Show that the natural map \( A_f \to \OO_{\spec A} (D(f)) \) is an isomorphism. (Possible hint: Exercise 3.5.E).
Proof:
Consider the multiplicative subset \( S = \{ g \in A \mid D(f) \subset D(g) \} \subset A\). By Exercise 3.5.E, \( D(f) \subset D(g) \Leftrightarrow f^n \in (g) \) for some \( n \geq 1 \Leftrightarrow g\) invertible in \(A_f\). By the last equivalent definition and the universal property of the localization, there is a natural map \( A_f \to S^{-1} A = \OO_{\spec A}(D(f)) \) which we will denote as \(\phi\). To see that this is an isomorphism, suppose \( \phi( \frac{g_1}{f^{r_1}} ) = \phi( \frac{g_2}{f^{r_2}} ) \) so that for each \( \pp \in D(f) \) ( i.e. \( f \notin \pp \) ) there is some \( h \notin \pp \) with \( h(g_1 f^{r_1} - g_2f^{r_2}) = 0 \). As this is true for each \( \pp \in D(f) \) so that \( V(h) \subset V(f) \), we have again by Exercise 3.5.E that \( h = f^k \) for some \( k \). Thus, the two are necessarily equal in \( A_f \). The map is clearly surjective since \( \{ 1, f, f^2, \dots \} \subset S \) so that \( S^{-1}A \subset A_f \).
$$\tag*{$\blacksquare$}$$
Exercise 4.1.B:
Make tiny changes to the above argument to show base identity for any distinguished open \(D(f)\). (Hint: judiciously replace \(A\) by \(A_f\) in the above argument.)
Proof:
Using quasicompactness from Exercise 3.5.C, there exists some \(n \) with
$$
D(f) = \bigcup_{i=1}^n D(h_i)
$$
By the previous exercise, we know that \( \OO_{\spec A}(D(f)) = A_f \) — thus, suppose \( \frac{s}{f^m} \in A_f \) is such that \( \res{D(f)}{D(h_i)} \frac{s}{f^m} = 0 \) — in other words, \( h_i^{k_i} \frac{s}{f^m} = 0 \) in \( A_{h_i} \). Now since \( \langle h_1, \dots, h_n \rangle = A_f \), there are some \( \frac{b_1}{f^{a_1}}, \dots, \frac{b_n}{f^{a_n}} \in A_f \) such that
$$
\sum_{i=1}^n \frac{b_i}{f^{a_i}} h_i = 1
$$
so that
$$
f^N s = \sum_{i=1}^n f^N \frac{b_i}{f^{a_i}} h_i s = = \sum_{i=1}^n b_i \cdot 0 = 0
$$
for large enough \(N\) — as \( f \) is invertible, this implies \( s = 0 \).
$$\tag*{$\blacksquare$}$$
Exercise 4.1.C:
Alter this argument appropriately to show base gluability for any distinguished open \(D(f)\).
Proof:
As was the case in the previous exercise, this simply requires copying the above argument and multiplying by a large enough power of \( f \).
$$\tag*{$\blacksquare$}$$
Exercise 4.1.D:
Suppose \(M\) is an \(A\)-module. Show that the following construction describes a sheaf \( \widetilde{M} \) on the distinguished base. Define \( \widetilde{M}(D(f)) \) to be the localization of \(M\) at the multiplicative set of all functions that do not vanish outside of \(V(f)\). Define restriction maps \(\res{D(f)}{D(g)}\) in the analogous way to \(OO_{\spec A}\). Show that this defines a sheaf on the distinguished base, and hence a sheaf on \(\spec A\). Then show that this is an \(\OO_{\spec A} \)-module.
Section 4.2: Visualizing Schemes II: Nilpotents
There are no exercises in this section, just commentary.
Section 4.3: The Definition of Schemes
Exercise 4.3.A:
Describe a bijection between the isomorphisms \(\spec A \to \spec A^{\prime}\) and the ring isomorphisms \(A^\prime \to A\). Hint: the hardest part is to show that if an isomorphism \(\pi: \spec A \to \spec A^\prime\) induces an isomorphism \(π^\sharp : A^\prime \to A\), which in turn induces an isomorphism \(\rho: \spec A \to \spec A^\prime\), then \(\pi = \rho\). First show this on the level of points; this is (surprisingly) the trickiest part. Then show \(\pi = \rho\) as maps of topological spaces. Finally, to show \(\pi = \rho\) on the level of structure sheaves, use the distinguished base. Feel free to use insights from later in this section, but be careful to avoid circular arguments. Even struggling with this exercise and failing (until reading later sections) will be helpful.
Proof:
We clearly have that every ring isomorphism \( A^\prime \to A \) induces an isomorphism \( \spec A \to \spec A^\prime \) as \( \spec (\cdot) \) is a contravariant functor. To see that an isomorphism \( \pi : \spec A \to \spec A' \) induces an isomorphism of rings \( A^\prime \to A \), notice that \( A = \OO_{\spec A} (\spec A) \) and \( A^\prime = \OO_{\spec A^\prime}(\spec A^\prime) \). Since \( \pi \) is an isomorphism of sheaves, we indeed get that the map \( \pi_* : \OO_{\spec A^\prime} \to \pi_* \OO_{\spec A} \) is an isomorphism of sheaves — let \( \phi = \pi_*(D(1)) :A^\prime \to A \) be the isomorphism of global sections. Letting \( \pp \in \spec A \) and \( \mathfrak{q} = \pi( \pp) \), by Exercise 2.2.I we get an isomorphism of stalks \( \phi_{\mathfrak{q}} : A^{\prime}_{\mathfrak{q}} \to A_\pp \) — pulling back prime ideals under \( \phi \), we get \( \pp = \phi^{-1}(\mathfrak{q}) \). But then \( \pi = \spec \phi \).
$$\tag*{$\blacksquare$}$$
Exercise 4.3.B:
Suppose \(f \in A\). Show that under the identification of \(D(f)\) in \(\spec A\) with \(\spec A_f\) (§3.5), there is a natural isomorphism of ringed spaces \((D(f), \OO_{\spec A\vert D(f)}) \cong (\spec A_f, \OO_{\spec A_f} )\). Hint: notice that distinguished open sets of \(\spec A_f\) are already distinguished open sets in \(\spec A\).
Proof:
Let \( \psi : A \to A_f \) be the localization map at \(f \in A \). By Exercise 4.1.A, \( \Gamma (D(f), \spec A) = A_f \) so that \( D(f) \) is homeomorphic to \( \spec A_f \). Following the hint, we wish to show that every distinguished open set in \( \spec A_f \) is a distinguished open set in \( \spec A\). If we suppose \( \pp \in A_f \) then \( D(\pp) = \{ \mathfrak{q} \in \spec A_f \mid \pp \not\subset \mathfrak{q} \} \) — however, since \( \spec A_f \cong D(f) \). Since \( \pp \) corresponds to a prime ideal \( \overline{\pp} = \psi^{-1} (\pp) \) in \( A \) not containing \(f\) and \( \psi \) preserves inclusion, we get \( D(\pp) = D(f) \cap D( \overline{\pp} ) = D( f \overline{\pp} ) \).
$$\tag*{$\blacksquare$}$$
Exercise 4.3.C:
If \(X\) is a scheme, and \( U \) is any open subset, prove that \( (U, \OO_X\vert_U) \) is also a scheme.
Proof:
We have by Vakil §2.2.8 that \( \OO_X \vert_U \) is indeed a sheaf of rings. Now if \( p \in U \), then as \( U \subset X \) there exists some open set \( V_p \subset X \) containing \( p \) such that \( (V_p, \OO_{V_p}) \cong \spec A^p, \OO_{\spec A^p} \) for some ring \( A^p \) ( I wasn't sure what notation to use since \( A_p \) looks like a local ring ). By quasicompactness Exercise 3.5.C, \( \spec A^p = \bigcup_{i=1}^n D(f_{i,p}) \) — in particular, \( p \in D(f_{j, p}) \) for some \( j \), so that \( D(f_{j, p}) \cap U \) is affine. In other words, \( U \) may be covered by distinguished open sets as well.
$$\tag*{$\blacksquare$}$$
Exercise 4.3.D:
Show that if \( X \) is a scheme, then the affine open sets form a base for the Zariski topology.
Proof:
Since every scheme \(X\) is covered by affine open schemes isomorphic to \( \spec A \), and the Zariski topology on \( \spec A \) has a base of distinguished open sets \( D(f) \), \(X\) may be covered by affine open schemes by Exercise 4.3.B.
$$\tag*{$\blacksquare$}$$
Exercise 4.3.E:
The disjoint union of schemes is defined as you would expect: it is the disjoint union of sets, with the expected topology: (thus it is the disjoint union of topological spaces), with the expected sheaf. Once we know what morphisms are, it will be immediate (Exercise 9.1.A) that (just as for sets and topological spaces) disjoint union is the coproduct in the category of schemes.
Show that the disjoint union of a finite number of affine schemes is also an affine scheme. (Hint: Exercise 3.6.A.)
(a first example of a non-affine scheme) Show that an infinite disjoint union of (nonempty) affine schemes is not an affine scheme. (Hint: affine schemes are quasicompact, Exercise 3.6.G(a). This is basically answered in Remark 3.6.6.)
Proof:
By Exercise 3.6.A we have that \( \coprod_{i=1}^n \spec A_i = \spec \prod_{i=1}^n A_i \), so that our sheaf is given on distinguished open subsets by localizations \(f \in \prod_{i=1}^n A_i \).
Consider \( X = \coprod_{n \geq 1} \spec A_n \) — then each \( \spec A_i \) is open in \( X \), however the collection \( \{ \spec A_i \} \) cannot have a finite subcover. Since every affine open set is quasicompact by Exercise 3.6.G, it must be the case that \( X \) quasicompact.
$$\tag*{$\blacksquare$}$$
Exercise 4.3.F:
Show that the stalk \( \OO_{\spec A} \) at the point \( [\pp] \) is the local ring \( A_\pp \).
Proof:
Let us denote \( X = \spec A \), and suppose \( \frac{f}{g} \) with \( g \notin \pp \). Then \( [\pp] \in D(g) \) so that \( \frac{f}{g} \in A_g = \OO_X(D(g)) \). Thus, we consider the map \( \phi : A_\pp \to \OO_{X,\pp} \) given by \( \frac{f}{g} \mapsto (D(f), \frac{f}{g}) \).
To see that the map is injective, suppose that \( \phi( \frac{f}{g} ) = 0 \) in \( \OO_{X,\pp} \). Then there exists some \( g^\prime \in A \) with \( [\pp] \in D(g^\prime) \subset D(g) \) so that the restriction \( \res{D(g)}{D(g^\prime)} \frac{f}{g} = 0 \). Then there exist some \( n \) with \( (g^\prime)^n f = 0 \). Since \( D(g^\prime) \subset D(g) \), there exists some \( m \) with \( (g^\prime)^m = ag \) for some \( a \in A \). Multiplying by large enough powers, we get that \( a^kg^{mk} f = 0 \) so that \( \frac{f}{g} = 0 \)
To see surjectivity, fix \( \psi \in \OO_{X,\pp} \). Then we may take some local representative with a distinguished open subset \( D(g) \ni \pp \). Then \( \psi = \frac{f^\prime}{g} \) for some \( f^{\prime} \in A_g \) — since \( g \in A - \pp \) we have \( \frac{f}{g} \in A_\pp \) with \( \phi(\frac{f}{g}) = \psi \).
$$\tag*{$\blacksquare$}$$
Exercise 4.3.G:
If \(f\) is a function on a locally ringed space \(X\), show that the subset of \(X\) where f vanishes is closed. (Hint: show that if \(f\) is a function on a ringed space \(X\), show that the subset of \(X\) where the germ of \(f\) is invertible is open.)
Show that if \(f\) is a function on a locally ringed space that vanishes nowhere, then \(f\) is invertible.
Though this was meant to be shown in part (a), we wish to show that the subset of \(X\) where the germ of \(f\) is invertible is open. Suppose \( f_\pp \) is invertible in \( \OO_{X,\pp} \) such that there exists \( g_\pp \in \OO_{X,\pp} \) with \( f_\pp g_\pp = 1_\pp \). Taking representatives, there exists some open set \( V \ni \pp \) and \( g \in \OO_{X}(V) \) with \( fg \vert_V = f\vert_V g\vert_V = 1_V \) — in fact, by the sheaf axiom we necessarily have that all stalks are invertible on \( V \), so that invertiblity is an open condition. Now if \( f \) doesn't vansish at \( \pp \), \( f_\pp \notin \mm_\pp \) so that \( f_\pp \) invertible ( if it weren't, \((f_\pp)\) would necessarily be contained in some maximal ideal which must be \( \mm_\pp \) ). We can conclude that if \( \textrm{Supp}\,(f) = \emptyset \), then \( f_\pp \) at all \( \pp \in X \) — using gluability on \(X\) we get that \( f \) is invertible.
$$\tag*{$\blacksquare$}$$
Section 4.4: Visualizing Schemes II: Nilpotents
Exercise 4.4.A:
Show that you can glue an arbitrary collection of schemes together. Suppose we are given:
Schemes \(X_i\) (as \(i\) runs over some index set \(I\), not necessarily finite,)
open subschemes \(X_{ij} \subset X_i\) with \(X_{ii} = X_i\),
isomorphisms \(f_{ij} : X_{ij} \to X_{ji}\) with \(f_{ii}\) the identity
such that
(the cocycle condition) the isomorphisms "agree on triple intersections", i.e., \(f_{ik}\vert_{X_{ij}\cap X_{ik}} = f_{jk}\vert_{X_{ji}\cap X_{jk}} \circ f_{ij}\vert_{X_{ij}\cap X_{ik}}\) (so implicitly, to make sense of the right side, \(f_{ij}(X_{ik} \cap X_{ij}) \subset X_{jk}\)).
(The cocycle condition ensures that \(f_{ij}\) and \(f_{ji}\) are inverses. In fact, the hypothesis that \(f_{ii}\) is the identity also follows from the cocycle condition.) Show that there is a unique scheme \(X\) (up to unique isomorphism) along with open subsets isomorphic to the \(X_i\) respecting this gluing data in the obvious sense. (Hint: what is \(X\) as a set? What is the topology on this set? In terms of your description of the open sets of \(X\), what are the sections of this sheaf over each open set?)
Proof:
We construct our topological space in the same way to the prior discussion, via \( X = (\coprod_i X_i) / \sim \) with \( X_{ij} \ni x \sim f_{ij}(x) \) for all \( i, j \in I \). Then the topology on \(X\) is precisely the quotient topology — in particular, if \( \overline{\sigma}_i : X_i \hookrightarrow \coprod_{i} X_i \) are the canonical inclusions and \( \sigma_i : X_i \hookrightarrow X \) are the induced maps on the quotient, then \( U \subset X \) is open in \( X\) if and only if \( \sigma^{-1}_i (U) \) is open in \( X_i \) for all \(i\).
To construct the sheaf \( \OO_X \) over \( X \), we want \( \OO_{X} \) to agree with \( \OO_{X_i} \) trivially when there are no intersections, and \( \OO_X \) to be independent of choice of \( i, j \) when there is some overlap on \( X_{i} \) and \( X_{j} \). However, the cocycle condition guarantees that if there were some other set \( X_k \) intersecting \( X_i \cap X_j \) (really there should be an isomorphism of open subsets here), our choice of isomorphisms is independent of indices \( i, j \). In other words, we may construct a sheaf over a base of open subsets on \(X\) by \( (\sigma_i)_*( \OO_{X_i} ) \) and this extends to a sheaf on \(X\) by §2.5.
$$\tag*{$\blacksquare$}$$
Exercise 4.4.B:
Show that the affine line with doubled origin is not affine. Hint: calculate the ring of global sections, and look back at the argument for \(\A^2 − \{(0, 0)\}\).
Proof:
The two schemes that we are gluing are \( X_1 = X_{11} = \spec k[t] \) and \( X_2 = X_{22} = \spec k[u] \), with open sets \( X_{12} = D(t) \cong \spec k[t, 1/t] \) and \( X_{21} = D(u) \cong \spec k[u, 1/u] \) and isomorphisms \( f_{12} : X_{12} \to X_{21} \), \( f_{21} : X_{21} \to X_{12} \) given by \( t \mapsto u \) and \( u \mapsto t \) respectively. By the previous exercise, this indeed gives us a scheme. Let \( s \in \Gamma(X, \OO_X) \) be a global section — by the construction of our structure sheaf from the previous exercise, this pulls back to global sections on \( X_1, X_2 \) i.e. polynomials \( g_1 \in k[t] \) and \( g_2 \in k[u] \) with the same coefficients. In other words, \( \Gamma(X, \OO_X) \cong k[t] \cong k[u]\). If we suppose to the contrary that \( X \cong \spec A \) for some ring \( A \), then by Exercise 4.3.A we would get some isomorphisms \( A \cong k[t] \) and \( A \cong k[u] \), which is impossible since \( D(u) \cong D(t) \) so that whatever point in \( A \) maps to \( u\) must also map to \(t\), and thus is 2-to-1.
$$\tag*{$\blacksquare$}$$
Exercise 4.4.C:
Do the same construction with \(\A^1\) replaced by \(\A^2\). You will have defined the affine plane with doubled origin. Describe two affine open subsets of this scheme whose intersection is not an affine open subset. (An "infinite-dimensional" version comes up in Exercise 5.1.J.)
Proof:
As before, we take \( X_1 = X_{11} = \spec k[x, y] \), \( X_2 = X_{22} = \spec k [u, v] \), \( X_{12} = D(x) \cup D(y) \cong \spec k[x, y, 1/x, 1/y] \) and \( X_{21} = D(u) \cup D(v) \cong k[u, v, 1/u, 1/v] \). Then we get an isomorphism \( f_{12} : X_{12} \to X_{21} \) by \( x \mapsto u, y \mapsto v \) (and vice-versa for inverse). By Exercise 4.4.A we obtain a scheme \(X\) which corresponds to two copies of the affine plane glued together.
By the same reasoning as the exercise above, \(X\) cannot be isomorphic to an affine scheme \(\spec A\) as that would imply by Exercise 4.3.A that we have an injective morphism but \( (x,y) \mapsto \pp \) and \( (u, v) \mapsto \pp \) for some prime \( \pp \).
$$\tag*{$\blacksquare$}$$
Exercise 4.4.D:
Check this, as painlessly as possible. (Possible hint: the triple intersection is affine; describe the corresponding ring.)
Proof:
For the sake of notation and sanity, we will restrict our attention to \( \P^2_k \), and consider the three affine open subsets
Our open sets \( X_{ij} \) in this case are just the intersections of these open sets (i.e. \( X_{02} = X_0 \cap X_2 \)). In this case, our isomorphism \( f_{ij} : X_{ij} \to X_{ji} \) is given by multiplication by \( \frac{x_i}{x_j} \). Then the cocycle condition is fairly immediate since \( f_{ij} \circ f_{jk} \) is multiplication by \( \frac{x_i}{x_j} \) composed with multiplication by \( \frac{x_j}{x_k} \), i.e. multiplication by
Show that the only functions on \(\P^n_k\) are constants ( \(\Gamma(\P^n_k,\OO) \cong k\)), and hence that \(P^n_k\) is not affine if \(n > 0\). Hint: you might fear that you will need some delicate interplay among all of your affine open sets, but you will only need two of your open sets to see this. There is even some geometric intuition behind this: the complement of the union of two open sets has codimension 2. But "Algebraic Hartogs’s Lemma" (discussed informally in §4.4.2, and to be stated rigorously in Theorem 11.3.11) says that any function defined on this union extends to be a function on all of projective space. Because we are expecting to see only constants as functions on all of projective space, we should already see this for this union of our two affine open sets.
Proof:
Suppose we have \( f(x_{1/0}, \dots, x_{n/0}) \) on \( D(x_0) \) and \( g( x_{0/1}, \dots, x_{n/1} ) \) on \( D(x_1) \). In order for these to define a global section, they must agree on overlaps; by applying our transition function \( f_{01} : D(x_0) \to D(x_1) \) given by \( x_{i/0}/x_{1/0} \mapsto x_{i/1} \) means we must have
But then it must be the case that \( \operatorname{deg}(f) = 0 \) so that \( f \) and \( g \) are constants.
$$\tag*{$\blacksquare$}$$
Exercise 4.4.F:
Show that if \(k\) is algebraically closed, the closed points of \(\P^n_k\) may be interpreted in the traditional way: the points are of the form \([a_0,\dots ,a_n]\), where the \(a_i\) are not all zero, and \([a_0,...,a_n]\) is identified with \([\lambda a_0, \dots , \lambda a_n]\) where \( \lambda \in k^\times \).
Proof:
Let \( P^n_k \) denote \( (k^{n+1} - \{ (0, \dots, 0) \}) / \sim \) (where \( \sim \) is the above relation). For any \( (a_0, \dots, a_n) \neq 0 \) in \( k^{n+1} \), there must exist some \( a_i \neq 0 \) so we define \( \phi( a_0, \dots, a_n ) = [( \frac{x_0}{x_i} - \frac{a_0}{a_i}, \dots, \frac{x_n}{x_i} - \frac{a_n}{a_i} )] \); since \( \frac{a_j}{a_i} = \frac{\lambda a_j}{\lambda a_i} \) for any \( \lambda \neq 0 \), this map preserves equivalence classes under \( \sim \) so it uniquely factors through to \( \overline{\phi} : P^n_k \to \P^n_k \).
We additionally wish to show that \( \overline{\phi} \) does not depend on choice of \( a_i \neq 0 \). If we suppose that \( a_j \neq 0 \) as well, then
Since \( \lambda = \frac{a_i}{a_j} \neq 0 \), our map \( \overline{\phi} \) is well-defined on intersections.
To construct the inverse, since \( k \) is algebraically closed our points are all of the form \( [ \frac{x_0}{x_i} - a_0, \dots, 1, \dots \frac{x_n}{x_i} - a_n ] \) on some affine open \( D(x_i) \). Then we define \( \overline{\psi} : \P^n_k \to P^n_k \) by
Section 4.5: Projective Schemes, and the \( \proj \) Construction
Exercise 4.5.A:
Consider \(\P^2_k\), with projective coordinates \(x_0, x_1\) and \(x_2\). (The terminology "projective coordinate" will not be formally defined until §4.5.8, but you should be able to solve this problem anyway.) Think through how to define a scheme that should be interpreted as \(x^2_0 + x^2_1 − x^2_2 = 0\) "in \( \P^2_k \)". Hint: in the affine open subset corresponding to \(x_2 \neq 0\),it should (in the languageof 4.4.9) be cut out by \(x^2_{0/2} + x^2_{1/2} − 1 = 0\), i.e., it should "be" the scheme \(\spec k[x_{0/2} , x_{1/2} ] / (x^2_{0/2} + x^2_{1/2} − 1) \). You can similarly guess what it should be on the other two standard open sets, and show that the three schemes glue together.
Proof:
From the hint, there is very little to do here. We consider the three affine schemes
Then our transition functions are the same as in the case of projective space (as we are simply considering a closed subset of projective space), so that these glue together in a natural way by Exercise 4.4.A.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.B:
More generally, consider \( \P^n_A \) with projective coordinates \(x_0, \dots , x_n\). Given a collection of homogeneous polynomials \(f_i \in A[x_0, \dots, x_n]\), make sense of the scheme "cut out in \(P^n_A\) by the \(f_i\)." (This will later be made precise as an example of a "vanishing scheme", see Exercise 4.5.P.) Hint: you will be able to piggyback on Exercise 4.4.D to make this quite straightforward.
Proof:
Given any homogeneous polynomial \( f \in A[x_{0/i}, \dots, x_n] \) of degree \(d\), we have that for any unit \( u \in A\), \(f[ux_0, \dots, u x_d] = u^{d} f[x_0, \dots, x_n] \) — thus, \( f \) factors through the quotient \( [x_0, \dots, x_n] \sim [ux_0, \dots, ux_n] \). Notice that in the ring \( A[x_0, \dots, x_n] \) our indeterminites \( x_i \) are non-nilpotent, so that on the affine open sets \( D(x_i) \) we have
is a polynomial (no longer homogeneous) of degree \(d\) in \( A[x_{0/i}, \dots, x_{n/i}] \) (which we will denote \( f_i \). In fact, when dividing through by \( x_i^{d} \) above, any term in \( x_{i/i} \) should become constant, so this is the same as a polynomial in \( A[x_{0/i}, \dots, x_{n/i}] / (x_{i/i} - 1) \). In particular, if \( f \) vanishes at some point \( P = [x_0, \dots, x_n] \), then it must also vanish at each \( [u x_0, \dots, u x_n] \) (for any unit \(u \in A\)). Thus, if \( P \in D(x_i) \) for some \(i\), we must necessarily have that \( f_i \) vanishes at \( P \). In fact, if \( D(x_j) \) is another affine open set which contains \( P \), then
$$
\frac{x_i^d}{x_j^d} f_j = f_i
$$
so that our \( f_i \) and \( f_j \) necessarily glue under the transition functions.
As mentioned in Exercise 3.4.I, the vanishing set \(V (I) \) for \( I \subset B \) may be interpreted as \( \spec B / I \). Thus, we consider the locus of our homogeneous polynomial \( f \) in \( \P^n_A \) as the gluing together of affine sets \( \spec A[x_0, \dots, x_n] / (f_i) \) — by an identical argument to Exercise 4.4.D, these glue together using our standard transition functions.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.C:
Show that an ideal \(I\) is homogeneous if and only if it contains the degree \(n\) piece of each of its elements for each \(n\). (Hence \(I\) can be decomposed into homogeneous pieces, \(I = \bigoplus I_n\), and \(S_\bullet / I\) has a natural \(\Z\)-grading. This is the reason for the name homogeneous ideal.)
Show that the set of homogeneous ideals of a given \(\Z\)-graded ring \(S_\bullet\) is closed under sum, product, intersection, and radical.
Show that a homogeneous ideal \(I \subset S_\bullet\) is prime if \(I \neq S•\), and if for any homogeneous \(a, b \in S_\bullet\), if \( ab \in I \), then \(a \in I\) or \(b \in I\).
Proof:
For the forward direction, suppose \( I \) a homogeneous ideal so that it is by definition generated by some homogeneous elements \( f_i \in S_\bullet \). We may suppose that each \( f_i \in S_{d_i} \) for some \( d_i \in \Z \). Then for any element \( f \in I \), \( f \) is a finite sum over our generators, so we may write
$$
f = \sum_{i=1}^N a_i f_i
$$
for sufficiently large \(N\) and some \( a_i \in S_\bullet \). Now each \( a_i \) may be written \( a_i = \sum_{d \in \Z} (a_i)_{(d))} \). Thus, the degree \( n \) component of \( f \) may be expressed as
which is clearly generated by our \(f_i\) so that it is an element of \( I \).
Conversely, if \( I \) is generated by some \( \{ f_i \}_{i \in J} \) (not necessarily finite), then each \( f_i \) may be decomposed into its degree \( n \) components \( f_i = \bigoplus_{d \in \Z} (f_i)_{(d)} \). By assumption, \( (f_i)_{(d)} \in I \) for each \( d \in \Z \) and \(i \in J\), so that \( I \) may be generated by \( \{ (f_i)_{(d)} \mid i \in J, d \in \Z \} \) — thus \( I \) is homogeneous.
Suppose \( I \) and \( J \) are homogeneous ideals in \( S_\bullet \) so that \( I \) is generated by some homogeneous elements \( I = \{ f_i \} \) and \( J = \{ g_j \} \). Then \( I + J \) is generated by \( \{ f_i\} \cup \{g_j\} \) and \( IJ \) is generated by \( \{ f_ig_j \} \) — the former is clearly homogeneous, and by the definition of a graded ring multiplication must respect grading, so \( f_i g_j \in S_{\deg f_i \cdot \deg g_j} \) and thus \( IJ \) homogeneous.
To see that \( I \cap J \) is homogeneous, we utitlize part (a) above. If \( f = \sum_{d \in \Z} f_{(d)} \in I \cap J \), then as \( I \) and \( J \) are homogeneous ideals every degree \( d \) component of \( f \) is contained in \( I, J \) and thus \( I \cap J \). Then by part (a) we have that \( I \cap J \) is homogeneous.
Lastly, we wish to show that \( \sqrt{I} \) is homogeneous by again utilizing part (a) above. Suppose \( f \sqrt{I} \) so that there exists some \( n \geq 1 \) with \( f^n \in I \). As \( f \) is an element in the direct sum, only finitely many of its homogeneous terms / components are non-zero — taking \( d_0\) to be the smallest such degree and \( N \) the largest, we have \( f = \sum_{d = d_0 }^N f_{(d)}\). Then \( f^n \) has lowest degree homogeneous term \( f^n_{(d_0)} \) of degree \( n \cdot d_0 \). As \( I \) is homogeneous, by part (a) \( f^n_{(d_0)} \in I \) so that \( f_{(d_0)} \in \sqrt{I} \) by definition. As \( \sqrt{I} \) is indeed an ideal, this implies that \( f - f_{(d_0)} \in \sqrt{I} \). Proceeding inductively, we can see that \( f_{(d)} \in \sqrt{I} \) for each degree \( d \) term of \( f \). Since \( f \in I\) was arbitrary, by part (a) we have \( \sqrt{I} \) is homogeneous.
Assume to the contrary we have \( fg \in I \) with \( f \notin I \) and \( g \notin I \), and write \( f = \sum_{ d = a }^n f_{(d)} \), \( g = \sum_{d = b}^m g_{(d)} \). Suppose we have chosen \( f, g \) with \( \deg f + \deg g \) minimal. Now the leading term of \( fg \) is \( f_{(n)}g_{(m)} \) — by part (a) we know that \( f_{(n)}g_{(m)} \in I \), and by assumption it must be the case that \( f_{(n)} \in I \) or \( g_{(m)} \in I \). If we assume without loss of generality it is the former, notice that it must be the case that \( f - f_{(n)} \notin I \) (as otherwise if \(f = (f - f_{(n)}) + f_{(n)} \) would be the sum of elements in \( I \) and thus in \(I\)). Then \( (f - f_{(d)}) \cdot g = fg - f_{(d)}g \in I \), though neither of the components are — however, this contradicts our minimality assumption of the degrees of \(f \) and \(g\).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.D:
Show that a graded ring \(S_\bullet\) over \(A\) is a finitely generated graded ring (over \(A\)) if and only if \(S_\bullet\) is a finitely generated graded \(A\)-algebra, i.e., generated over \(A = S_0\) by a finite number of homogeneous elements of positive degree. (Hint for the forward implication: show that the generators of \(S_+\) as an ideal are also generators of \(S_\bullet\) as an algebra.)
Show that a graded ring \(S_\bullet\) over \(A\) is Noetherian if and only if \(A = S_0\) is Noetherian and \(S_\bullet\) is a finitely generated graded ring.
Proof:
For the forward direction, suppose that \( S_\bullet \) is a finitely-generated graded ring over \(A\) so that the irrelevant ideal \( S_+ \) is finitely generated by some \( f_1, \dots, f_n \), and let \( f \in S \) be arbitrary. Using the grading on \(S_\bullet\), \(f\) may be written as \( f = a_0 + s_1 f_1 + \dots s_n f_n \) where \( a_0 \in S_0 = A \) and \( s_1, \dots, s_n \in S_{\bullet} \). Now for each \( s_i \), if \( d_i = \deg(s_i) > 0 \) then \( s_i \in S_+ \) so that it can be expressed as a linear combination \(f_1, \dots, f_n \) with coefficients having degree strictly less than \(d_i\). Proceeding by induction, it follows that \(f\) may be written as a polynomial in the \( f_1, \dots, f_n \) so that \( S_\bullet \) is a finitely-generated \(A\)-algebra.
Conversely, if \( S_\bullet \) is a finitely generated \(A\)-algebra, with generators \( f_1, \dots, f_n \), then \( S_+ \) is generated by those \( f_i \) with positive degree.
For the forward direction, suppose that \( S_\bullet \) is Noetherian. Then every ideal is finitely generated, so in particular \( S_+ \) is finitely generated (i.e. \( S_+ \) is a finitely generated graded ring over \( A \)). Additionally, \( 0 \to S_+ \to S_\bullet \to S_0 \to 0 \) exact, so by Exercise 3.6.V \( S_0 \) is Noetherian.
Conversely, suppose that \( A = S_0\) is Noetherian and \( S_\bullet \) is a finitely generated ring over \( A \). By part (a) above, this implies that \( S_\bullet \) is a finitely generated \(A\)-algebra. Then there is some surjection from \( S_\bullet \to A[x_0, \dots, x_m] \) for some \( m \geq 1 \). Since \( A \) is Noetherian, \( A[x_0, \dots, x_m] \) is necessarily Noetherian (Hilbert Basis Theorem), and thus \( S_\bullet \) is as well.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.E:
Suppose \( f \in S_+ \) is homogeneous.
Give a bijection between the prime ideals of \(((S_\bullet)_f)_0\) and the homogeneous prime ideals of \((S_\bullet)_f\). Hint: Avoid notational confusion by proving instead that if \(A\) is a \(\Z\)-graded ring with a homogeneous invertible element \(f\) in positive degree, then there is a bijection between prime ideals of \(A_0\) and homogeneous prime ideals of \(A\). Using the ring map \(A_0 \to A\), from each homogeneous prime ideal of \(A\) we find a prime ideal of \(A_0\). The reverse direction is the harder one. Given a prime ideal \(P_0 \subset A_0\), define \(P \subset A\) (a priori only a subset) as \(\oplus Q_i\), where \(Q_i \subset A_i\), and \(a \in Q_i\) if and only if \(a^{\deg f}/f_i \in P_0\). Note that \(Q_0 =P_0\). Show that \(a\in Q_i\) if and only if \(a_2 \in Q_{2i}\); show that if \(a1,a2 \in Q_i\) then \(a_{1}^2 +2a_1a_2 +a_2^2 \in Q_{2i}\) and hence \(a_1 + a_2 \in Q_i\); then show that \(P\) is a homogeneous ideal of \(A\); then show that \(P\) is prime.
Interpret the set of prime ideals of \( ((S_\bullet)_f)_0\) as a subset of \(Proj S_\bullet\).
Proof:
For the forward direction, let \( \pp \subset A \) be a prime ideal and let \( i : A_0 \hookrightarrow A \) denote the inclusion morphism; by Exercise 3.2.M we know that \( i^{-1}(P) \) is necessarily a prime ideal in \( A_0 \)
In the reverse direction, we utilize the hint given — that is, given a prime ideal \( \pp_0 \subset A_0 \), we define \( \pp = \bigoplus Q_i \) where \( a \in Q_i \Leftrightarrow a^{\deg f} / f^i \in \pp_0 \). Obviously if \( a \in Q_i \), then \( a^{\deg f} / f^i \in \pp_0 \) and thus \( a^{ 2\deg f} / f^{2i} \in \pp_0 \) so that \( a^2 \in Q_{2i} \); the converse is proved the same way. By homogeneity, if \( a \in Q_i \) and \( b \in Q_j \), then
so that \( ab \in Q_{i+j} \). It follows that if \( a_1, a_2 \in Q_i \), then \( 2a_ia_j \in Q_{2i} \) and thus \( (a_1 + a_2)^2 = a_1^2 + 2a_1 a_2 + a_2^2 \in Q_{2i} \) which implies \( a_1 + a_2 \in Q_i \). It follows that \( \pp \) is necessarily an ideal; it remains to show that it is prime. By the preceding exercise, it suffices to check only on homogeneous elements. Thus, if we have some \( a \in A_i, b \in A_j\) with \( ab \in P \) (in particular \( ab \in Q_{i + j} \) by looking at degrees), then
However, since \( \pp_0\) is assumed to be prime, this implies that either \( \frac{a^{\deg f}}{f^i} \in \pp_0 \) or \( \frac{b^{\deg f}}{f^j} \in \pp_0 \) from which the result follows.
To see the two are inverse, notice that if \( \pp \subset A \) is a homogeneous prime ideal, then \( a \in \pp_i \Leftrightarrow \pp_i^{\deg f} \in \pp_{(\deg f) i} \Leftrightarrow a^{\deg f}/f^i \in \pp_0 \)
By part (a) above, the prime ideals of \( ((S_\bullet)_f)_0 \) correspond to the homogeneous prime ideals of \( (S_\bullet)_f \), which we identify with \( D(f) \subset \proj S_\bullet \)
$$\tag*{$\blacksquare$}$$
Exercise 4.5.F:
Show that \(D(f)\) "is" (or more precisely, "corresponds to") the subset \(\spec((S_\bullet)_f)_0\) you described in Exercise 4.5.E(b). For example, in §4.4.9, the \(D(x_i)\) are the standard open sets covering projective space.
Proof:
Using the definition from the preceeding paragraph, \( \pp \) is an element of \( D(f) \) iff \( \pp \) is not an element of \( V(f) \) iff \( f \notin \pp \) and \( \pp \) does not contain \( S_+ \). By an identical argument to the affine case, this corresponds precisely to the (now homogeneous) prime ideals in the localization \( (S_\bullet)_f \), which by the preceding exercise are in correspondence with the prime ideals in \( ((S_\bullet)_f)_0 \). Therefore the two may be identified (up to a bijective correspondence).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.G:
Verify that the projective distinguished open sets \(D(f)\) (as \(f\) runs through the homogeneous elements of \(S_+\)) form a base of the Zariski topology.
Proof:
Similar to the affine case, if \( U \subset \proj S_\bullet \) is an open subset in the Zariski topology then \( \proj S_\bullet \backslash U = V(T)\) for some homogeneous ideal \(T\). Then \( V(T) = \bigcap_{f \in T} V(f) \) so that \( U = \bigcup_{f \in T} D(f) \).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.H:
Suppose \(I\) is any homogeneous ideal of \(S_\bullet\) contained in \(S_+\), and \(f\) is a homogeneous element of positive degree. Show that \(f\) vanishes on \(V(I)\) (i.e., \(V(I) \subset V(f)\)) if and only if \(f^n \in I\) for some n. (Hint: Mimic the affine case; see Exercise 3.4.J.) In particular, as in the affine case (Exercise 3.5.E), if \(D(f) \subset D(g)\), then \(f^n \in (g) \) for some \(n\), and vice versa. (Here \(g\) is also homogeneous of positive degree.)
If \(Z \subset \proj S_\bullet\), define \(I(Z) \subset S_+\). Show that it is a homogeneous ideal of \( S_\bullet \). For any two subsets, show that \( I(Z_1 \cup Z_2) = I(Z_1) \cap I(Z_2) \).
For any subset \(Z \subset Proj S_\bullet\), show that \(V(I(Z)) = \overline{Z}\).
Proof:
Similar to the affine case (as indicated in the hint), we know that for each \( I \subset \pp \in \proj S_\bullet \) contained in \( S_+ \), \( V(I) \subset V(f) \Leftrightarrow f \in \pp \). Since \( I \subset \pp \) was arbitrary, this tells us \( V(I) \subset V(f) \Leftrightarrow f \in \bigcap_{I \subset \pp} \pp \). Thus, we wish to show that \( \sqrt{I} = \bigcap_{I \subset \pp} \pp \).
For the forward direction, suppose that \( g \in \sqrt{I} \) — by Exercise 4.5.C, we may suppose without loss of generality that \(g\) is in fact homogeneous, as we may otherwise consider each degree \( n\) homogeneous component. Then there exists some \( n \geq 1 \) with \( g^n \in I \). Thus, for every homogeneous prime \( \pp \) (not containing \( S_+ \)) we have \( g^n \in \pp \). Using Exercise 4.5.C(c), this implies that \( g \in \pp \) for each prime \( \pp \) and thus \( g \in \bigcap_{I \subset \pp} \pp \).
Conversely, suppose \( g \in \bigcap_{I \subset \pp} \pp \) — again, we may assume without loss of generality that \( g \) is homogeneous. Suppose to the contrary that \( g^n \notin I \) for any \( n \) so that the homogeneous prime \( (g) \) is disjoint from \( I \). By Zorn's lemma, this implies the existence of some ideal \( \qq \) maximal among ideals disjoint from \( (g) \). We must have \( \qq \) is prime, since otherwise we could find homogeneous \( ab \in \qq \) with \( a, b \notin \qq \) — by maximality of \( \qq \) we would have \( \qq + (a) \) and \( \qq + (b) \) intersect \( (g) \) so that \( g^m = q_1 + ra \), \( g^n = q_2 + sb\) for some \( m, n \geq 1 \) — since \( ab \in \qq \) we must also have \( g^{m + n} \in \qq \)
As in the affine case, we define \( I(Z) = \bigcap_{\pp \in Z} \pp \) — however, we now require that our prime (homogeneous) ideals now do not contain \( S_+ \) (notice this implies \( I(Z) \subset S_+ \)). By Exercise 4.5.C(b), this \( I(Z) \) is also necessarily a homogeneous ideal. It follows from definition that \( I(Z_1 \cup Z_2) = I(Z_1) \cap I(Z_2) \) since we may possibly repeat any prime ideals \( \pp, \pp^\prime \) in the intersection of \( Z_1 \cap Z_2 \) without affecting \( \bigcap_{\pp \in Z_1} \pp \cap \bigcap_{\pp^\prime \in Z_2} \pp^\prime \).
For reverse inclusion, by definition we have that for all \( \pp \in Z \), \( I(Z) \subset \pp \). Since \( V(I(Z)) \) is the union of all homogeneous prime ideals containing \( I(Z) \), this implies that \( Z \subset V(I(Z)) \). Since \( V(I(Z)) \) is closed, this implies \( \overline{Z} \subset V(I(Z)) \). Conversely, since \( \overline{Z} \) is closed we have that \( \overline{Z} = V(J) \supset Z \) for some homogeneous ideal \( J \) (by definition of the Zariski topology). Then \( J \subset I(V(J)) \subset I(Z) \) so that \( V(I(J)) \subset V(J) = \overline{Z} \) as desired.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.I:
Fix a graded ring \(S_\bullet\), and a homogeneous ideal \(I\). Show that the following are equivalent.
\(V(I) = \emptyset\)
For any \(f_i\) (as \(i\) runs through some index set) generating \(I\), \( \bigcup D(f_i) =
\proj S_\bullet\).
\( \sqrt{I} \supset S_+ \)
Proof:
Notice that (a) and (b) are clearly equivalent from definitions, since if \( I = (f_i) \) so that \( V(I) = \bigcap V(f_i) \), then
To see that (b) implies (c) suppose that \( f \in S_+ \) so that \( D(f) \subset \bigcup D(f_i) \) and thus \( V(I) = V( (f_i)) \subset V(f) \). By the previous exercise, this is equivalent to \( f \in \sqrt{I} \). Since \( f \in S_+\) was arbitrary, this tells us \( S_+ \subset \sqrt{I} \).
Lastly, to see that (c) implies (a), suppose \( \pp \in V(I) \) so that \( \pp \) is a homogeneous prime ideal containing \( I \) that doesn't contain \( S_+ \). Since prime ideals are radical and radicalization preserves containment, we have \( \sqrt{I} \subset \pp \). By assumption, however, this implies \( S_+ \subset \sqrt{I} \subset \pp \) — a contradiction.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.J:
Suppose some homogeneous \(f \in S_+\) is given. Via the inclusion
$$
D(f) = \spec ((S_\bullet)_f)_0
$$
of Exercise 4.5.F, show that the Zariski topology on \( \proj S_\bullet \) restricts to the Zariski topology on \( \spec ((S_\bullet)_f)_0 \)
Proof:
Let \( R \) denote the graded ring \( (S_\bullet)_f \), and suppose \( V(I_0) \subset \spec R_0 \) is a closed subset. Since \( V(I) = V(\sqrt{I}) \) we may assume without loss of generality \( I \) is radical so that \( I_0 = \bigcap_{ I_0 \subset \pp_0 } \pp_0 \subset R_0 \) (where the prime ideals are not necessarily homogeneous here). From our corresponence in Exercise 4.5.E there exists some homogeneous \( \pp \in R \) for each \( \pp_0 \in R_0 \) such that \( \pp \cap R_0 = \pp_0 \) — if we let \( I \) denote the intersection of all such \( \pp \), then \( I \) is again homogeneous by Exercise 4.5.C(b) and more importantly \( I \cap R_0 = I_0 \). By part (b) of Exercise 4.5.E the (not necessarily homogeneous) prime ideals of \( R_0 \) corresponds to \(D(f)\) so that \( V(I_0) = V(I) \cap D(f) \).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.K:
If \( f, g \in S_+ \) are homogeneous and nonzero, describe an isomorphism between \( \spec((S_\bullet)_{fg})_0 \) and the distinguished open subset \( D(g^{\deg f}/f^{\deg g}) \) of \( \spec ((S_\bullet)_f)_0 \).
Proof:
Since \( f \) is invertible in \( (S_\bullet)_{fg} \), by the universal property of the localization \( (S_\bullet)_f \) there exists a unique morphism \( \phi : (S_\bullet)_f \to (S_\bullet)_{fg} \) so that \( \phi(\frac{h}{f^n}) = \frac{hg^n}{(fg)^n} \) and thus \( \phi \) preserves grading. Thus, we get a morphism \( \phi_0 : ((S_\bullet)_f)_0 \to ((S_\bullet)_{fg})_0 \). Since
has inverse \( \frac{ f^{\deg f + \deg g} }{ (fg)^{\deg g} } \), its degree 0 parts become invertible and we may apply the universal property of localization again to get a unique morphism \( \phi_0^\prime : \frac{ g^{\deg f} }{ f^{\deg g} } \subset \spec ((S_\bullet)_f)_0 \to \spec ((S_\bullet)_{fg})_0 \). We will not show that this is necessarily an isomorphism, as it can be easily checked.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.L:
By checking that these gluings behave well on triple overlaps (see Exercise 2.5.D), finish the definition of the scheme \( \proj S_\bullet \) .
Proof:
If we now suppose \( f, g, h \in S_+\) nonzero homogeneous polynomials, we have by the previous exercise that \( \spec ((S_\bullet)_{fgh})_0 \) is isomorphic to \( D( \frac{ h^{\deg f + \deg g} }{(fg)^{\deg h}}) \) in \( \spec ((S_\bullet)_{fg}) \), \( D( \frac{ g^{ \deg f + \deg h } }{ (fh)^{\deg g} } ) \) in \( \spec ((S_\bullet)_{fh}) \), and \( D( \frac{ f^{\deg h + \deg g} }{ (gh)^{\deg f} } ) \) in \( \spec ((S_\bullet)_{gh}) \). Each of \( f, g \) and \(h\) are invertible in \( \spec ((S_\bullet)_{fgh}) \) (since, for example, \( \frac{1}{f} = \frac{gh}{fgh} \)), and thus the obvious isomorphism between the first and the second distinguished open sets is multiplication by \( \frac{ g^{\deg f + 2\deg h} }{ h^{\deg f + 2 \deg g}} \) — the isomorphisms are similar. Each of these isomorphisms induces an isomorphism on the degree 0 part of our graded schemes, so that these gluings agree on triple overlaps
Since we have well-defined structure sheaves for each distinguished open set \( \spec ((S_\bullet)_{ab})_0 \), by Exercise 2.5.D there exists a sheaf \( \OO_{\proj S_\bullet} \) which restricts to \( \OO_{\spec (S_bullet)_{x_i}} \) for each of our homogenous projective coordinate \( f, g, h = x_i \).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.M:
(Some will find this essential, others will prefer to ignore it.) (Re)interpret the structure sheaf of \(\proj S_\bullet\) in terms of compatible germs.
Proof:
I assume this is interpreted as \( \OO_{\proj S_\bullet} \) consisting of collections of germs \( (f_i / s_i)_{i \in I} \) where each \( f_i / s_i \in \OO_{\spec ((S_\bullet)_{s_i})_0} \) and the \( (s_i) \) generate \( S_\bullet \) (i.e. if \( S_\bullet = A[x_0, \dots, x_n] \) we take the \( s_i = x_i \) ) which are compatible in the sense that they agree on overlap.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.M:
(Some will find this essential, others will prefer to ignore it.) (Re)interpret the structure sheaf of \(\proj S_\bullet\) in terms of compatible germs.
Proof:
I assume this is interpreted as \( \OO_{\proj S_\bullet} \) consisting of collections of germs \( (f_i / s_i)_{i \in I} \) where each \( f_i / s_i \in \OO_{\spec ((S_\bullet)_{s_i})_0} \) and the \( (s_i) \) generate \( S_\bullet \) (i.e. if \( S_\bullet = A[x_0, \dots, x_n] \) we take the \( s_i = x_i \) ) which are compatible in the sense that they agree on overlap.
$$\tag*{$\blacksquare$}$$
Exercise 4.5.N:
Check that this agrees with our earlier construction of \( \P^n_A \) (§4.4.9). (How do you know that the \(D(x_i)\) cover \( \proj A[x_0, \dots, x_n] \)?)
Proof:
This effectively follows from the discussion preceeding Exercise 4.5.E (in the text). We originally defined \( \P^n_A \) as \( n + 1 \) copies of \( \A_A^n \), each with local coordinates \( x_{0/i}, \dots, x_{n/i} \) and \( x_{i/i} = 1 \). From the discussion mentioned above, in the \( \proj \) construction we take \( S_\bullet = A[x_0, \dots, x_n] \) graded by degree and \( f = x_i \). Then \( \spec (S_\bullet)_f = \spec A[x_0/ x_i, \dots, x_n/ x_i] \) and taking the zero-degree components, we obtain scalar multiples of \( x_i \) on the \(i^{th} \) affine open set. On intersections, this agrees with \( D(\frac{x_i}{x_j}) \) by Exercise 4.5.K.
Our \( D(x_i) \) must necessarily cover \( \proj A[x_0, \dots, x_n] \) since if there were some point \( P \in \P^n_A \) not contained in any \( D(x_i) \), then \( P \) would vanish on \( S_+ \) implying that the corresponding prime ideal \( \pp \) contains \( S_+ \) — a contradiction from our definition of \( \proj\).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.O:
Suppose that \(k\) is an algebraically closed field. We know from Exercise 4.4.F that the closed points of \(\P^n_k\) , as defined in §4.4.9, are in bijection with the points of "classical" projective space. By Exercise 4.5.N, the scheme \(\P^n_k\) as defined in §4.4.9 is isomorphic to \(\proj k[x_0, \dots, x_n ]\). Therefore, each point
\([a_0, \dots , a_n]\) of classical projective space corresponds to a homogeneous prime ideal of \(k[x_0 , \dots , x_n ]\). Which homogeneous prime ideal is it?
Proof:
If \( [a_0 : \dots : a_n] \) is a closed ("classical") point of \( \P_k^n \), then the corresponding ideal on \( \spec (k[x_0, \dots, x_n]_{x_i})_0 \) is \( \left( \frac{ x_0 }{x_i} - \frac{a_0}{a_i}, \dots, \frac{x_n}{x_i} - \frac{a_n}{a_i} \right) \). In \( S_\bullet = k[x_0, \dots, x_n]\), this corresponds to the homogeneous prime ideal \( (a_ix_0 - a_0x_i, \dots, a_ix_n - a_nx_i) \).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.P:
If \(S_\bullet\) is generated in degree 1, and \(f \in S_+\) is homogeneous, explain how to define \(V(f)\) "in" \( \proj S_\bullet \), the vanishing scheme of \(f\). (Warning: \(f\) in general isn’t a function on \(\proj S_\bullet\). We will later interpret it as something close: a section of a line bundle, see for example §14.1.2.) Hence define \(V(I)\) for any homogeneous ideal \(I\) of \( S_+ \). (Another solution in more general circumstances will be given in Exercise 13.1.I.)
Proof:
Since \( f \) is homogeneous, \( S_\bullet / f \) is a graded ring so we define \( V_+(f) := \proj S_\bullet / (f) \).
On every affine open set \( U = \spec R\), \( V_+(f) \cap U \) is clearly a closed subset by Exercise 4.5.J so that \( V_+(f) \cap U = \spec R / J\) for some ideal \( J \).
$$\tag*{$\blacksquare$}$$
Exercise 4.5.Q:
Suppose \(k\) is algebraically closed. Describe a natural bijection between one-dimensional subspaces of \(V\) and the closed points of \(\P V\). Thus this construction canonically (in a basis-free manner) describes the one-dimensional subspaces of the vector space \(V\).
Proof:
If \( x \in \P V \) is a closed point corresponding to a homogeneous maximal ideal \( \mm \subset \textrm{Sym}^\bullet\,V^\vee \), which must correspond to the span of linear forms by maximality. By linearity, this corresponds to the vanishing set of some one-dimensional subspace. Conversely, any one dimensional subspace corresponds to the vanishing set of some linear forms. It is a standard linear algebra fact that this is independent of choice of dual basis.