--- title: Definitions --- # LIST OF ALGEBRAIC GEOMETRY DEFINITIONS AND THEOREMS ## Definitions :::{.definition title="abelian category" .flashcard} An abelian category is a category $\mathscr{A}$ such that (i) for all objects $A, B, \operatorname{Hom}(A, B)$ is an abelian group, (ii) composition distributes: $f \circ(g+h)=(f \circ g)+(f \circ h)$ and $(f+g) \circ h=(f \circ h)+(g \circ h)$, (iii) biproducts $A \oplus B$ exist (hence so do finite sums), (iv) for every $f: A \rightarrow B$, $\ker f$ and $\coker f$ exist, (v) every monomorphism is the kernel of its cokernel, and every epimorphism is the cokernel of its kernel (the first isomorphism theorem), (vi) every morphism $f: A \rightarrow B$ can be factored as $f=g \circ h$ where $h$ is an epimorphism and $g$ is a monomorphism (surjecting onto the image, which injects into the target). ::: :::{.definition title="acyclic" .flashcard} Let $\mathscr{A}$ be an abelian category with enough injectives and $F: \mathscr{A} \rightarrow \mathscr{B}$ a left exact additive functor, i.e. precisely the conditions necessary for the right derived functors $R^{i} F$ to exist. An object $A \in \mathscr{A}$ is $F$-acyclic (often just "acyclic" when the functor is clear) if the higher right derived functors vanish, $$ R^{i} F(A)=0 \text { for all } i>0 . $$ This also works for left derived functors and $F$ contravariant. It's worth noting that injective (resp. projective) objects are acyclic, and in fact acyclic resolutions can be used to compute derived functors. ::: :::{.definition title="additive functor" .flashcard} A (covariant) functor $F: \mathscr{A} \rightarrow \mathscr{B}$, where $\mathscr{A}, \mathscr{B}$ are abelian categories, is additive if the induced map $\operatorname{Hom}_{\mathscr{A}}\left(A, A^{\prime}\right) \rightarrow \operatorname{Hom}_{\mathscr{B}}\left(F(A), F\left(A^{\prime}\right)\right)$ is a homomorphism of abelian groups. That is, $F(f+g)=F(f)+F(g)$ for all $f, g \in \operatorname{Hom}\left(A, A^{\prime}\right)$. ::: :::{.definition title="adjoint" .flashcard} Suppose $\mathscr{A}$ and $\mathscr{B}$ are categories with functors $F: \mathscr{A} \rightarrow \mathscr{B}$ and $G: \mathscr{B} \rightarrow \mathscr{A} . F$ and $G$ are adjoint if there is a natural bijection $$ \tau_{A B}: \operatorname{Mor}_{\mathscr{B}}(F(A), B) \rightarrow \operatorname{Mor}_{\mathscr{A}}(A, F(B)) $$ for all $A \in \mathscr{A}$ and $B \in \mathscr{B}$. We say $F$ is left adjoint and $G$ is right adjoint. The adjective "natural" here refers to the fact that if $f: A \rightarrow A$ is a map in $\mathscr{A}$ then we have ![](https://cdn.mathpix.com/cropped/2022_10_17_cd4f806c438565f8bbceg-02.jpg?height=198&width=638&top_left_y=570&top_left_x=798) and a similar diagram given a map $g: B \rightarrow B^{\prime}$. ::: :::{.definition title="affine" .flashcard} A scheme $X$ is affine if $X \simeq \operatorname{Spec} A$, as ringed spaces, for some ring $A$. A map of schemes $\pi: X \rightarrow Y$ is affine if for every open $U \subseteq Y, \pi^{-1}(U)$ is an affine open subscheme of $X$. ::: :::{.definition title="affine-local" .flashcard} A property $P$ is said to be affine-local if it is sufficient to check it on an affine cover. See local on the for more. ::: :::{.definition title="affine space" .flashcard} Classically, affine $n$-space, $\mathbb{A}_{k}^{n}$, over a field $k$ is the vector space of $n$-tuples, $k^{n}$. As a scheme, we take $\mathbb{A}_{k}^{n}=\operatorname{Spec} k\left[x_{1}, \ldots, x_{n}\right]$, whose closed points are identified with the classical points when $k$ is algebraically closed. Over an arbitrary ring, $\mathbb{A}_{A}^{n}=\operatorname{Spec} A\left[x_{1}, \ldots, x_{n}\right]$. We can extend this to an arbitrary base scheme $Y$ by taking the fiber product, $\mathbb{A}_{Y}^{n}=\mathbb{A}_{\mathbb{Z}}^{n} \times \mathbb{Z} Y$. ::: :::{.definition title="ample" .flashcard} Let $X$ be a projective (proper is sufficient) A-scheme and $\mathscr{L}$ a line bundle on $X$. We say $\mathscr{L}$ is very ample if there exist $n+1$ sections with no common zero such that the linear system defines a closed embedding $X \hookrightarrow \mathbb{P}_{A}^{n}$. Note that very ampleness implies base point freeness. An equivalent definition (see Exercise 16.6.A) is that $X \simeq \operatorname{Proj} S \bullet$ where $S \bullet$ is a finitely generated graded ring over $A$ and $\mathscr{L} \simeq \mathscr{O}_{\operatorname{Proj} S_{\bullet}}(1)$. If $\pi: X \rightarrow$ Spec $A$, we may refer to this as $\pi$-very ample. A line bundle $\mathscr{L}$ on $X$ is ample if $\mathscr{L}^{\otimes n}$ is very ample for some $n>0$. This is equivalent to $\mathscr{L}^{\otimes n}$ very ample for all $n \gg 0$. ::: :::{.definition title="base point" .flashcard} Let $X$ be a $k$-scheme and $\mathscr{L}$ a line bundle on $X$. A base point $P \in X$ of $\mathscr{L}$ (or of a linear system $V)$ is defined to be a point on which all elements of $\Gamma(X, \mathscr{L})$ vanish. The base locus of $\mathscr{L}$ is defined to be the scheme-theoretic intersection of the vanishing loci of each element of $\Gamma(X, \mathscr{L})$ We say $\mathscr{L}$ is base point free if it has no base points, and define the base point free locus as the complement of the base locus in $X$. It is useful to note that base point freeness of $\mathscr{L}$ is equivalent to $\mathscr{L}$ being globally generated. The utility of this is that an $n+1$ dimensional subset of $\Gamma(X, \mathscr{L})$ defines a morphism $X-\{$ base locus $\} \rightarrow \mathbb{P}_{k}^{n}$. This map is given by evaluating (a basis of) the sections at $P$ to obtain and $(n+1)$-tuple not all zero. ::: :::{.definition title="Canonical bundle/sheaf" .flashcard} Suppose $X$ is a smooth $k$-variety of dimension $n$. The canonical sheaf $\mathscr{K}_{X}$ is taken to be the top wedge power of $\Omega_{X / k}$, $$ \mathscr{K}_{X}=\wedge^{n} \Omega_{X / k}=\operatorname{det} \Omega_{X / k}, $$ which is an invertible sheaf, i.e. a line bundle, on $X$. (See also exterior algebra, determinant, (co)tangent sheaf.) When $X$ is projective (proper is sufficient), $\mathscr{K}_{X}$ is a dualizing sheaf. This is particularly useful in computing the genus, and other cohomological properties. The canonical sheaf/divisor also plays a central role in Riemann-Roch and Riemann-Hurwitz. ::: :::{.definition title="Cartier divisor" .flashcard} Let $\left(X, \mathscr{O}_{X}\right)$ be a scheme with sheaf of total quotients $\mathscr{K}$. A Cartier divisor on $X$ is a global section of the sheaf $\mathscr{K}^{\times} / \mathscr{O}_{X}^{\times}$. This is equivalent to giving an open cover $X=\cup U_{i}$ and elements $f_{i} \in \Gamma\left(U_{i}, \mathscr{K}^{\times}\right)$such that $\left.f_{i} / f_{j} \in \Gamma\left(U_{i} \cap U_{j}\right), \mathscr{O}_{X}^{\times}\right)$. Another characterization of Cartier divisors is to take an effective divisor to be a closed subscheme $D \hookrightarrow X$ such that the ideal sheaf of $D$ is an invertible sheaf on $X$. The sum of such divisors is seen to be the product of ideal sheaves. A Cartier divisor is principal if it is in the image of the map $\Gamma\left(X, \mathscr{K}^{\times}\right) \rightarrow \Gamma\left(X, \mathscr{K}^{\times} / \mathscr{O}_{X}^{\times}\right)$. We can define an equivalence relation on Cartier divisors by $D \simeq D^{\prime}$ if and only if $D-D^{\prime}$ is a principal divisor. Then we define the Cartier divisor class group $\mathrm{CaCl} X$ to be the Cartier divisors mod this equivalence relation. ::: :::{.definition title="category of open sets" .flashcard} Let $X$ be a topological space. The category of open sets on $X$ is a category whose objects are the open sets of $X$ and whose morphisms are inclusions $V \subseteq U$. ::: :::{.definition title="Čech cohomology" .flashcard} Let $\mathscr{F}$ be a sheaf of abelian groups on a quasicompact, separated scheme $X$ with $X=\cup U_{i}$ a finite open cover. The Čech complex of $\mathscr{F}$ with respect to the cover is $$ 0 \rightarrow \prod_{i} \Gamma\left(U_{i}, \mathscr{F}\right) \rightarrow \prod_{i0$ and $N$, and these are equivalent to $M$ being flat. We can compute the Tor functors via free, or indeed projective module resolutions. ::: :::{.definition title="total quotient ring" .flashcard} Let $A$ be a ring and $S$ the multiplicative subset of elements which are not zero divisors. The total quotient ring of $A$ is the localization $S^{-1} A$. This is the "closest thing to the field of fractions" when $A$ is not an integral domain. Let $X$ be a scheme. For each open $U \subseteq X$, let $S(U)$ denote the multiplicative subset of $\Gamma\left(U, \mathscr{O}_{X}\right)$ consisting of elements which are not zero divisors in $\mathscr{O}_{X, p}$ for all $p \in U$. Then the sheaf associated to the presheaf $U \mapsto S(U)^{-1} \Gamma\left(U, \mathscr{O}_{X}\right)$ is called the sheaf of total quotient rings of $\mathscr{O}_{X}$. This is an analog of the function field for an integral scheme. ::: :::{.definition title="twist" .flashcard} A twist of a variety $X / k$ is another variety $T / k$ such that they become isomorphic upon base extension, i.e. $X \times_{k} K \simeq T \times_{k} K$ for a field extension $K / k$. A prototypical example is a quadratic twist of an elliptic curve $E: y^{2}=f(x)$. The twist is given by $E_{d}: d y^{2}=f(x)$. If $\sqrt{d} \notin k$ then $E \not E_{d}$, but they are always isomorphic over $k(\sqrt{d})($ or $\bar{k})$. ::: :::{.definition title="twisting sheaf" .flashcard} Consider the projective space $\mathbb{P}_{k}^{m}$ (or $\mathbb{P}_{A}^{m}$ ). We may define a sheaf $\mathscr{O}_{\mathbb{P}_{k}^{m}}(n)$ by taking it to be the degree $n$ functions on each of the usual affine patches. The transition functions are then multiplying appropriately by $n$-th powers. These glue into a sheaf called $\mathscr{O}_{\mathbb{P}_{k}^{m}}(n)$. In particular, $\mathscr{O}_{\mathbb{P}_{k}^{m}}(1)$ is sometimes called the twisting sheaf. It can be shown that $\mathscr{O}_{\mathbb{P}_{k}^{m}}(n) \otimes_{\substack{\mathcal{O}_{k}^{m} \\ k}} \mathscr{O}_{\mathbb{P}_{k}^{m}}\left(n^{\prime}\right)=\mathscr{O}_{\mathbb{P}_{k}^{m}}\left(n+n^{\prime}\right)$. When $k$ is a field, all line bundles on $\mathbb{P}_{k}^{m}$ arise in this way, so we have $\mathbb{Z} \simeq \operatorname{Pic} \mathbb{P}_{k}^{m}$. Moreover, for $n \geq 0$, the global sections of $\mathscr{O}_{\mathbb{P}_{k}^{m}}(n)$ correspond to $n$-forms, allowing us to easily compute the dimension $$ \Gamma\left(\mathbb{P}_{k}^{m}, \mathscr{O}_{\mathbb{P}_{k}^{m}}(n)\right)=\left(\begin{array}{c} m+n \\ n \end{array}\right) . $$ If $X \hookrightarrow \mathbb{P}_{k}^{m}$ is a projective variety, we define $\mathscr{O}_{X}(n)=\mathscr{O}_{X} \otimes_{\mathscr{O}_{\mathrm{p}}^{m}} \mathscr{O}_{\mathbb{P}_{k}^{m}}(n)$, the pullback of $\mathscr{O}(n)$ to $X$. This can be further generalized by considering a Cartier divisor $D$ on $X$. We define $\mathscr{O}(D)$ to be the dual of the (invertible) ideal sheaf of $D$. ::: :::{.definition title="universal" .flashcard} The $\delta$-functor $T=\left(T^{i}\right): \mathscr{A} \rightarrow \mathscr{B}$ is universal if for any other $\delta$-functor $T^{\prime}=\left(T^{\prime i}\right)$ and any morphism $f^{0}: T^{0} \rightarrow T^{\prime 0}$, there exist unique $f^{i}: T^{i} \rightarrow T^{\prime i}$ for all $i$ which commute with both sets of $\delta$ maps for all exact sequences. ::: :::{.definition title="unramified" .flashcard} If $\pi: X \rightarrow Y$ is a map of schemes, and $\Omega_{\pi}=\Omega_{X / Y}$ is the sheaf of relative differentials, we say $\pi$ is formally unramified if $\Omega_{\pi}=0$. We say $\pi$ is unramified if it is formally unramified and locally of finite type. The ramification locus is the support of $\Omega_{\pi}$, which is a subset of $X$. The branch locus is the image in $Y$ of the ramification locus. ::: :::{.definition title="Valuation" .flashcard} Let $K$ be a field. A valuation on $K$ with values in a totally ordered abelian group $G$ is a group homomorphism $v: K-0 \rightarrow G$ satisfying $v(x+y) \geq \min (v(x), v(y))$. The valuation ring is the subring of $K$ consisting of elements with nonnegative valuation and zero. We call a ring $R$ a valuation ring if $R$ is the valuation ring for a valuation $v$ on $\Frac R$, and we say a valuation is discrete if $G=\mathbb{Z}$. If $Y$ is a prime divisor of $X$, with generic point $\eta$ then the stalk $\mathscr{O}_{X, \eta}$ is a discrete valuation ring with quotient field $K$, the function field of $X$. We call $v_{Y}$ the valuation of $Y$. ::: :::{.definition title="vanishing scheme" .flashcard} Given a scheme $Y$ and a global section $s \in \Gamma\left(Y, \mathscr{O}_{Y}\right)$, the vanishing scheme $V(s)$ is a closed subscheme. On an affine open $\operatorname{Spec} B \subseteq Y$, it corresponds to $\operatorname{Spec} B /\left(s_{B}\right)$ where $s_{B}=\left.s\right|_{\operatorname{Spec} B}$. Given a set $S$ of global sections, we can take $V(S)$ to be defined by Spec $B /\left(S_{B}\right)$ on an affine open. ::: :::{.definition title="variety" .flashcard} An affine scheme over a field $k$ which is reduced and of finite type is called an affine $k$-variety. A reduced (quasi)projective $k$-scheme is called a (quasi)projective $k$-variety. More generally, a variety over a field $k$ is defined to be a reduced, separated scheme of finite type over $k$. Note that affine schemes and (quasi)projective schemes are always separated, so this additional hypothesis is not necessary. ::: :::{.definition title="vector bundle" .flashcard} A vector bundle is another name for a locally free sheaf. Classically, a vector bundle on $X$ is a topological space with a map to $X$ such that the fibers are vector spaces, "continuously varying" as we move along $X$. The trivial vector bundle is $X \times V$, and a general vector bundle is locally trivial, in that on some open cover it is isomorphic to a trivial bundle. As an nontrivial example, think of the Möbius strip as a rank one vector bundle on the circle - it is not globally trivial because it has a twist, but it is trivial if we remove a point on the circle. ::: :::{.definition title="Veronese" .flashcard} Given $\mathbb{P}^{n}$ and the (very ample) line bundle $\mathscr{O}(d)$, we have an embedding $\mathbb{P}^{n} \hookrightarrow \mathbb{P}^{N}$ given by $|\mathscr{O}(d)|$, which we call a Veronese embedding. In particular, $$ N=h^{0}\left(\mathbb{P}^{n}, \mathscr{O}(d)\right)-1=\left(\begin{array}{c} n+d \\ n \end{array}\right)-1 . $$ In the special case of $n=1$, the image curve is called a rational normal curve. As in that case, writing down a basis for $H^{0}\left(\mathbb{P}^{n}, \mathscr{O}(d)\right)$ can give explicit coordinates for the map. ::: :::{.definition title="Weil divisor" .flashcard} Let $X$ be a scheme satisfying $(*)$. A prime divisor on $X$ is a closed integral subscheme $Y$ of codimension one. A Weil divisor is an element of the free abelian group on prime divisors, Div $X$. That is, a Weil divisor $D$ is a formal integer linear combination of prime divisors, $\sum n_{i} Y_{i}$. Vakil takes a Weil divisor to mean a (formal) $\mathbb{Z}$-linear combination of irreducible closed subsets of codimension 1 on a Noetherian scheme $X$. Note this doesn't use the full strength of condition $\left(^{*}\right)$, in that it doesn't assume integrality, separatedness, or regularity, though he quickly adds back in reducedness and regularity. ::: :::{.definition title="Zariski sheaf" .flashcard} A contravariant functor $F: \mathcal{C} \rightarrow$ Sch is a Zariski sheaf if for any scheme $Y$, the assignment $U \mapsto F(U)$ forms a sheaf on $Y$. ::: :::{.definition title="zeros" .flashcard} Let $X$ be a scheme with function field $K$ and let $Y$ be a prime divisor. We say $f \in K^{\times}$has a zero at $Y$ if $v_{Y}(f)>0$, where $v_{Y}$ is the valuation at $Y$. ::: ## Results :::{.proposition title="Adjunction formula" .flashcard} Let $X$ be a smooth variety over $k$ and $Z$ a smooth (closed) subvariety. Then we can compute the canonical bundle $\mathscr{K}_{Z}$ by $$ \left.\mathscr{K}_{Z} \simeq \mathscr{K}_{X}\right|_{Z} \otimes \operatorname{det} \mathscr{N}_{Z / X} . $$ Furthermore, if $Z$ has codimension 1 (i.e. a divisor) then we have $\left.\mathscr{N}_{Z / X} \simeq \mathscr{O}_{X}(Z)\right|_{Z}$, so this is sometimes written $$ \left.\mathscr{K}_{Z} \simeq\left(\mathscr{K}_{X} \otimes \mathscr{O}_{X}(Z)\right)\right|_{Z} . $$ In the notation of divisors, we have $$ K_{Z}=\left.\left(K_{X}+Z\right)\right|_{Z} $$ Proof sketch. Use the conormal exact sequence first, which is exact on the left by smoothness. $$ 0 \rightarrow \mathscr{N}_{Z / X}^{\vee} \rightarrow i^{*} \Omega_{X / k} \rightarrow \Omega_{Z / k} \rightarrow 0 . $$ This induces an alternating product of determinants $$ \operatorname{det} \mathscr{N}_{Z / X}^{\vee} \otimes\left(\operatorname{det} i^{*} \Omega_{X / k}\right) \otimes \operatorname{det} \Omega_{Z / k} \simeq \mathscr{O}_{Z} . $$ Dualizing appropriately and using the definition of $\mathscr{K}$, we have $$ \left.\mathscr{K}_{Z} \simeq \mathscr{K}_{X}\right|_{Z} \otimes \operatorname{det} \mathscr{N}_{Z / X} . $$ ::: :::{.proposition title="Affine communication lemma" .flashcard} Let $P$ be a property of affine open subsets of a scheme $X$. Suppose (i) every affine open $\operatorname{Spec} A \subseteq X$ that has $P$ implies Spec $A_{f} \subseteq X$ also has $P$ and (ii) if $A=\left(f_{1}, \ldots, f_{n}\right)$ and $\operatorname{Spec} A_{f_{i}} \subseteq X$ has $P$ for all $i$ then $\operatorname{Spec} A \subseteq X$ also has $P$. Then if $X=\cup \operatorname{Spec} A_{i}$ where Spec $A_{i}$ has $P$ for all $i$, then every affine open Spec $A \subseteq X$ has $P$ as well. The proof involves intersecting $\operatorname{Spec} A$ with $\operatorname{Spec} A_{i}$ and covering by affine opens simultaneously distinguished in both. Morally, properties $P$ satisfying (i) and (ii) above need only be checked on an open cover. Here are some examples of affine local conditions: - Noetherian-ness (c.f. Proposition 5.3.3), - finite type-ness (c.f. Proposition 5.3.3), - reducedness (and other stalk-local conditions) A non example of such a property is integrality. For example, take $\operatorname{Spec}(A \times B)$ where $A$ and $B$ are arbitrary integral domains. This is a disjoint union (not irreducible) but can be covered by the integral schemes $\operatorname{Spec} A$ and $\operatorname{Spec} B$. ::: :::{.proposition title="Affine reduction iff affine" .flashcard} Let $X$ be a scheme. Then $X$ is affine if and only if the reduced subscheme $X_{\text {red }}$ is affine. Proof. If $X=\operatorname{Spec} A$, then the reduction $X_{\text {red }}=\operatorname{Spec} A / N$, where $N$ is the ideal of nilpotents. There is surprising content in the converse. Suppose $X_{\text {red }}$ is affine. Recall we have a closed embedding $i: X_{\text {red }} \hookrightarrow X$, with the ideal sheaf of $X_{\text {red }}$ given by $\mathscr{N}$, the ideal of nilpotents. Let $\mathscr{F}$ be a coherent sheaf of ideals on $X$. We have a filtration Notice that for any $i \geq 0$ we have $$ \mathscr{F} \supset \mathscr{F} \mathcal{N} \supset \mathscr{F} \mathcal{N}^{2} \supset \cdots . $$ $$ 0 \rightarrow \mathscr{F} \mathcal{N}^{i+1} \hookrightarrow \mathscr{F} \mathcal{N}^{i} \rightarrow \mathscr{F} \mathcal{N}^{i} / \mathscr{F} \mathcal{N}^{i+1} \rightarrow 0, $$ where the rightmost sheaf has the natural structure of an $\mathscr{O}_{X} / \mathcal{N} \simeq \mathscr{O}_{X_{\text {red }}}$-module. Since $H^{1}\left(X_{\text {red }}, \mathscr{F} \mathcal{N}^{i} / \mathscr{F} \mathcal{N}^{i+1}\right)=0$ by Serre's cohomological criterion for affineness, it is enough to prove that $H^{1}\left(X, \mathscr{F} \mathcal{N}^{i}\right)=0$ for some $i$, as then we have $H^{1}\left(X, \mathscr{F} \mathcal{N}^{i-1}\right)=0$, since it's sandwiched between two vanishing $H^{1}$ 's. Going up the filtration, we have $H^{1}(X, \mathscr{F})=0$, and again by Serre's criterion we have $X$ is affine. To show that $\mathscr{F} \mathcal{N}^{i}$ has vanishing $H^{1}$ for some $i$, we argue instead that $\mathcal{N}^{i}=0$ for $i \gg 0$. Covering $X$ by finitely many affine opens $\operatorname{Spec} A$, we see that $\left.\mathcal{N}\right|_{\text {Spec } A} \simeq N$, where $N$ is the nilradical of $A$. Let's further assume $A$ is Noetherian. In the local ring $A_{\mathfrak{p}}$, where we know $\cap_{i=1}^{\infty} \mathfrak{p}^{i}=0$, we have $\cap_{i=1}^{\infty} \mathfrak{p}^{i}=0$ and the proof of this fact shows that $\mathfrak{p}^{i}=0$ for some $i \gg 0$. Since Spec $A$ has finitely many irreducible components (Noetherian), this means that we can take $n$ to be the maximum such $i$ for finitely many generic points. Then for $f \in N^{n}$, we have $f^{n}$ vanishes in all local rings, so it must be that $f=0$, and hence $N^{n}=0$. Doing this for each of the affine opens covering $X$, we have $\mathcal{N}^{i}=0$ for some $i \gg 0$, and as desired, $\mathscr{F} \mathcal{N}^{i}=0$ as well, allowing us to use the previous argument. ::: :::{.proposition title="Bezout's theorem" .flashcard} Let $X \in \mathbb{P}_{k}^{n}$ be a projective scheme of degree $\operatorname{deg} X$ and $H \subset \mathbb{P}_{k}^{n}$ a hypersurface of degree $d$ not containing a component of $X$. The (scheme theoretic) intersection $H \cap X$ has degree $$ \operatorname{deg}(H \cap X)=d(\operatorname{deg} X) . $$ In the special case of curves in $\mathbb{P}^{2}$ (here curves are hypersurfaces) we have the classical result that the number of intersection points of curves $C_{1}$ and $C_{2}$ which don't overlap on an irreducible component is $\left(\operatorname{deg} C_{1}\right)\left(\operatorname{deg} C_{2}\right)$. One way to prove this is with Hilbert polynomials. The key insight is that the ideal sheaf of $H \cap X$ in $X$ is $\mathscr{O}_{X}(-d)$. Another is to intersect $X$ with $\operatorname{dim} X$ many general hyperplanes and count intersection points to define degree, then use Bertini's theorem to make sense of this. ::: :::{.proposition title="Closed embedding exact sequence of sheaves" .flashcard} Let $i: Z \rightarrow X$ be a closed embedding of schemes with ideal sheaf $\mathscr{I}$. We have an exact sequence of $\mathscr{O}_{X}$-modules $$ 0 \rightarrow \mathscr{I} \rightarrow \mathscr{O}_{X} \rightarrow i_{*} \mathscr{O}_{Z} \rightarrow 0 . $$ In the special case where $X$ is a Noetherian scheme which is regular in codimension one (and possibly normal too) and $Z$ is closed of codimension one - i.e. a divisor $D$ - then we have $\mathscr{I}=\mathscr{O}(-D)=\mathscr{O}(D)^{\vee}$, giving us the exact sequence $$ 0 \rightarrow \mathscr{O}_{X}(-D) \rightarrow \mathscr{O}_{X} \rightarrow \mathscr{O}_{D} \rightarrow 0 . $$ Tensoring this exact sequence is often useful. ::: :::{.proposition title="Conormal exact sequence" .flashcard} (affine version) Let $B \rightarrow A \rightarrow A / I$ be maps of rings and $\Omega_{A / B}$ the module of differentials. We can extend the cotangent exact sequence to the left (recognizing that $\Omega_{(A / I) / A}=0$ by $$ I / I^{2} \rightarrow \Omega_{A / B} \otimes_{B} A / I \rightarrow \Omega_{(A / I) / B} \rightarrow 0, $$ where the leftmost map sends $i+I^{2} \mapsto d i \otimes 1$, and extending $(A / I)$-linearly. (global version) If $i: Z \hookrightarrow X$ is a closed embedding and $\pi: X \rightarrow Y$, then we have an exact sequence of sheaves on $Z$, $$ \mathscr{N}_{Z / X}^{\vee} \rightarrow i_{*} \Omega_{Z / Y} \rightarrow \Omega_{Z / X} \rightarrow 0 . $$ Note the discrepancy with the subscripts of the cotangent exact sequence. ::: :::{.proposition title="Cotangent exact sequence" .flashcard} (affine version) Let $C \rightarrow B \rightarrow A$ be maps of rings and $\Omega_{A / B}$ the module of $A$-differentials. We have a right-exact sequence $$ A \otimes_{B} \Omega_{B / C} \rightarrow \Omega_{A / C} \rightarrow \Omega_{A / B} \rightarrow 0 . $$ (global version) Let $X \rightarrow Y \rightarrow Z$ be morphisms of schemes and $\Omega_{X / Y}$, etc. the module of relative differentials, i.e. the co(co)tangent sheaf. We have a right-exact sequence $$ \pi^{*} \Omega_{Y / Z} \rightarrow \Omega_{X / Z} \rightarrow \Omega_{X / Y} \rightarrow 0 $$ where $\pi$ denotes the map $X \rightarrow Y$. ::: :::{.proposition title="Criteria for basepoint free and very ample on curves" .flashcard} Let $X$ be a projective regular integral curve over an algebraically closed field $k=\bar{k}$. We record three observations: (i) Twisting by a point drops the dimension by $\leq 1$ : Let $\mathscr{L}$ be a line bundle on $X$ and $p$ a closed point. Then $$ h^{0}(X, \mathscr{L})-h^{0}(X, \mathscr{L}(-p)) \leq 1 . $$ (ii) Criteria for basepoint freeness: A line bundle $\mathscr{L}$ on $X$ is basepoint free if and only if for all closed points $p \in X$, the inequality in (i) is an equality, $$ h^{0}(X, \mathscr{L})-h^{0}(X, \mathscr{L}(-p))=1 . $$ (iii) Criteria for very ampleness: A line bundle $\mathscr{L}$ on $X$ is very ample if and only if for all closed points $p, q \in X$, not necessarily distinct, the dimension drops maximally, $$ h^{0}(X, \mathscr{L})-h^{0}(X, \mathscr{L}(-p-q))=2 . $$ Proof sketch. For (i), we use the closed subscheme exact sequence $$ \left.0 \rightarrow \mathscr{O}_{X}(-p) \rightarrow \mathscr{O}_{X} \rightarrow \mathscr{O}_{X}\right|_{p} \rightarrow 0, $$ suitably twisted by $\mathscr{L}$ $$ \left.0 \rightarrow \mathscr{L}(-p) \rightarrow \mathscr{L} \rightarrow \mathscr{L}\right|_{p} \rightarrow 0 . $$ This gives the LES in cohomology, $$ 0 \rightarrow H^{0}(X, L(-p)) \rightarrow H^{0}(X, \mathscr{L}) \rightarrow H^{0}\left(p,\left.\mathscr{L}\right|_{p}\right) \rightarrow \cdots $$ Since $p$ is a degree one point and $\mathscr{L}$ is locally free, we have $\left.\mathscr{L}\right|_{p} \simeq \mathscr{O}_{p}$, so $H^{0}\left(p,\left.\mathscr{L}\right|_{p}\right)=k$. Therefore the map $H^{0}(X, \mathscr{L}) \rightarrow H^{0}\left(p,\left.\mathscr{L}\right|_{p}\right)$ is either surjective, in which case the dimension drops by 1 , or it's the zero map, in which case the dimensions are equal. For (ii) we interpret $H^{0}(X, \mathscr{L}(-p))$ as the sections of $H^{0}(X, \mathscr{L})$ which vanish at $p$. If the dimension always drops by one, then we have a section not vanishing at $p$. If this occurs for all $p$, then $\mathscr{L}$ is basepoint free. Conversely, basepoint freeness implies the existence of such a section for each $p$, so the dimension must drop by one. For (iii), we will show very ampleness by arguing that the map associated to the complete linear series $|\mathscr{L}|$ is injective on both points and tangent vectors (see Theorem 19.1.1). For injectivity on points, consider distinct points $p, q$. The dimension of sections drops as we go from $\mathscr{L}(-q)$ to $\mathscr{L}(-p-q)$, indicating the existence of a section vanishing on $q$, but not on $p$. Hence there is a hyperplane in $\mathbb{P}^{\operatorname{dim}|\mathscr{L}|-1}$ containing (the image of) $q$ but not $p$. In particular, the map is injective on points. To see injectivity on tangent vectors consider a closed point $p$. Dualizing, it suffices to show that the map on cotangent vectors is surjective. Since $\Omega_{X / k}$ is one dimensional, we need only see that the map is nonzero. Recall the (Zariski) cotangent space is $\mathfrak{m}_{p} / \mathfrak{m}_{p}^{2}$, so we need to find a section vanishing at $p$ to degree precisely one. However, this exists because $$ h^{0}(X, \mathscr{L}(-p))-h^{0}(X, \mathscr{L}(-2 p))=1 . $$ Thus such an $\mathscr{L}$ is very ample. Conversely, very ample $\mathscr{L}$ implies (i) and (ii), and because it separates points and tangent vectors, we find (iii) holds. ::: :::{.proposition title="Dimensions of $\Gamma$ (and $H^{i}$ ) for $\mathscr{O}_{\mathbb{P}^{n}}$" .flashcard} Let $\mathbb{P}_{k}^{n}$ be projective $A$-space and $\mathscr{O}_{\mathbb{P}_{k}^{n}}(m)=\mathscr{O}(m)$. Then we can compute the dimensions of the (Čech) cohomology groups $H^{i}\left(\mathbb{P}_{k}^{n}, \mathscr{O}(m)\right)$ as follows: $$ \begin{array}{lr} \operatorname{dim}_{k} H^{0}\left(\mathbb{P}_{k}^{n}, \mathscr{O}(m)\right)=\left(\begin{array}{c} m+n \\ m \end{array}\right)=\left(\begin{array}{c} m+n \\ n \end{array}\right), & 0n . \end{array} $$ Note that the above also work over a ring $A$ in place of $k$, where instead of dimension as a $k$-vector space, we take the rank of $H^{\bullet}$ as a free $A$-module. The key idea here is to interpret sections of $\mathscr{O}(m)$ as homogeneous degree $m$ polynomials. See e.g. the discussion in $\S 14.1$ of Vakil. This immediately allows us to compute the dimension of $H^{0}$ by counting degree $m$ monomials in $n+1$ variables. If $U_{i}=D\left(x_{i}\right)$ is the usual affine patch, then $\Gamma\left(U_{i}, \mathscr{O}(m)\right)$ may be identified with the degree $m$ piece of $k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{i}}\right]$ (which in turn has a natural identification with $\left.k\left[x_{0 / i}, \ldots, \widehat{x_{i / i}}, \ldots, x_{n / i}\right]\right)$. Moreover, if $I \subset\{0, \ldots, n\}$ the restriction maps $\Gamma\left(U_{I}, \mathscr{O}(m)\right) \rightarrow \Gamma\left(U_{J}, \mathscr{O}(m)\right)$ with $I \subset J$ are interpreted as natural inclusions of polynomials where $x_{i}$ appears with nonnegative exponent for all $i \notin I$. Now we have the Čech complex of $\oplus \mathscr{O}(m)$, where taking the degree $m$ piece gives the Čech complex of $\mathscr{O}(m)$. \[ &0 \rightarrow k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{0}}\right] \times \cdots \times k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{n}}\right] \rightarrow k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{0}}, \frac{1}{x_{1}}\right] \times \cdots k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{n-1}}, \frac{1}{x_{n}}\right] \rightarrow \cdots \\ &\ldots \rightarrow k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{0}}, \ldots, \frac{1}{x_{n-1}}\right] \times \cdots \times k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{1}}, \ldots, \frac{1}{x_{n}}\right] \rightarrow k\left[x_{0}, \ldots, x_{n}, \frac{1}{x_{0}}, \ldots, \frac{1}{x_{n}}\right] \rightarrow 0 . \] We can more neatly abbreviate this $$ 0 \rightarrow \prod A_{i} \rightarrow \prod A_{i j} \rightarrow \cdots \rightarrow \prod_{\# I=n} A_{I} \rightarrow A_{0 \ldots n} \rightarrow 0 $$ and extend it to $$ 0 \rightarrow A \rightarrow \prod A_{i} \rightarrow \prod A_{i j} \rightarrow \cdots \rightarrow \prod_{\# I=n} A_{I} \rightarrow A_{0 \ldots n} \rightarrow 0 $$ Note that in the above sequence, exactness at the first two places is equivalent to the computation of $H^{0}$. At each place thereafter, taking kernel mod image computes the desired cohomology groups. Importantly, these inclusions preserve both degree and multidegree (i.e. degree of each factor $x_{i}$ ), allowing us to compute the cohomology monomial by monomial, following the strategy outlined for $\mathbb{P}^{2}$ in $\S 18.3$ of Vakil. We note here that this also shows the final assertion, that $H^{j}=0$ for $j>n$ (in other words, by affine cover vanishing, since $\mathbb{P}^{n}$ is covered by $n+1$ affine opens). (all negative) Let's consider monomials $x_{0}^{a_{0}} \cdots x_{n}^{a_{n}}$ where all $a_{i}<0$. The exact sequence above becomes $$ 0 \rightarrow 0_{H^{0}} \rightarrow \prod 0_{i} \rightarrow \cdots \prod_{\# I=n} 0_{I} \rightarrow A_{0 \ldots n} \rightarrow 0 . $$ It's easy to see that the sequence is exact everywhere except the final, $n$-th place. (one nonnegative) Consider now the case that all $a_{i}<0$ except for $a_{n}$. The sequence becomes $$ 0 \rightarrow 0_{H^{0}} \rightarrow \prod 0_{i} \rightarrow \cdots \rightarrow \prod_{\# I=n-1} 0_{I} \rightarrow A_{0 \ldots(n-1)} \times \prod_{\# I=n, n \in I} 0_{I} \rightarrow A_{0 \ldots n} \rightarrow 0 . $$ Since the second to rightmost map is the inclusion $A_{0 \ldots(n-1)} \rightarrow A_{0 \ldots n}$, we have exactness in the $(n-1)$-th place. This is seen to surject for monomials of the stated form, giving exactness of the entire sequence. Hence there is no cohomology here! (at least one negative) Suppose $a_{0}<0$. The strategy above essentially works to show that the sequence is exact once again, though we have to take some care with the signs of the maps. For concreteness, consider a monomial where $a_{0}, \ldots, a_{k}<0$ and $a_{k+1}, \ldots, a_{n} \geq 0$. It's straightforward to see that the leftmost map of $$ A_{0 \ldots k} \rightarrow \prod_{k1$. The multiplication is given by multiplication in $A$, and this is clearly finitely generated as a $B$-module in degree 1. To check $\operatorname{Spec} A \simeq \operatorname{Proj}_{B} S_{\bullet}$, we look at the affine opens $\operatorname{Spec}\left(\left(S_{\bullet}\right)_{f}\right)_{0}$ for $f \in A$ considered as degree $n$; in fact we can assume $n=1$. The ring $\left(\left(S_{\bullet}\right)_{f}\right)_{0}$ is generated by elements of the form $g / f^{k}$ where $g \in A$ has degree $k$. We see that two such elements being equal is the same as them being equal in $A_{f}$, so $\operatorname{Proj} S_{\bullet}$ is covered by affine opens that are isomorphic to distinguished opens in Spec $A$, hence they are isomorphic. We have seen projective implies proper, so we have finite morphisms are proper. For finite is integral and finite type, we first observe that finite trivially implies finite type. Integrality follows from seeing that if $B \rightarrow A$ is a finite map of algebras, it is an integral one. The converse comes from using finite type to give algebra generators $a_{1}, \ldots, a_{n}$, and recognize that integrality implies a (finite) power basis of the $a_{i}$ 's generates $A$ as a $B$-module. For finite implies quasifinite, consider a point $q \in \operatorname{Spec} B \subseteq Y$, i.e. a map Spec $\kappa(q) \rightarrow Y$ coming from $B \rightarrow B_{\mathfrak{q}} / \mathfrak{q}$. Since finite morphisms are affine, the preimage of Spec $B$ is Spec $A \subseteq X$ and we have that the fiber is precisely $$ \pi^{-1}(q)=\operatorname{Spec} A \otimes_{B} B_{\mathfrak{q}} / \mathfrak{q} . $$ Now since $A \otimes_{B} B_{\mathfrak{q}} / \mathfrak{q}$ is finitely generated as a $\kappa(q)$ vector space, we have reduced to the case $B=k$ a field and $A$ a finite $k$-algebra. Since $A$ is Noetherian (rings are Noetherian and finitely generated modules are Noetherian) Spec $A$ has finitely many irreducible components, i.e. it's covered by $\operatorname{Spec} A_{i}$ for integral domains $A_{i}$. If $A$ is an integral domain and a finite $k$-algebra, it is a field because if $f \neq 0$ then the multiplication by $f$ map is an injection $A \rightarrow A$ of $k$-modules, hence surjective by linear algebra, giving an inverse. Thus $A$ is a field and has only one prime. ::: :::{.proposition title="Grothendieck's vanishing theorem" .flashcard} Let $X$ be a scheme of dimension $n$. Then $$ H^{i}(X, \mathscr{F})=0 $$ for all $i>n$ and quasicoherent $\mathscr{F}$ on $X$. Proof idea (projective case). Use affine cover cohomology vanishing. Find $n+1$ hypersurfaces in $\mathbb{P}^{n}$ whose intersections miss $X$, thus $X$ is contained in the union of $n+1 D(f)$ 's. Closed embeddings are affine, so the preimages of these are affine and cover $X$. Thus the cohomology vanishes for all $i \geq n+1$ ::: :::{.proposition title="Integral = reduced + irreducible" .flashcard} $X$ is integral if and only if $X$ is reduced and irreducible. $(\Longrightarrow)$ Suppose $X$ is integral. Then $\mathscr{O}_{X}(U)$ is an integral domain for all $U$. In particular, all rings of sections are reduced, so $X$ is reduced. Suppose $X$ is reducible. Then $X=Z_{1} \cup Z_{1}$ for nonempty sets $Z_{i}$. Assume $p_{i} \in Z_{i}-Z_{j}$, and $p_{i} \in \operatorname{Spec} A_{i}$ for affine opens in $X$. Moreover, we can assume $\operatorname{Spec} A_{i} \cap Z_{j}=\emptyset$. But then $\operatorname{Spec} A_{i}$ are disjoint, so $\mathscr{O}_{X}\left(\operatorname{Spec} A_{1} \cup \operatorname{Spec} A_{2}\right)=$ $A_{1} \times A_{2}$, which is only an integral domain if one of them is the zero ring (it isn't because of nonemptyness!). $(\Longleftarrow)$ Suppose $X$ is reduced and irreducible. Let's do the affine case first. Suppose $f, g \in A$ such that $f g=0$. Then $V(f g)=V(f) \cup V(g)=\operatorname{Spec} A$. By irreducibility, one of $V(f)$ or $V(g)$ is zero, so assume it's $f$. Then $f$ is nilpotent, but by reducedness, we have $f=0$, therefore $A$ is an integral domain. For the general case, we'll need to cover $X$ by affine opens $\operatorname{Spec} A_{i}$. If $f, g \in \Gamma\left(X, \mathscr{O}_{X}\right)$ with restrictions $f_{i}, g_{i}$, suppose $f g=0$, which implies $f_{i} g_{i}=0$. By the previous paragraph (and the observation that a dense open set of an irreducible scheme is irreducible) we have $f_{i} g_{i}=0 \Longrightarrow f_{i}=0$ or $g_{i}=0$. If $f_{i}, g_{j} \neq 0$ for some $i, j$, then we have $V(f) \cup V(g)=X$ is a reduction. Conclude $f$ (or $g)$ must be zero, hence $\mathscr{O}_{X}(X)$ is an integral domain, and hence $\mathscr{O}_{X}(U)$ is for all $U$. ::: :::{.proposition title="Lying over/going up/down" .flashcard} This flavor of result relates primes on either side of ring maps. These are often useful in proving things about dimension. - (lying over) Suppose $\phi: B \rightarrow A$ is an integral extension of rings. Then for all primes $\mathfrak{q} \subset B$, there exists a prime $\mathfrak{p} \subset A$ such that $\mathfrak{p} \cap B=\mathfrak{q}$, i.e. $\mathfrak{p}$ "lies over" $\mathfrak{q}$. Geometrically, this is saying that $\operatorname{Spec} A \rightarrow \operatorname{Spec} B$ is surjective, because the image of $\mathfrak{p} \in \operatorname{Spec} A$ is precisely its preimage under $\phi$, which is intersection with $B$ since the map is an inclusion of rings. - (going up) Suppose $\phi: B \rightarrow A$ is an integral map of rings (not necessarily an extension). Then if $\mathfrak{q}_{1} \subset \mathfrak{q}_{2} \subset \cdots \subset \mathfrak{q}_{n}$ is a chain in $B$ and $\mathfrak{p}_{1} \subset \mathfrak{p}_{2} \subset \cdots \subset \mathfrak{p}_{m}$ is a chain in $A$ with $\mathfrak{p}_{i}$ lying over $\mathfrak{q}_{i}$, the chain upstairs can be extended to $\mathfrak{p}_{n}$ lying over $\mathfrak{q}_{n}$. We are "going up" in the sense that we are extending the chain of primes in $A$ lying over those in $B$ in the upward direction. - (going down) Suppose $\phi: B \hookrightarrow A$ is a finite extension of integral domains with $B$ integrally closed. Given a chain $\mathfrak{q} \subset \mathfrak{q}^{\prime}$ in $B$ and a prime $\mathfrak{p}^{\prime}$ lying over $\mathfrak{q}^{\prime}$, there exists $\mathfrak{p} \subset \mathfrak{p}^{\prime}$ such that $\mathfrak{p}$ lies over $\mathfrak{q}$. Here we are "going down" by going the other way along the chain of inclusions. ::: :::{.proposition title="Map of proper curves is constant or surjective" .flashcard} Let $C, C^{\prime}$ be proper irreducible curves over $k$. Then any map $C \rightarrow C^{\prime}$ over $k$ is surjective or a constant map. Proof idea. By "property $\mathrm{P}$ " arguments, such a map is proper, since $C^{\prime}$ is separated and closed embeddings are proper. The image of $C$ must be an irreducible closed subset, hence either a closed point or all of $C^{\prime}$. ::: :::{.proposition title="Maps to $\mathbb{P}^{n}$ using line bundles" .flashcard} Let $X$ be a $k$-scheme and $V \subseteq \Gamma(X, \mathscr{L})$ where $\mathscr{L}$ is a line bundle on $X$. We could just as well take $V \subseteq \Gamma\left(X, \mathscr{O}_{X}(D)\right)$ where $D$ is a Weil divisor, if they are well defined. Let $\operatorname{dim} V=n+1$. In general, we have $$ X-B \rightarrow \mathbb{P} V, $$ where $\mathbb{P} V$ is the coordinate-free projectivization of $V$. We'll prove this with coordinates. Let $s_{0}, \ldots, s_{n}$ be a basis of sections and take $B$ to be the base locus, i.e. the scheme theoretic intersection of $V\left(s_{i}\right)$. (If $V$ is base point free we may ignore this altogether.) We define $$ X-B \rightarrow \mathbb{P}^{n}, \quad P \mapsto\left[s_{0}(P): \cdots: s_{n}(P)\right] $$ informally. More precisely, consider the open subscheme $D\left(s_{i}\right) \subseteq X$ where $s_{i}$ doesn't vanish. Then we have on any trivializing open subset $U$ of $X, D\left(s_{i}\right) \rightarrow D\left(x_{i}\right)$, where $D\left(x_{i}\right)$ is the standard affine open patch of $\mathbb{P}^{n}$. There is a natural map $k\left[x_{0}, \ldots, x_{n}\right]_{x_{i}} \rightarrow \Gamma\left(D\left(s_{i}\right), \mathscr{O}_{U}\right)$ sending $x_{j} \mapsto s_{j}$, and thus we glue to obtain $U \rightarrow \mathbb{P}^{n}$. We can glue along the trivializing open sets as well to get $X-B \rightarrow \mathbb{P}^{n}$. Some other useful comments: - If $\mathscr{L}$ (or $D$ ) is very ample, then $|\mathscr{L}|$ (or $|D|$ ) defines an embedding of $X$ into $\mathbb{P}^{n}$ where the image is a degree $\operatorname{deg} \mathscr{L}$ (or $\operatorname{deg} D)$ variety. In this case, $\mathscr{L}$ is the pullback of $\mathscr{O}(1)$ along the embedding. - If $X$ is a curve and $\operatorname{dim} V=2$ then the induced map $X \rightarrow \mathbb{P}^{1}$ has degree $\operatorname{deg} \mathscr{L}$ (or $\operatorname{deg} D)$ - In fact, all maps to $\mathbb{P}^{n}$ arise in this way, up to isomorphism (of the bundle and the sections). To see this, given $X \rightarrow \mathbb{P}^{n}$, we pull back the $n+1$ hyperplane sections to $s_{i} \in \Gamma\left(X, \pi^{*} \mathscr{O}(1)\right)$ and check that the above construction gives back the map to $\mathbb{P}^{n}$. ::: :::{.proposition title="Nakayama's Lemma" .flashcard} There are several related results that go by the name "Nakayama's lemma." Here are some of them. Unless otherwise noted, $A$ denotes a ring, $I$ an ideal of $A$, and $M$ an $A$-module. (i) Suppose $M$ is finitely generated and $M=I M$. Then there exists $a \equiv 1(\bmod I)$ such that $a M=0$. Equivalently, since $a=1+i$ for $i \in I$, we have $(-i) m=m$ for all $m \in M$. (ii) Suppose $I$ is contained in all maximal ideals (i.e. $I \subseteq \operatorname{Jac} A$ ) and $M$ is finitely generated, such that $M=I M$. Then $M=0$. (iii) Suppose $I$ is contained in all maximal ideals (i.e. $I \subseteq \operatorname{Jac} A$ ) and $M$ is finitely generated, with a submodule $N \subseteq M$. If $N / I N \rightarrow M / I M$ is a surjection, then $M=N$. (iv) Suppose $(A, \mathfrak{m})$ is local, $M$ is finitely generated, and (the images of) $f_{1}, \ldots, f_{n} \in M$ generate $M / \mathfrak{m} M$ as an $A / \mathfrak{m}$-vector space. Then the $f_{i}$ 's generate $M$ as an $A$-module. Proof. To prove (i), we can use the determinant trick. Let $m_{1}, \ldots, m_{n}$ be generators. Then $m_{i}=\sum_{j} a_{i j} m_{j}$ for $a_{i j} \in I$. Thus $\left(a_{i j}\right)$ times the vector $\left(m_{1}, \ldots, m_{n}\right)^{T}$ is precisely $\left(m_{1}, \ldots, m_{n}\right)^{T}$. Equivalently, $I-\left(a_{i j}\right)$ induces the zero map $M \rightarrow M$. Multiplying by the adjoint matrix, we have $\operatorname{det}\left(I-\left(a_{i j}\right)\right)$ induces the zero map as well, but upon inspection we see that the determinant comes out to an element of $A$ which is $1(\bmod I)$. For (ii), we need only recognize that $a \equiv 1(\bmod I)$ implies $a \in A^{\times}$when $I$ is contained in all maximal ideals. Since $a \equiv 1(\bmod \mathfrak{m})$ for all maximal $\mathfrak{m}$, we have $a \notin \mathfrak{m}$. All nonunits are in a maximal ideal, so $a$ is invertible, and hence $M=0$. For (iii) we leverage (ii) as follows. Let $L$ be the cokernel of the inclusion $M \rightarrow N$. Since taking quotients is right exact (tensoring by $A / I$ ) we have an exact sequence $$ N / I N \rightarrow M / I M \rightarrow L / I L \rightarrow 0 $$ The fact that $N / I N \rightarrow M / I M$ is a surjection means that $L / I L=0$, i.e. $L=I L$. Since $L$ is finitely generated (it's the quotient of a f.g. module) and $I$ is contained in all maximal ideals, by (ii) we have $L=0$, so $M=N$. (iv) is just (iii) applied to the case of $N=\left\langle f_{1}, \ldots, f_{n}\right\rangle$ and $I=\mathfrak{m}$, which is trivially contained in all maximal ideals. ::: :::{.proposition title="Pic of projective space" .flashcard} The Picard group of projective $n$-space $\mathbb{P}_{k}^{n}$ is given by $$ \operatorname{Pic} \mathbb{P}_{k}^{n} \simeq \mathrm{Cl} \mathbb{P}_{k}^{n} \simeq \mathbb{Z} $$ Explicitly, this group is generated by the class of a hyperplane, or equivalently the line bundle $\mathscr{O}(1)$ Proof. First note that $\mathbb{P}^{n}$ is covered by copies of $\mathbb{A}^{n}$, and $k\left[x_{1}, \ldots, x_{n}\right]$ is a UFD, so its stalks are as well. Hence we are free to identify the Picard group with the Weil divisor class group. Next we use the excision exact sequence, $$ 0 \rightarrow \mathbb{Z} \rightarrow \operatorname{Div} X \rightarrow \operatorname{Div}(X-Z) \rightarrow 0 $$ where $Z$ is an irreducible codimension one subscheme (i.e. a prime divisor) and the map sends $1 \mapsto Z$. For this problem, we take $X=\mathbb{P}^{n}$ and $Z$ to be a hyperplane isomorphic to $\mathbb{P}^{n-1}$ (the hyperplane at infinity, if you like). Quotienting by principal divisors, we have $$ \mathbb{Z} \rightarrow \mathrm{Cl} \mathbb{P}^{n} \rightarrow \mathrm{Cl} \mathbb{A}^{n} \rightarrow 0 $$ Again, using that $\mathbb{A}^{n} \simeq \operatorname{Spec} k\left[x_{1}, \ldots, x_{n}\right]$ which is a UFD, the class group vanishes, so $Z \rightarrow \mathrm{Cl} \mathbb{P}^{n}$ is a surjection. Thus the class of the hyperplane, $[H]$ generates $\mathrm{Cl} \mathbb{P}^{n}$. It remains to see that $n[H] \neq 0$ for all $n>0$. Suppose $n H=\operatorname{div} t$ for $t \in K\left(\mathbb{P}^{n}\right)^{\times}$and $n \geq 0$. Then since $n H$ has nonnegative degree, $t$ has no poles, and thus is regular. But the only regular functions on $\mathbb{P}^{n}$ are the constants, which have degree zero (when nonzero). Hence $[n H] \neq 0$ for $n>0$, i.e. $\mathrm{Cl} \mathbb{P}^{n} \simeq\langle[H]\rangle \simeq \mathbb{Z}$. It's useful to identify $[H]$ with the line bundle $\mathscr{O}(1)$, and thus $[n H]$ with $\mathscr{O}(n)$. To see this, consider the global section $x_{0}$ of $\mathscr{O}(1)$. Its image in the standard affine patch $U_{i}$ is 1 for $i=0$ and $x_{0 / i}$ for $00$, the image of $x_{0}$ has a zero at the $V\left(x_{0 / i}\right)$ hyperplane, and we have div $x_{0}=H$. In fact, any two hyperplanes are linearly equivalent, a fact we can realize by passing through $\div x_{0}=H_{0}$ : if $H$ is some other hyperplane, it is given by the vanishing of a linear form $s$, and hence div $\frac{s}{x_{0}}=H-H_{0}$ is principal. ::: :::{.proposition title="Pic as cohomology: $Pic(X) \cong H^1(X; \OO_X\units)$" .flashcard} Let $X$ be a quasicompact separated scheme (i.e. a scheme where Čech cohomology makes sense). Then $$ \operatorname{Pic} X \simeq H^{1}\left(X, \mathscr{O}_{X}^{\times}\right) $$ where $\mathscr{O}_{X}^{\times}$is the sheaf of abelian groups given by taking units: $U \mapsto \Gamma\left(U, \mathscr{O}_{X}\right)^{\times}$. Proof idea. Use Čech cohomology on some affine open cover $X=\cup \operatorname{Spec} A_{i}$, where the intersection $\operatorname{Spec} A_{i} \cap \operatorname{Spec} A_{j}=\operatorname{Spec} A_{i j}$ is affine. Giving a line bundle involves giving transition functions, which will correspond to units in the rings $A_{i j}$. Then show these units sit in the kernel of the map in the Čech complex $$ \prod A_{i j}^{\times} \rightarrow \prod A_{i j k} $$ giving a map $$ \mathscr{L} \mapsto H^{1}\left(X, \mathscr{O}_{X}^{\times}\right) $$ Next, show that if these transition functions are coboundaries, i.e. in the image of $\prod A_{i}^{\times}$, then the original line bundle is trivial. Hence the map above is an isomorphism. Note, this uses the fact that we can compute $H^{1}$ via Čech cohomology, which is nontrivial in this case since $\mathscr{O}_{X}^{\times}$is not a quasicoherent sheaf. However, we have now seen that Pic $X$ maps to the Čech $H^{1}$ for any open cover, so it has a map to the limit which must also be an isomorphism. An exercise in Hartshorne (III.4.4) shows that derived functor $H^{1}$ is isomorphic to the limit of Čech $H^{1}$ 's, completing the proof. ::: :::{.proposition title="Properties of cohomology" .flashcard} Let $X$ be a scheme. Given a quasicoherent $\mathscr{O}_{X}$-module $\mathscr{F}$, recall $H^{i}(X, \mathscr{F})$ is the $i$-th sheaf cohomology group, which we may compute via Čech cohomology if $X$ is quasicompact and separated. (i) functoriality: $H^{i}(X, \cdot)$ is a covariant functor from $\mathrm{QCoh}_{X} \rightarrow \mathrm{Ab}$. If $X$ is a scheme over $A$ then it is a functor $\mathrm{QCoh}_{X} \rightarrow \operatorname{Mod}_{A}$. In particular, $H^{i}(X, \mathscr{F})$ is a vector space for a scheme over a field. (ii) $\boldsymbol{H}^{0}$ is global sections: For any sheaf $\mathscr{F}$ on $X$, we have $H^{0}(X, \mathscr{F})=\Gamma(X, \mathscr{F})$. (iii) long exact sequence: Let $0 \rightarrow \mathscr{F}^{\prime} \rightarrow \mathscr{F} \rightarrow \mathscr{F}^{\prime \prime} \rightarrow 0$ be a short exact sequences of sheaves on $X$. There is a long exact sequence in cohomology, $0 \rightarrow H^{0}\left(X, \mathscr{F}^{\prime}\right) \rightarrow H^{0}(X, \mathscr{F}) \rightarrow H^{0}\left(X, \mathscr{F}^{\prime \prime}\right) \rightarrow H^{1}\left(X, \mathscr{F}^{\prime}\right) \rightarrow H^{1}(X, \mathscr{F}) \rightarrow \cdots$ $\cdots \rightarrow H^{n}\left(X, \mathscr{F}^{\prime}\right) \rightarrow H^{n}(X, \mathscr{F}) \rightarrow H^{n}\left(X, \mathscr{F}^{\prime \prime}\right) \rightarrow H^{n+1}\left(X, \mathscr{F}^{\prime}\right) \rightarrow H^{n+1}(X, \mathscr{F}) \rightarrow \cdots$ (iv) contravariant in space: If $\pi: X \rightarrow Y$ is a map of schemes (possibly over a base) then there are natural maps $$ \begin{aligned} H^{i}\left(Y, \pi_{*} \mathscr{F}\right) & \rightarrow H^{i}(X, \mathscr{F}) \\ H^{i}(Y, \mathscr{G}) & \rightarrow H^{i}\left(X, \pi^{*} \mathscr{G}\right), \end{aligned} $$ allowing us to reasonably say that $H^{i}$ is "contravariant in the space." (v) affine morphisms: Suppose $\pi: X \rightarrow Y$ is an affine morphism. Then the natural map of (iv) is an isomorphism $$ H^{i}\left(Y, \pi_{*} \mathscr{F}\right) \stackrel{\sim}{\longrightarrow} H^{i}(X, \mathscr{F}) . $$ (vi) affine cover vanishing: Suppose $X$ is covered by $n$ affine open sets. Then $$ H^{i}(X, \mathscr{F})=0 $$ for all $i \geq n$. In particular, if $X$ is affine then all higher cohomology vanishes for all quasicoherent $\mathscr{F}$. (vii) $\boldsymbol{H}^{i}$ commutes with (filtered) colimits: Let $X$ be an $A$-scheme. Then for a filtered direct system $\mathscr{F}_{j}$, we have ![](https://cdn.mathpix.com/cropped/2022_10_17_cd4f806c438565f8bbceg-35.jpg?height=105&width=514&top_left_y=2192&top_left_x=797) In particular, cohomology commutes with direct sums $$ \oplus_{j} H^{i}\left(X, \mathscr{F}_{j}\right) \simeq H^{i}\left(X, \oplus_{j} \mathscr{F}_{j}\right) . $$ (viii) Dimensions of $\Gamma$ (and $H^{i}$ ) for $\mathscr{O}_{\mathbb{P}^{n}}$ : Consider $\mathbb{P}_{k}^{n}$. Then $$ \begin{aligned} h^{0}\left(\mathbb{P}^{n}, \mathscr{O}(m)\right) &=\left(\begin{array}{c} m+n \\ n \end{array}\right) \\ h^{n}\left(\mathbb{P}^{n}, \mathscr{O}(m)\right) &=\left(\begin{array}{c} -m-1 \\ n \end{array}\right) \\ h^{i}\left(\mathbb{P}^{n}, \mathscr{O}(m)\right) &=0 \text { for all } i \neq 0, n . \end{aligned} $$ (ix) $\boldsymbol{H}^{i}$ finitely generated: Let $X$ be a projective $A$-scheme for a Noetherian ring $A$ and $\mathscr{F}$ a coherent sheaf on $X$. Then $H^{i}(X, \mathscr{F})$ is finitely generated for all $i$. In particular, if $A=k$ then $H^{i}(X, \mathscr{F})$ is a finite dimensional vector space over $k$. (x) Serre vanishing: Let $X$ be a projective $A$-scheme for a Noetherian ring $A$ and $\mathscr{F}$ a coherent sheaf on $X$. Then for $m \gg 0$, we have $H^{i}(X, \mathscr{F}(m))=0$ for all $i>0$. This holds without Noetherian hypotheses as well. (xi) base change: Let $X$ be a quasicompact and separated over a field $k$. Then for any field extension $K / k$, we have $$ H^{i}(X, \mathscr{F}) \otimes_{k} K \simeq H^{i}\left(X \times_{k} \operatorname{Spec} K, \mathscr{F} \otimes_{k} K\right) . $$ (xii) dimensional vanishing, aka. ??: Let $X$ be a projective $k$-scheme. Then $H^{i}(X, \mathscr{F})=$ 0 for all $i>\operatorname{dim} X$ and any quasicoherent $\mathscr{F}$. (xiii) Serre's cohomological criterion for affineness: $X$ is affine if and only if $H^{i}(X, \mathscr{F})=$ 0 for all $i>0$ and all quasicoherent $\mathscr{F}$ on $X$. This is a converse to (vi) when $n=1$. Proof sketch(es). (i) This is obvious from the derived functor setup. If you setup with Čech cohomology, a map of sheaves gives a map on Čech complexes, which admits maps on cohomology. (ii) In the derived functor setup, this is by definition, as $H^{0}(X, \cdot)=\Gamma(X, \cdot)$. For Čech cohomology, this is precisely the sheaf axiom. (iii) Again, if you take the derived functor approach this is implied by the LES for derived functors. In general, if we have an exact sequence of complexes $0 \rightarrow C_{\bullet}^{\prime} \rightarrow C_{\bullet} \rightarrow C_{\bullet}^{\prime \prime} \rightarrow 0$ we can take any two "rows" ![](https://cdn.mathpix.com/cropped/2022_10_17_cd4f806c438565f8bbceg-36.jpg?height=184&width=727&top_left_y=1640&top_left_x=794) and quotient the top row by the image, while taking the kernel of the bottom, because these are complexes. This removes left exactness on top and right exactness on the bottom: ![](https://cdn.mathpix.com/cropped/2022_10_17_cd4f806c438565f8bbceg-36.jpg?height=192&width=1114&top_left_y=1937&top_left_x=598) One observes the kernels of the vertical maps are $H^{i}$ 's while the cokernels are $H^{i+1}$ 's. Applying the snake lemma gives the LES. (iv) Using Čech cohomology, it's straightforward to see there's a map on Čech complexes $C \bullet\left(Y, \pi_{*} \mathscr{F}\right) \rightarrow C \bullet(X, \mathscr{F})$, by choosing covers appropriately and taking restriction maps. For the pullback, if $\mathscr{G}$ is a quasicoherent sheaf on $Y$, there is a natural map $\mathscr{G} \rightarrow \pi_{*} \pi^{*} \mathscr{G}$ by adjointness, so by the above and (i) we have $$ H^{i}(Y, \mathscr{G}) \rightarrow H^{i}\left(Y, \pi_{*} \pi^{*} \mathscr{G}\right) \rightarrow H^{i}\left(X, \pi^{*} \mathscr{G}\right) \text {. } $$ (v) When computing the natural map on Čech cohomology in (iv), observe that if $\cup U_{i}=Y$ is a finite affine cover of $Y$, then $\cup \pi^{-1}\left(U_{i}\right)$ is a finite affine cover of $X$. Thus the Čech complexes are identical. (vi) Computing via Čech cohomology, the complex stops at the $n$-th position, so the $n$-th and higher cohomology vanish if $H^{i}(X, \mathscr{F})$ is independent of the cover we compute with. In order to prove this, we actually need to first show the $n=1$ case, that $H^{i}(X, \mathscr{F})$ vanishes for $i>0$ when $X$ is affine (for any cover). The proof goes as follows. First assume that $\operatorname{Spec} A$ itself is in the cover. Then the Čech complex for this cover (with $\Gamma(X, \mathscr{F})$ appended) sits in the middle of an exact sequence with the complex forced to contain $\operatorname{Spec} A$ on top and the cover with $\operatorname{Spec} A$ removed on the bottom. The top and bottom are identical, but shifted, and the maps on cohomologies between them are isomorphisms, forcing $H^{i}(X, \mathscr{F})=0$ for $i>0$. In general, if $X=\operatorname{Spec} A$ and $\cup U_{i}$ is an affine cover, we choose a distinguished open cover of $X$ such that each distinguished open is contained in a $U_{i}$. This allows us to compute locally on our distinguished cover, which is in the previous case by our choices, so the cohomology vanishes. We can then show that adding an open set to a cover doesn't change the Čech cohomology, so any two affine covers compute the same $H^{i}$ 's. (vii) This is a consequence of the fact that filtered colimits are exact in $\operatorname{Mod}_{A}$ and the FHHF theorem, which states that exact functors commute with cohomology. See Exercises 1.6.H and 1.6.L. (viii) See Dimensions of $\Gamma\left(\right.$ and $H^{i}$ ) for $\mathscr{O}_{\mathbb{P}^{n}}$. (ix) We're free to use (v) to compute $\pi_{*} \mathscr{F}$ on $\mathbb{P}_{A}^{n}$ instead, allowing us to use (viii) above. We'll also need that since $\mathscr{F}$ is coherent, $\mathscr{F}(m)$ is globally generated for some $m \gg 0$ (this has to do with (very) ampleness). What we need is that $$ 0 \rightarrow \mathscr{G} \rightarrow \mathscr{O}(m)^{\oplus j} \rightarrow \mathscr{F} \rightarrow 0 $$ is exact for some $m$ and $j$, and coherence implies that $\mathscr{G}$ is also coherent. By (vi) we have vanishing of $H^{i}\left(\mathbb{P}^{n}, \mathscr{F}\right)$ for $i>n$. The LES of (iii) ends in $$ \cdots \rightarrow H^{n}\left(\mathbb{P}^{n}, \mathscr{G}\right) \rightarrow H^{n}\left(\mathbb{P}^{n}, \mathscr{O}(m)\right)^{\oplus j} \rightarrow H^{n}\left(\mathbb{P}^{n}, \mathscr{F}\right) \rightarrow 0 . $$ Thus $H^{n}\left(\mathbb{P}^{n}, \mathscr{F}\right)$ is a quotient of a finitely generated $A$-module, and hence is finitely generated itself. This holds for $H^{n}\left(\mathbb{P}^{n}, \mathscr{G}\right)$ as well, as we haven't used anything about $\mathscr{F}$ here. Now we see $H^{n-1}\left(\mathbb{P}^{n}, \mathscr{F}\right)$ is sandwiched between finitely generated modules in the LES, so it too must be finitely generated. Inducting downwards, we are done. (x) Repeat the above proof, but twist by $\mathscr{O}(N)$ for $N$ sufficiently large that $H^{n}\left(\mathbb{P}^{n}, \mathscr{O}(m+\right.$ $N))=0$, which is possible by (viii). Then the argument from (ix), combined with the fact that $H^{i}\left(\mathbb{P}^{n}, \mathscr{O}(m+N)\right)=0$ always for $00$ and all quasicoherent sheaves $\mathscr{F}$ on $X$, (iii) $H^{1}(X, \mathscr{I})=0$ for all coherent sheaves of ideals $\mathscr{I}$ on $X$. Proof idea (Noetherian case). (i $\Longrightarrow$ ii) is true by Čech cohomology; see Properties of cohomology. (ii $\Longrightarrow$ iii) is clear. The content is to show (iii $\Longrightarrow$ i). Suppose $X$ satisfies (iii). One shows that there exist sections $f_{1}, \ldots, f_{r} \in \Gamma\left(X, \mathscr{O}_{X}\right)$ for which $D\left(f_{i}\right)$ are all affine and the $f_{i}$ 's generate the unit ideal in $\Gamma\left(X, \mathscr{O}_{X}\right)$. Then $X=\operatorname{Spec} A$. This amounts to checking that the natural map $X \rightarrow \operatorname{Spec} \Gamma\left(X, \mathscr{O}_{X}\right)$ is an affine morphism, hence $X$ is affine. Let $p \in X$ be a closed point (which exist for Noetherian schemes!) and $Y=X-U$ for an affine open $U$ containing $p$. Then we have an exact sequence $$ 0 \rightarrow \mathscr{I}_{Y \cup\{p\}} \rightarrow \mathscr{I}_{Y} \rightarrow \kappa(p) \rightarrow 0 . $$ Interpret this as "the functions vanishing on both $Y$ and $p$ inject into $Y$, the quotient of which is $\kappa(p)$, i.e. the skyscraper sheaf on $p$. Taking the long exact sequence on cohomology, using hypothesis (iii), we see that $$ \Gamma\left(X, \mathscr{I}_{Y}\right) \rightarrow \kappa(p) \rightarrow 0, $$ so there exists a function $f$ on $X$ which doesn't vanish at $p$. Since $D(f) \subseteq U$ and $f$ is also a function on $U$, we have that $D(f)$ is affine, containing $p$. By quasicompactness, we obtain a finite open cover $\cup D\left(f_{i}\right)=X$. It remains to show these generate the unit ideal. Consider $$ 0 \rightarrow \mathscr{F} \rightarrow \mathscr{O}_{X}^{r} \rightarrow \mathscr{O}_{X} \rightarrow 0, $$ where the map sends $1 \mapsto f_{i}$ on the basis of $\mathscr{O}_{X}^{r}$. We need to show that this is surjective on global sections, i.e. $H^{1}(X, \mathscr{F})=0$. While this isn't true a priori, as (iii) applies only to coherent sheaves of ideals, we filter by $$ \mathscr{F} \supseteq F \cap \mathscr{O}_{X}^{r-1} \supseteq \cdots \supseteq \mathscr{F} \cap \mathscr{O}_{X} . $$ The rightmost sheaf in the sequence is indeed a coherent sheaf of ideals, hence it has vanishing cohomology, and each of the successive quotients may be interpreted as a coherent sheaf of ideals, allowing us to "walk up" the filtration (much like how we did to show $X_{\text {red }}$ affine $\Longleftrightarrow X$ affine!). ::: :::{.proposition title="Serre duality" .flashcard} (existence) Let $X$ be a projective scheme over $k$. Then a dualizing sheaf, $\omega$, exists. That is, for all coherent $\mathscr{F}$, we have a pairing $$ \operatorname{Hom}(\mathscr{F}, \omega) \times H^{n}(X, \mathscr{F}) \rightarrow k $$ with $$ \operatorname{Hom}(\mathscr{F}, \omega) \simeq H^{n}(X, \mathscr{F})^{\vee} . $$ When $X$ is Cohen-Macaulay and of equidimension $n$ and $\mathscr{F}$ is locally free, we have $$ H^{i}(X, \mathscr{F}) \simeq H^{n-i}\left(X, \omega \otimes \mathscr{F}^{\vee}\right)^{\vee} . $$ (coincidence with canonical sheaf) When $X$ is smooth, we have that the dualizing sheaf $\omega$ coincides with the canonical bundle/sheaf $\mathscr{K}_{X}=\operatorname{det} \Omega_{X / k}$. In this case, Serre duality refers to the perfect pairing $$ H^{i}(X, \mathscr{F}) \times H^{n-i}\left(X, \mathscr{K}_{X} \otimes \mathscr{F}^{\vee}\right) \rightarrow H^{n}\left(X, \mathscr{K}_{X}\right) \simeq k, $$ which implies $$ h^{i}(X, \mathscr{F})=h^{n-i}\left(X, \mathscr{K}_{X} \otimes \mathscr{F}^{\vee}\right) . $$ (curves) In the special case of $X$ a smooth curve $(n=1)$, we have the oft-used formula $$ h^{1}\left(X, \mathscr{O}_{X}(D)\right)=h^{0}\left(X, \mathscr{O}_{X}(K-D)\right), $$ used in a common formulation of Riemann-Roch. This also allows us to that the arithmetic genus $p_{a}=1-\chi\left(X, \mathscr{O}_{X}\right)$ agrees with the geometric genus $g=h^{0}\left(X, \omega_{X}\right)$, $$ p_{a}=1-\chi\left(X, \mathscr{O}_{X}\right)=1-h^{0}\left(X, \mathscr{O}_{X}\right)+h^{1}\left(X, \mathscr{O}_{X}\right)={ }^{(S D)} h^{0}\left(X, \omega_{X}\right)=g . $$ ::: :::{.proposition title="Sheaves which are not quasicoherent" .flashcard} Let $X=\operatorname{Spec} k[x]_{(x)}$, which has a generic point $\eta$ and a closed point corresponding to $(x)$. The open sets consist of $\emptyset, U=\{\eta\}, X$. Consider the sheaf of abelian groups $\mathscr{F}$ obtained by assigning $$ \Gamma(X, \mathscr{F})=k(x), \quad \Gamma(U, \mathscr{F})=0, $$ where the restriction map is the obvious: the zero map. This is an $\mathscr{O}_{X}$-module, because each open set is a $k[x]_{(x)}$-module (we can make $x$ act trivially on $k(x)$ ) and this agrees with restriction. If this were quasicoherent, we'd have $\Gamma(X, \mathscr{F})_{x}=\Gamma(U, \mathscr{F})$, since $U=D(x)$. But of course, $k(x)_{x}=k(x) \neq 0$. Therefore this isn't quasicoherent! A similar strategy can be used to show the pushforward of the sheaf $k(x)$ on the closed point at the origin (a.k.a. the skyscraper sheaf at the origin) is not quasicoherent on Spec $k[x]$. Here the idea is that $\Gamma\left(\operatorname{Spec} k[x], i_{*} k(x)\right)=k(x)$, and $\Gamma\left(D(x), i_{*} k(x)\right)=0$. Thus the powers of $x$ must annihilate the module $k(x)$ if this were to be quasicoherent; this is clearly seen to be false! ::: :::{.proposition title="Smoothness characterizations" .flashcard} Let $X$ be a (necessarily) finite type $k$-scheme. The following conditions are equivalent for $X$ to be $k$-smooth, of (pure) dimension $n$. (i) $X$ has a cover by affine open sets $\operatorname{Spec} k\left[x_{1}, \ldots, x_{m}\right] /\left(f_{1}, \ldots, f_{r}\right)$ where the Jacobian matrix has corank $n$ at every point (we took this to be the definition of smooth). (ii) The cotangent bundle $\Omega_{X / k}$ is locally free of rank $n$. (iii) If $k$ is a perfect field, $X$ is regular and finite type. Note that for (iii), we always have that smooth schemes are regular. When $k$ is perfect, it turns out that regularity implies smoothness. Consider now a morphism of schemes $\pi: X \rightarrow Y$. The following conditions about smoothness (of some relative dimension $n$ ) are equivalent. (i) $X$ and $Y$ have covers by open sets such that $\pi$ locally looks like the map induced by $B \rightarrow B\left[x_{1}, \ldots, x_{n+r}\right] /\left(f_{1}, \ldots, f_{r}\right)$, where the Jacobian in the first $r$ variables is invertible (we took this to be the definition). (ii) $\pi$ is locally finitely presented, flat of relative dimension $n$, and $\Omega_{\pi}=\Omega_{X / Y}$ is locally free of rank $n$. (iii) $\pi$ is locally finitely presented, flat of relative dimension $n$, and the fibers are smooth $k$-schemes of pure dimension $n$. (iv) $\pi$ is locally finitely presented, flat of relative dimension $n$, and the geometric fibers are smooth $k$-schemes of pure dimension $n$. Recalling that a morphism is étale if it is smooth of relative dimension zero, we find that étaleness is equivalent to flat + loc. fin. pres. + unramified, or simply smooth and unramified. Conditions (iii) and (iv) above imply that the fibers of etale morphisms look like a disjoint union of copies of Spec $K$, where $K / \kappa(p)$ is a finite separable extension (and in fact this is enough to be etale). ![](https://cdn.mathpix.com/cropped/2022_10_17_cd4f806c438565f8bbceg-43.jpg?height=49&width=1390&top_left_y=1203&top_left_x=427) that $\Omega_{\text {Spec } k\left[x_{1}, \ldots, x_{m}\right] /\left(f_{1}, \ldots, f_{r}\right) / k}$ computes the cokernel of the Jacobian matrix (see Exercise 21.2.E). Thus if $X$ satisfies (i), then the stalks of $\Omega_{X / k}$ have rank $n$, and since constant rank implies locally free (for finite type quasicoherent sheaves on a reduced scheme, see Exercise 13.7.K) we have (ii). Conversely, if $\Omega_{X / k}$ is locally free, we have that after covering by affine open sets of the form in (i), the module of differentials - which computes the cokernel of the Jacobian - has the correct rank. For regularity implies smoothness, first we use that $X$ is regular implies $X_{\bar{k}}$ is regular. See Exercise 12.2.O; the idea is that regular local rings are preserved under base extension, provided the residue field is separated over $k$ (hence the perfection hypothesis!). We then recognize that if $X_{\bar{k}}$ is regular at its closed points, it must be $\bar{k}$-smooth, as the points failing to satisfy the Jacobian criterion are in the vanishing set of a certain determinant, hence this set contains closed points. In fact this is an if and only if. Finally, we have that $\Omega_{X / k} \otimes_{k} \bar{k} \simeq \Omega_{X_{\bar{k}} / \bar{k}}$ by pullback of differentials; this preserves rank, so one is locally free if and only if the other is. For smoothness implies regular, we don't need the perfect hypothesis on $k$. Smoothness means we have $X$ is an etale cover of $\mathbb{A}_{k}^{n}$, which is regular. Exercise 25.2.D shows that the preimage of a regular point under an etale map is regular. ::: :::{.proposition title="Useful adjoint pairs" .flashcard} Below are several useful left/right adjoint functor pairs. - "-ify" (left adjoint) and "forget" (right adjoint), e.g. - sets to groups (here "-ify" is the functor producing the free group on a set), - presheaves to sheaves, - $\Gamma\left(X, \mathscr{O}_{X}\right)$-modules to $\mathscr{O}_{X}$-modules (here "-ify" is the $\widetilde{\text { functor), etc. }}$ - Tensor $\cdot \otimes_{A} N$ (left adjoint) and $\operatorname{Hom}_{A}(N, \cdot)$ (right adjoint) as functors from $A$-modules to $A$-modules, for a fixed $A$-module $N$. - Inverse image $\pi^{-1}$ (left adjoint) and pushforward $\pi_{*}$ (right adjoint), as functors to/from sheaves on $X$ to sheaves on $Y$ for a fixed map $\pi: X \rightarrow Y$. - Pullback $\pi^{*}$ (left adjoint) and pushforward $\pi_{*}$ (right adjoint) as functors to/from $\mathscr{O}_{X^{-}}$ modules to $\mathscr{O}_{Y}$-modules, for fixed map $\pi: X \rightarrow Y$. - Extension by zero $i_{!}$(left adjoint) and inverse image $i^{-1}$ (right adjoint), for an open embedding $i: U \hookrightarrow X$. This handily implies that $i^{-1}$ is exact in this case. ::: :::{.proposition title="Valuative criterion for properness" .flashcard} Let $\pi: X \rightarrow Y$ be a quasiseparated finite type map of schemes and $K$ a valued field with valuation ring $A$. Then $\pi$ is proper if and only if for every such $K, A$ with outer diagram below, there exists exactly one diagonal arrow: ![](https://cdn.mathpix.com/cropped/2022_10_17_cd4f806c438565f8bbceg-44.jpg?height=176&width=285&top_left_y=557&top_left_x=974) If $X$ and $Y$ are locally Noetherian, then we need only consider discrete valuation rings $A$ in the diagram above. Note that the existence of a dashed arrow implies universal closedness. Since separatedness is part of the definition of properness, this combined with the uniqueness from the valuative criterion for separatedness gives properness. ::: :::{.proposition title="Valuative criterion for separatedness" .flashcard} Let $\pi: X \rightarrow Y$ be a map of schemes and $K$ a valued field with valuation ring $A$. Then $\pi$ is separated if and only if for every such $K, A$ with outer diagram below, there exists at most one diagonal arrow: ![](https://cdn.mathpix.com/cropped/2022_10_17_cd4f806c438565f8bbceg-44.jpg?height=171&width=285&top_left_y=1102&top_left_x=974) If $\pi$ is finite type and $X$ and $Y$ are locally Noetherian, then we need only consider discrete valuation rings $A$ in the diagram above. :::