ & =
\codim(X \cap \mathbb{A}^n, Z \cap \mathbb{A}^n) \\ & = \codim(X, Z)
\end{align*}
because $\mathfrak{k}^n$ is catenary and the first point follows.
The remaining assertions can easily be derived from the first two.
\end{proof}
\subsection{The cone $C(X)$}
\begin{definition}
If $X \subseteq \mathbb{P}^n$ is closed, we define the
\vocab{affine cone over
$X$}
\[
C(X) = \{0\} \cup \{(x_0,\ldots,x_n) \in \mathfrak{k}^{n+1} \setminus
\{0\} | [x_0,\ldots,x_n] \in X\}
\]
If $X = \Vp(I)$ where $I \subseteq A_+ =
\mathfrak{k}[X_0,\ldots,X_n]_+$ is homogeneous, then $C(X) =
\Va(I)$.
\end{definition}
\begin{proposition}
\label{conedim}
\begin{itemize}
\item
$C(X)$ is irreducible iff $X$ is irreducible or $X = \emptyset$.
\item
If $X$ is irreducible, then
$\dim(C(X)) = \dim(X) + 1$ and
$\codim(C(X), \mathfrak{k}^{n+1}) = \codim(X, \mathbb{P}^n)$
\end{itemize}
\end{proposition}
\begin{proof}
The first assertion follows from
\ref{bijproj} and
\ref{bijiredprim}
(bijection of irreducible subsets and prime ideals in the projective
and affine case).
Let $d = \dim(X)$ and
\[
X_0 \subsetneq \ldots \subsetneq X_d = X \subsetneq
X_{d+1} \subsetneq \ldots \subsetneq X_n =
\mathbb{P}^n
\]
be a chain of irreducible subsets of $\mathbb{P}^n$.
Then
\[
\{0\} \subsetneq C(X_0) \subsetneq \ldots \subsetneq C(X_d) = C(X)
\subsetneq \ldots \subsetneq C(X_n) = \mathfrak{k}^{n+1}
\]
is a chain of irreducible subsets of $\mathfrak{k}^{n+1}$.
Hence $\dim(C(X)) \ge 1 + d$ and
$\codim(C(X), \mathfrak{k}^{n+1}) \ge n-d$.
Since
$\dim(C(X)) + \codim(C(X), \mathfrak{k}^{n+1}) = \dim(\mathfrak{k}^{n+1})
= n+1$,
the two inequalities must be equalities.
\end{proof}
\subsubsection{Application to hypersurfaces in $\mathbb{P}^n$}
\begin{definition}[Hypersurface]
Let $n > 0$.
By a \vocab{hypersurface} in $\mathbb{P}^n$ or
$\mathbb{A}^n$ we understand an irreducible closed subset
of codimension $1$.
\end{definition}
\begin{corollary}
If $P \in A_d$ is a prime element,
then $H = \Vp(P)$ is a hypersurface in $\mathbb{P}^n$
and every hypersurface $H$ in $\mathbb{P}^n$ can be obtained in this way.
\end{corollary}
\begin{proof}
If $H = \Vp(P)$ then $C(H) = \Va(P)$
is a hypersurface in $\mathfrak{k}^{n+1}$
by \ref{irredcodimone}.
By \ref{conedim}, $H$ is irreducible and of codimension $1$.
Conversely, let $H$ be a hypersurface in $\mathbb{P}^n$.
By \ref{conedim}, $C(H)$ is a hypersurface in
$\mathfrak{k}^{n+1}$,
hence $C(H) = \Vp(P)$ for some prime element $P \in A$
(again by \ref{irredcodimone}).
We have $H = \Vp(\fp)$ for some $\fp \in \Proj(A)$ and $C(H) = \Va(\fp)$.
By the bijection between closed subsets of $\mathfrak{k}^{n+1}$ and ideals
$I = \sqrt{I} \subseteq A$ (\ref{antimonbij}),
$\fp = P \cdot A$.
Let $P = \sum_{k=0}^{d}P_k$ with $P_d \neq 0$ be the decomposition
into homogeneous components.
If $P_e $ with $e < d$ was $\neq 0$,
it could not be a multiple of $P$ contradicting the homogeneity of
$\fp = P \cdot A$.
Thus, $P$ is homogeneous of degree $d$.
\end{proof}
\begin{definition}
A hypersurface $H \subseteq \mathbb{P}^n$ has \vocab{degree $d$}
if $H = \Vp(P)$,
where $P \in A_d$ is an irreducible polynomial.
\end{definition}
\subsubsection{Application to intersections in $\mathbb{P}^n$
and Bezout's theorem}
\begin{corollary}
Let $A \subseteq \mathbb{P}^n$ and $B \subseteq \mathbb{P}^n$ be
irreducible subsets of dimensions $a$ and $b$.
If $a+ b \ge n$,
then $A \cap B \neq \emptyset$
and every irreducible component of $A \cap B$
has dimension $\ge a + b - n$.
\end{corollary}
\begin{remark}
This shows that $\mathbb{P}^n$ indeed fulfilled the goal of allowing for
nicer results of algebraic geometry because ``solutions at infinity''
to systems of algebraic equations are present in $\mathbb{P}^n$
(see \ref{affineproblem}).
\end{remark}
\begin{proof}
The lower bound on the dimension of irreducible components of $A \cap B$ is
easily derived from the similar affine result
(corollary of the principal ideal theorem, \ref{codimintersection}).
From the definition of the affine cone it follows that
$C(A \cap B) = C(A) \cap C(B)$.
We have $\dim(C(A)) = a+1$ and $\dim(C(B)) = b + 1$ by
\ref{conedim}.
If $A \cap B = \emptyset$, then $C(A) \cap C(B) = \{0\}$ with $\{0\} $ as an
irreducible component, contradicting the lower bound $a + b + 1 - n > 0$ for
the dimension of irreducible components of $C(A) \cap C(B)$
(again \ref{codimintersection}).
\end{proof}
\begin{remark}[Bezout's theorem]
If $A \neq B$ are hypersurfaces of degree $a$ and $b$
in $\mathbb{P}^2$, then $A \cap B$ has $ab$ points counted by
(suitably defined) multiplicity.
\end{remark}
%TODO Proof of "Dimension of P^n"
% SLIDE APPLICATION TO HYPERSURFACES IN $\P^n$
%ERROR: C(H) = V_A(P)
%If n = 0, P = 0, V_P(P) = \emptyset is a problem!