Fact-checked by Grok 2 weeks ago

Projective variety

A projective variety is an irreducible closed subset of \mathbb{P}^n over an , defined as the zero set of a collection of homogeneous polynomials, equipped with the induced . \mathbb{P}^n itself consists of lines through the in affine (n+1)-space, with points represented by [x_0 : \cdots : x_n], where not all coordinates vanish. This structure ensures that projective varieties are "compact" in the classical sense when defined over the complex numbers, analogous to compact Riemann surfaces, and they incorporate points at infinity, making them suitable for studying global geometric properties without boundary issues inherent in affine varieties. The coordinate ring of a projective variety X \subseteq \mathbb{P}^n is the homogeneous k[X] = k[x_0, \dots, x_n]/I(X), where I(X) is the homogeneous of polynomials vanishing on X, and by the projective Nullstellensatz, this ideal is the radical of the vanishing ideal for non-empty sets. Projective varieties can be covered by finitely many affine open sets via the standard affine charts U_i = X \cap D(x_i), where D(x_i) is the principal excluding the V(x_i), allowing the use of affine techniques while preserving projective properties. Regular functions on the entire variety are constant, reflecting their rigidity, with the sheaf of regular functions \mathcal{O}_X providing a local description. In , projective varieties form the core objects for investigating phenomena like , (defined as the transcendence degree of the function field), and (where the Zariski tangent space dimension equals the variety's at smooth points). Notable examples include projective curves such as elliptic curves in \mathbb{P}^2, defined by homogeneous cubics, and higher-dimensional objects like hypersurfaces. Their study traces back to foundational work by figures like Dedekind on ideals (1876), , and Noether's developments, evolving into modern tools for and moduli problems. Theorems like Bézout's, stating that two plane curves of degrees m and n without common components intersect in mn points ( multiplicity), underscore their role in counting geometric invariants.

Definitions and Basic Constructions

Projective Space and Homogenization

Projective space \mathbb{P}^n_k over an k is defined as the set of all one-dimensional subspaces (lines through the ) of the k^{n+1}, or equivalently, the quotient space (k^{n+1} \setminus \{0\}) / k^\times, where k^\times acts by . Points in \mathbb{P}^n_k are represented using [x_0 : x_1 : \dots : x_n], where not all x_i = 0, and two tuples (x_0, \dots, x_n) and (y_0, \dots, y_n) represent the same point if there exists \lambda \in k^\times such that x_i = \lambda y_i for all i. This construction identifies \mathbb{P}^n_k with the space of lines in k^{n+1}, providing a geometric interpretation that generalizes the classical . To define projective varieties from affine ones, the process of homogenization extends polynomials from affine space \mathbb{A}^n_k to projective space \mathbb{P}^n_k. For a polynomial f \in k[x_1, \dots, x_n] of degree d, its homogenization F with respect to a new variable X_0 is obtained by writing f = f_d + f_{d-1} + \dots + f_0, where each f_i is the homogeneous component of degree i, and setting F(X_0, x_1, \dots, x_n) = X_0^d f(x_1/X_0, \dots, x_n/X_0) = f_d(x_1, \dots, x_n) + X_0 f_{d-1}(x_1, \dots, x_n) + \dots + X_0^d f_0. For an affine variety V \subset \mathbb{A}^n_k defined by an ideal I(V) = (f_1, \dots, f_m), the projective closure \overline{V} \subset \mathbb{P}^n_k is the zero set of the ideal generated by the homogenizations F_1, \dots, F_m, ensuring \overline{V} \cap \mathbb{A}^n_k = V. For example, the affine curve V(y^2 - x^3) \subset \mathbb{A}^2_k homogenizes to V(y^2 z - x^3) \subset \mathbb{P}^2_k, where z is the homogenizing variable. Homogenization incorporates points at infinity into the affine variety, which are points in \overline{V} where the homogenizing coordinate vanishes (e.g., z = 0 in \mathbb{P}^2_k). These points resolve behaviors at "infinity" in the affine setting, such as asymptotic directions of curves. For instance, the homogenization of the hyperbola V(xy - 1) \subset \mathbb{A}^2_k adds two points at infinity, [1:0:0] and [0:1:0], closing the curve in \mathbb{P}^2_k. Topologically, \mathbb{P}^n_k serves as a compactification of \mathbb{A}^n_k, embedding \mathbb{A}^n_k as the dense where x_0 \neq 0 (via the chart [1 : x_1 : \dots : x_n]), with the at infinity \mathbb{P}^{n-1}_k = V(x_0) compactifying the in the . This structure ensures that projective varieties, as closed subsets of \mathbb{P}^n_k, are compact in the classical topology when k = \mathbb{C}, providing a foundation for studying global properties.

Homogeneous Ideals and Proj Construction

A graded ring S is a commutative ring equipped with a direct sum decomposition S = \bigoplus_{d \geq 0} S_d, where each S_d is an abelian group and the multiplication map satisfies S_m \cdot S_n \subseteq S_{m+n} for all m, n \geq 0. Elements of S_d are called homogeneous of degree d, and the decomposition allows for a natural \mathbb{Z}_{\geq 0}-grading on the ring. A prototypical example is the polynomial ring S = k[x_0, \dots, x_n] over a field k, graded by total degree, where each S_d consists of homogeneous polynomials of degree d. A homogeneous ideal I in a S is an ideal generated by homogeneous elements, equivalently, I = \bigoplus_{d \geq 0} (I \cap S_d). The irrelevant ideal, denoted S_+, is the homogeneous ideal \bigoplus_{d > 0} S_d, which consists of all elements of positive degree. For the k[x_0, \dots, x_n], S_+ is the ideal generated by x_0, \dots, x_n. Homogeneous ideals play a central role in defining subschemes, as the S/I inherits a grading whenever I is homogeneous. The Proj construction associates to a graded ring S (with S_+ \neq S) the space \operatorname{Proj} S, defined as the set of all homogeneous prime ideals of S that do not contain S_+. This set is equipped with the , whose basic open sets are the standard opens D_+(f) = \{ p \in \operatorname{Proj} S \mid f \notin p \} for homogeneous elements f \in S_d with d \geq 1. Each D_+(f) is homeomorphic to \operatorname{Spec} S_{(f)}, where S_{(f)} denotes the degree-zero part of the localization of S at the multiplicative set generated by f. To endow \operatorname{Proj} S with a scheme structure, the affine schemes \operatorname{Spec} S_{(f)} on the D_+(f) are glued along their intersections: D_+(f) \cap D_+(g) = D_+(fg) for homogeneous f, g, with the natural localization maps S_{(f)} \to S_{(fg)} and S_{(g)} \to S_{(fg)} ensuring compatibility. The structure sheaf \mathcal{O}_{\operatorname{Proj} S} is defined by \Gamma(D_+(f), \mathcal{O}_{\operatorname{Proj} S}) = S_{(f)} on basic opens, extended uniquely to all opens, yielding a ringed space that is a scheme. For twisting, the sheaf \mathcal{O}_{\operatorname{Proj} S}(d) arises from the graded S-module S(d) with components S(d)_i = S_{d+i}, and it satisfies \mathcal{O}_{\operatorname{Proj} S}(m) \otimes \mathcal{O}_{\operatorname{Proj} S}(n) \cong \mathcal{O}_{\operatorname{Proj} S}(m+n); in the case of projective space \mathbb{P}^n = \operatorname{Proj} k[x_0, \dots, x_n], these are the invertible sheaves \mathcal{O}_{\mathbb{P}^n}(d). Projective schemes are precisely those schemes isomorphic to \operatorname{Proj} S for some S, and there is an between closed subschemes of \mathbb{P}^n and quotients of k[x_0, \dots, x_n] by saturated homogeneous , where ensures the ideal is properly defined modulo units in degree zero. Specifically, a homogeneous I in k[x_0, \dots, x_n] defines a closed \operatorname{Proj} k[x_0, \dots, x_n]/I \hookrightarrow \mathbb{P}^n.

Relation to Affine Varieties

Projective space \mathbb{P}^n over an k admits a standard affine open cover consisting of n+1 sets U_i = \{ [x_0 : \cdots : x_n] \mid x_i \neq 0 \} for i = 0, \dots, n, where each U_i is isomorphic to affine n-space \mathbb{A}^n via the map sending [x_0 : \cdots : x_n] to (x_0/x_i, \dots, \hat{x_i}/x_i, \dots, x_n/x_i). This cover demonstrates that is locally affine, allowing the study of projective varieties through their intersections with these affine opens. Dehomogenization provides a explicit isomorphism between a projective variety X \subset \mathbb{P}^n intersected with U_i and an in \mathbb{A}^n. For a point in X \cap U_i, setting the i-th homogeneous coordinate to 1 yields affine coordinates, and the defining homogeneous equations of X restrict to polynomial equations on this affine chart after dehomogenization. Conversely, any embeds into via homogenization of its , yielding its projective closure. Projective varieties serve as compactifications of affine varieties by adjoining a " at ." Specifically, for the affine open U_0 \cong \mathbb{A}^n in \mathbb{P}^n, the complement is the H_\infty = \{x_0 = 0\} \cong \mathbb{P}^{n-1}, which adds points at to "complete" the affine space in the . The projective closure \tilde{Y} of an affine variety Y \subset \mathbb{A}^n is thus obtained by homogenizing the equations of Y and taking the zero set in \mathbb{P}^n, with Y = \tilde{Y} \cap U_0 and the points at forming \tilde{Y} \cap H_\infty. This construction ensures that \tilde{Y} is proper (compact in the classical sense over \mathbb{C}), unlike the non-compact affine Y. Quasi-projective varieties are defined as open subsets of , bridging the gap between and . For instance, every is quasi-projective, as it arises as the of its with an in . This class includes all varieties that can be embedded locally into while retaining affine-like behavior on opens.

Fundamental Properties

Completeness and Projective Morphisms

In , a X over a k is defined to be complete if, for every Y over k, the projection p_Y: X \times_k Y \to Y is a closed map. This ensures that the image of any from a complete X to another Y is closed in Y. Equivalently, over an , completeness implies that regular functions on a connected complete are constant. Properness generalizes completeness to morphisms. A morphism f: X \to Y of varieties over a field k is proper if it is of finite type, separated, and universally closed, meaning that for any base change Y' \to Y, the induced morphism X' \to Y' (where X' = X \times_Y Y') has closed image. An equivalent characterization, known as the valuative criterion of properness, states that for Noetherian schemes (or varieties), f is proper if and only if, for every R with fraction field K and any involving a \operatorname{Spec} K \to X and \operatorname{Spec} R \to Y, there exists a unique lift \operatorname{Spec} R \to X making the diagram commute. This criterion captures the "rigidity" of proper morphisms under extensions of valuation rings. Over an k, every projective variety is complete. Specifically, if X \subset \mathbb{P}^n_k is a closed subvariety, then the \mathbb{P}^n_k \times Y \to Y is closed for any Y, implying that X inherits this property as a closed subscheme. A f: X \to Y is projective if X is a closed subscheme of a bundle over Y, or equivalently, if there exists an on X such that f factors through the associated projective bundle. Projective s are proper: any such f is of finite type, separated, and universally closed, as the from satisfies these conditions. This properness yields a universal property for morphisms into projective varieties. If Z is a proper variety over an algebraically closed field k and f: Z \to X is a to a projective variety X \subset \mathbb{P}^n_k, then the f(Z) is closed in X, and f factors uniquely through the closed of f(Z) into X. Properness can be verified locally using affine covers of the target .

Homogeneous Coordinate Ring

For a projective variety X embedded in the projective space \mathbb{P}^n_k over an k, the homogeneous coordinate ring is the graded k- S(X) = \bigoplus_{d \geq 0} H^0(X, \mathcal{O}_X(d)), where \mathcal{O}_X(d) denotes the d-th power of the Serre twisting sheaf associated to the embedding, and H^0(X, \mathcal{O}_X(d)) is the k-vector space of global sections of this sheaf. This ring encodes the algebraic structure of X compatibly with its projective embedding, as the applied to S(X) recovers X by quotienting out the irrelevant ideal generated by the degree-1 elements. The defining X in the k[x_0, \dots, x_n] is a saturated homogeneous , meaning it equals its with respect to the maximal irrelevant (x_0, \dots, x_n), which ensures that the associated sheaf \tilde{I}(X) on \mathbb{P}^n is the ideal sheaf of X. Consequently, S(X) is isomorphic to k[x_0, \dots, x_n]/I(X), where I(X) is this saturated . The ring S(X) is particularly relevant when it is generated as a k- by its degree-1 component, a condition equivalent to X being projectively in the ; in this case, higher-degree sections are generated by linear forms, facilitating computations of invariants. The Hilbert function of X is defined as h_X(d) = \dim_k S(X)_d, measuring the dimension of the degree-d graded piece of the homogeneous coordinate ring. For small d, h_X(d) grows combinatorially, reflecting the initial constraints imposed by the , but for sufficiently large d, it stabilizes to a behavior that captures asymptotic growth related to the of X. This relates to the sheaf \tilde{I}(X) via the graded structure, where the minimal number of generators of I(X) as a homogeneous determines the of X in \mathbb{P}^n, providing a measure of how the variety sits in the ambient space.

Hilbert Polynomial and Degree

The Hilbert function h_X(d) of a projective variety X \subset \mathbb{P}^n over an is defined as the of the -d component of its homogeneous coordinate S(X) = k[x_0, \dots, x_n]/I(X), where I(X) is the homogeneous of X. This function measures the growth of sections of the \mathcal{O}_X(d). For large d, h_X(d) stabilizes and equals a P_X(t) \in \mathbb{Q} of equal to \dim X, known as the Hilbert of X. The leading coefficient of P_X(t) encodes key geometric invariants, specifically with leading term \frac{\deg X}{(\dim X)!} t^{\dim X}. The degree \deg X of an r-dimensional projective variety X is the integer such that the leading coefficient of P_X(t) is \frac{\deg X}{r!}. Geometrically, this is the intersection multiplicity of X with a general linear subspace of complementary dimension n - r, or the number of points in a general such intersection counted with multiplicity. This definition aligns the algebraic Hilbert polynomial with classical intersection theory on projective space. For curves (\dim X = 1), the Hilbert polynomial takes the form P_X(t) = (\deg X) t + 1 - p_a(X), where p_a(X) is the arithmetic genus, an invariant reflecting the topology of X via p_a(X) = 1 - \chi(\mathcal{O}_X). The arithmetic genus is non-negative and vanishes if and only if X \cong \mathbb{P}^1. Representative examples illustrate these concepts. The , the image of the Veronese embedding \nu_2: \mathbb{P}^2 \to \mathbb{P}^5, is a surface of degree 4, so its Hilbert polynomial has leading term \frac{4}{2!} t^2 = 2 t^2. Likewise, the \mathrm{Gr}(2,5), parametrizing 2-planes in \mathbb{C}^5 and embedded via the Plücker map into \mathbb{P}^9, is 6-dimensional with degree 5, yielding leading term \frac{5}{6!} t^6.

Examples and Invariants

Projective Curves

Projective curves are one-dimensional projective varieties, typically realized as closed subschemes of \mathbb{P}^n defined by homogeneous equations. These objects provide foundational examples in , bridging affine curves to compact settings through homogenization and exhibiting rich invariants like and . Unlike higher-dimensional varieties, curves admit complete by , with ensuring they are Riemann surfaces over the numbers. The simplest projective curve is the rational curve, exemplified by the projective line \mathbb{P}^1, which has genus 0 and serves as the model for all smooth projective curves of genus 0. \mathbb{P}^1 embeds into \mathbb{P}^2 as a line, a degree 1 curve, and any birational map from \mathbb{P}^1 to a plane curve of degree d parametrizes rational curves of that degree. Rational curves are characterized by their parametrization by rational functions, reflecting the function field isomorphic to k(t) for a field k. Elliptic curves represent the next level of complexity, defined as smooth projective curves of genus 1 equipped with a base point, though the often omits explicit reference to the point. A standard into \mathbb{P}^2 arises from homogenizing the affine Weierstrass form y^2 = x^3 + ax + b (with discriminant nonzero to ensure smoothness), yielding the projective equation Y^2 Z = X^3 + a X Z^2 + b Z^3. This cubic model has degree 3 and genus 1, capturing the curve's complete structure in . For smooth plane projective curves embedded in \mathbb{P}^2, the genus g relates directly to the degree d via the formula g = \frac{(d-1)(d-2)}{2}. This relation, classical in , quantifies how higher-degree embeddings increase topological complexity, with d=1 or d=2 yielding genus 0 (lines and conics, isomorphic to \mathbb{P}^1) and d=3 giving genus 1 (elliptic curves). The degree here aligns with the Hilbert polynomial's leading coefficient, aiding genus computations through . Singular projective curves, defined by homogeneous ideals where the scheme has non-reduced structure or self-intersections, require to recover a smooth model. The \tilde{C} \to C is the unique birational from a smooth projective curve \tilde{C} (the integral closure of the coordinate ) that resolves singularities, such as nodes or cusps, by separating branches at singular points. For instance, both a nodal cubic and a cuspidal cubic in \mathbb{P}^2 normalize to \mathbb{P}^1 ( 0), preserving the arithmetic genus but adjusting the geometric genus to reflect the smooth . This process is always possible for curves, as singularities lie in codimension 1.

Projective Hypersurfaces

A projective hypersurface in \mathbb{P}^n over an k is defined as the zero locus V(f) = \{ [x_0 : \cdots : x_n] \in \mathbb{P}^n \mid f(x_0, \dots, x_n) = 0 \}, where f \in k[x_0, \dots, x_n] is a of positive .$$] If f is irreducible, then V(f) is an irreducible projective variety of n-1 and \deg(f). These arise naturally from affine hypersurfaces via homogenization, where an affine g(y_1, \dots, y_n) = 0 in \mathbb{A}^n is extended to the projective closure by setting f(x_1, \dots, x_n, x_0) = x_0^{\deg(g)} g(x_1/x_0, \dots, x_n/x_0). Singularities on a projective hypersurface V(f) occur at points where f and all its partial derivatives \partial f / \partial x_i vanish simultaneously. For an isolated hypersurface singularity at the origin in the local affine model, the Milnor number \mu(f) quantifies the topological complexity of the singularity and is given by \mu(f) = \dim_k \mathcal{O}_{n,0} / (f, \partial f / \partial x_1, \dots, \partial f / \partial x_n), where \mathcal{O}_{n,0} is the local ring of the at the in \mathbb{C}^n; this invariant is constant under small deformations and equals the rank of the middle of the Milnor fiber.[ The Tyurina algebra, associated to the versal deformation of the singularity, is the quotient $\mathcal{O}_{n,0} / T(f)$, where the Tyurina ideal $T(f)$ is generated by $f$ and the partial derivatives together with relations from the [embedding](/page/Embedding); its dimension, the Tjurina number $\tau(f)$, satisfies $\tau(f) \leq \mu(f)$, with equality holding for quasi-homogeneous singularities.] These numbers classify the local equisingularity type and bound the dimension of the of deformations. Resolution of singularities for projective hypersurfaces over fields of characteristic zero can be achieved through a finite sequence of blow-ups along nonsingular subvarieties, yielding a smooth proper birational model \tilde{V} \to V(f) with exceptional divisors that are themselves projective hypersurfaces.[ Hironaka's theorem guarantees such a resolution exists for any algebraic variety, but for hypersurfaces, the process often simplifies due to codimension one, with blow-ups centered at the singular locus iteratively reducing multiplicity until smoothness is attained; for example, blowing up the maximal ideal at a point singularity replaces it with a projective space bundle over the exceptional divisor.] The scheme F_1(V(f)) of a projective parametrizes the 1-dimensional linear subspaces (lines) contained in V(f), constructed as the zero locus of a universal section of a on the \mathrm{Gr}(2, n+1).[ For a [smooth](/page/Smooth) cubic [hypersurface](/page/Hypersurface) in $\mathbb{P}^4$, this Fano surface is a [smooth](/page/Smooth) projective surface of geometric [genus](/page/Genus) 5 whose [geometry](/page/Geometry) encodes [rationality](/page/Rationality) properties of the threefold, as studied via its relation to the intermediate Jacobian.] In higher dimensions, such as cubics in \mathbb{P}^5, the Fano scheme of lines is a fourfold that aids in understanding birational invariants and period domains.

Abelian Varieties

An over a k is defined as a nonsingular projective A that is also a commutative algebraic group, meaning the group operations of addition and inversion are given by morphisms of varieties. This structure ensures that A is complete and geometrically connected, inheriting projectivity from its proper morphism properties as a . The commutativity follows from the rigidity of group laws on projective varieties, making the addition map symmetric. A key feature of abelian varieties is the notion of polarization, which provides a positivity structure compatible with the group law. A polarization on an abelian variety A of dimension g is an ample line bundle L on A, up to translation by points of A, that induces a homomorphism \lambda_L: A \to \hat{A} to the dual abelian variety \hat{A} = \Pic^0(A), where the kernel is finite and the induced map on the N-torsion is an isogeny of degree N^{2g}. A principal polarization occurs when this homomorphism is an isomorphism, corresponding to a degree-1 ample line bundle; in such cases, the zero section of L defines an effective ample divisor known as the theta divisor \Theta, which embeds A into projective space via the complete linear system |n\Theta| for sufficiently large n. For example, on a principally polarized abelian variety, the theta divisor \Theta is ample and its associated line bundle L(\Theta) satisfies \chi(L(\Theta)) = 1, highlighting the principal nature. A fundamental example of an is the of a smooth projective C of g \geq 1 over k. The J(C) is the parametrizing degree-zero line bundles on C, realized as a projective variety of g with a natural principal induced by the theta divisor, which corresponds to the of effective divisors of g-1 on C. This construction shows that every abelian variety over an infinite field admits a surjective from some , underscoring their role in the of s. Abelian varieties have dimension g \geq 1, where the tangent space at the identity is a vector space of that dimension, and the group law is analytic in local coordinates. The endomorphism ring \End(A) consists of all morphisms from A to itself as a group scheme, forming a ring that acts faithfully on the tangent space at the identity; over algebraically closed fields of characteristic zero, \End(A) \otimes \mathbb{Q} is a semisimple \mathbb{Q}-algebra of finite rank, often commutative for simple abelian varieties. For instance, when A is an elliptic curve (dimension 1), \End(A) is either \mathbb{Z} or an order in a quadratic imaginary field.

Morphisms and Embeddings

Projections and Veronese Embeddings

Projections from or provide essential rational maps in the study of projective varieties, allowing reduction of embedding dimensions while preserving key geometric properties. Consider a projective variety X \subset \mathbb{P}^n over an and C \subset \mathbb{P}^n disjoint from X. The projection \pi_C: \mathbb{P}^n \dashrightarrow \mathbb{P}^{n - \dim C - 1} with center C is a rational map defined by sending p \in \mathbb{P}^n \setminus C to the intersection of the line joining p to a generic point in C with a complementary linear subspace. Restricting to X, this induces a rational map \pi_C|_X: X \dashrightarrow Y \subset \mathbb{P}^{n - \dim C - 1}, where Y is the closure of the image, provided the center avoids the tangent spaces to X. Such projections are undefined along the cone over X with vertex C, but resolve to morphisms after blowing up the center. The Veronese embedding offers a canonical method to embed \mathbb{P}^n into a higher-dimensional using homogeneous polynomials of fixed degree. For d \geq 1, the d-th Veronese map v_d: \mathbb{P}^n \to \mathbb{P}^N, where N = \binom{n+d}{d} - 1, sends [x_0 : \cdots : x_n] to the point whose coordinates are all monomials of degree d in the x_i, up to scalar. This morphism is defined by the complete linear system | \mathcal{O}_{\mathbb{P}^n}(d) |, and the image v_d(\mathbb{P}^n) is a projective variety of dimension n and degree d^n. The map v_d is an , hence very ample, meaning \mathcal{O}_{\mathbb{P}^n}(d) generates the embedding and separates points and tangent vectors. More generally, for a projective variety X \subset \mathbb{P}^n, the restriction of v_d to X yields an embedding if the restriction of \mathcal{O}_{\mathbb{P}^n}(d) to X is very ample. A line bundle L on a projective variety X is very ample if the associated morphism \phi_{|L|}: X \to \mathbb{P}^H to , given by a basis of global sections H^0(X, L), is an . While the Nakai-Moishezon characterizes ampleness—a prerequisite for very ampleness—via intersection numbers, stating that L is ample L^{\dim V} \cdot V > 0 for every irreducible subvariety V \subset X, very ampleness requires additional separation conditions. Briefly, L is very ample if it is ample and the map separates points and tangents, with the providing a numerical test for the underlying positivity. Birational projections, particularly generic linear projections from a point or outside the , play a role in studying invariants under birational equivalence. For a non-degenerate projective X \subset \mathbb{P}^r of n and c \geq 2, a generic from a point not on X or its spaces induces a birational map onto its image in \mathbb{P}^{r-1}, provided c \geq 2. Such projections preserve the birational type of X, and consequently, the of X—defined as the multiplicity with a general of complementary —remains unchanged, as birational morphisms between projective of the same are degree-preserving isomorphisms in 1. This invariance facilitates computations of degrees in lower embeddings without altering intrinsic properties.

Linear Systems and Dual Varieties

In , the complete linear system associated to a Cartier divisor D on a projective variety X is the \mathbb{P} H^0(X, \mathcal{O}_X(D)), parametrizing the effective divisors linearly equivalent to D. This space consists of all global sections of the line bundle \mathcal{O}_X(D), up to scalar multiple, and its dimension is h^0(X, \mathcal{O}_X(D)) - 1. If the linear system is basepoint-free, it defines a \phi_{|D|}: X \to \mathbb{P}^H, where H = h^0(X, \mathcal{O}_X(D)) - 1, embedding X into via the evaluation map that sends a point x \in X to the of sections vanishing at x. By the on very ample line bundles, if \mathcal{O}_X(D) is very ample, this morphism is an . The variety of a X \subset \mathbb{P}^N is defined as the closure in the (\mathbb{P}^N)^\vee of the set of all s tangent to X at some point. A H \in (\mathbb{P}^N)^\vee is tangent to X at x \in X_{\text{smooth}} if it contains the embedded T_x X. For a irreducible X, the variety X^\vee is typically a , but its in (\mathbb{P}^N)^\vee may exceed 1, leading to the notion of defect. The defect of X, denoted \delta(X), is given by \delta(X) = \dim(\mathbb{P}^N)^\vee - \dim X^\vee - 1. with positive defect are ruled by linear spaces of dimension \delta(X), and \delta(X) > 0 implies that the K_X is not nef. For smooth projective varieties, the biduality theorem asserts that the dual of the dual variety recovers the original: X^{\vee\vee} = X. This reflexivity holds because the Gauss map \gamma: X \to \mathrm{Gr}(\dim X + 1, N + 1), sending each point to its , is finite and birational onto its image when X is , ensuring the incidence between points and tangent hyperplanes is proper. If both X and X^\vee are , then \dim X = \dim X^\vee. Examples of varieties achieving biduality with positive defect include quadrics and certain Segre embeddings, such as \mathbb{P}^n \times \mathbb{P}^1. Bertini's theorem provides a genericity result for s on projective varieties. For a basepoint-free |D| on a projective variety X over an of characteristic zero, there exists a Zariski-open dense U \subset |D| such that every effective E \in U is of codimension 1 in X. More generally, the theorem applies to ample line bundles, ensuring that general sections are and connected, away from the base locus. This result relies on the properness of the induced by |D| and the of the of singular loci in families. In positive characteristic, additional hypotheses like reducedness may be needed to avoid pathologies.

Coherent Sheaves and Cohomology

Structure of Coherent Sheaves

A coherent sheaf on a projective variety X is a sheaf \mathcal{F} of \mathcal{O}_X- that is locally of finite presentation, meaning that on every affine open subset U = \operatorname{Spec}(A) of X, the restriction \mathcal{F}|_U is a finitely presented A-. Since projective varieties are Noetherian schemes, coherent sheaves arise as the sheafification of finitely generated graded over the homogeneous coordinate of X, and they can be either locally free ( bundles) or torsion (supported on proper subvarieties). This structure ensures that coherent sheaves capture the essential algebraic data of subschemes and bundles on X, forming the category \operatorname{Coh}(X) which is abelian. The Hilbert syzygy theorem provides a key resolution property for coherent sheaves on projective space \mathbb{P}^n. It states that every coherent sheaf \mathcal{F} on \mathbb{P}^n admits a finite resolution by locally free sheaves of length at most n+1, where n is the dimension of \mathbb{P}^n. More precisely, associating to \mathcal{F} the graded module E = \bigoplus_m H^0(\mathbb{P}^n, \mathcal{F}(m)) over the polynomial ring S = k[z_0, \dots, z_n], the projective dimension of E is bounded by n+1, reflecting the global generation of sheaves on projective varieties embedded in \mathbb{P}^n. This bound extends to coherent sheaves on arbitrary projective varieties via embeddings, limiting the complexity of syzygies in their minimal free resolutions. For vector bundles on \mathbb{P}^n, the Beilinson monad offers an explicit resolution using exterior powers of the tautological bundle. Specifically, any coherent sheaf \mathcal{F} on \mathbb{P}^n can be resolved by a monad of the form [ 0 \to \bigoplus_i \Omega^j(n_j) \to \bigoplus_i \mathcal{O}{\mathbb{P}^n}(m_i) \to \bigoplus_i \mathcal{O}{\mathbb{P}^n}(l_i) \to \mathcal{F} \to 0, where $\Omega^j$ denotes the $j$-th exterior power of the [cotangent bundle](/page/Cotangent_bundle), and the shifts are determined by the [cohomology](/page/Cohomology) of $\mathcal{F}$.[](https://link.springer.com/article/10.1007/BF01681436) This construction, derived from the bounded [derived category](/page/Derived_category) of coherent sheaves, provides a uniform way to describe indecomposable vector bundles and facilitates computations of Ext groups.[](https://link.springer.com/article/10.1007/BF01681436) Stability notions refine the structure of coherent sheaves, enabling the study of moduli spaces. Slope stability, introduced for vector bundles, defines a torsion-free sheaf $\mathcal{E}$ on a polarized projective variety $(X, \mathcal{O}_X(1))$ as $\mu$-stable if for every proper subsheaf $\mathcal{F} \subset \mathcal{E}$, the slope $\mu(\mathcal{F}) < \mu(\mathcal{E})$, where $\mu(\mathcal{F}) = \frac{c_1(\mathcal{F}) \cdot H}{\operatorname{rk}(\mathcal{F})}$ with $H$ the hyperplane class.[](https://www.dam.brown.edu/people/mumford/alg_geom/papers/1977a--StabilityLecturesIHES-Swiss.pdf) Gieseker stability generalizes this to a Hilbert polynomial comparison: a sheaf $\mathcal{E}$ is Gieseker-stable if for every proper subsheaf $\mathcal{F} \subset \mathcal{E}$, the normalized Hilbert polynomial $P(\mathcal{F}, m)/\operatorname{rk}(\mathcal{F}) < P(\mathcal{E}, m)/\operatorname{rk}(\mathcal{E})$ for large $m$, prioritizing higher-degree terms over slope alone.[](https://projecteuclid.org/journals/journal-of-differential-geometry/volume-27/issue-1/A-construction-of-stable-bundles-on-an-algebraic-surface/10.4310/jdg/1214441654.pdf) These conditions ensure boundedness and the existence of projective moduli spaces for semistable sheaves on smooth projective varieties.[](https://projecteuclid.org/journals/journal-of-differential-geometry/volume-27/issue-1/A-construction-of-stable-bundles-on-an-algebraic-surface/10.4310/jdg/1214441654.pdf) ### Cohomology Groups on Projective Varieties Sheaf cohomology provides a fundamental tool for studying global properties of coherent sheaves on projective varieties, extending the notion of cohomology from topology to algebraic geometry. On a projective variety $X$, such as the projective space $\mathbb{P}^n$, sheaf cohomology groups $H^i(X, \mathcal{F})$ for a coherent sheaf $\mathcal{F}$ can be computed using Čech cohomology with respect to a suitable open cover. Specifically, Čech cohomology is defined via an open cover $\mathcal{U} = \{U_\alpha\}$ of $X$, where the Čech complex is formed by taking alternating sums of sections over intersections: the 0-th term is $\prod_\alpha \mathcal{F}(U_\alpha)$, the 1-st term is $\prod_{\alpha < \beta} \mathcal{F}(U_{\alpha\beta})$, and higher terms similarly, with the cohomology of this complex yielding the Čech groups $\check{H}^i(\mathcal{U}, \mathcal{F})$. For projective space $\mathbb{P}^n$ over an algebraically closed field $k$, the standard affine cover $\mathcal{U} = \{D(x_i)\}_{i=0}^n$, where $D(x_i)$ are the principal open sets defined by the homogeneous coordinates $x_i$, has the property that all intersections $U_{i_0 \dots i_j} = D(x_{i_0}) \cap \cdots \cap D(x_{i_j})$ are affine schemes. Since higher cohomology vanishes on affine schemes (i.e., $H^j(U, \mathcal{G}) = 0$ for $j > 0$ and quasi-coherent $\mathcal{G}$), the Leray theorem ensures that $\check{H}^i(\mathcal{U}, \mathcal{F}) \cong H^i(\mathbb{P}^n, \mathcal{F})$ for coherent $\mathcal{F}$, allowing explicit computations via the finite-dimensional Čech complex.[](https://stacks.math.columbia.edu/tag/01XS) A key result is the finite-dimensionality of these cohomology groups on projective varieties. For $\mathbb{P}^n$ over a [field](/page/Field) $k$ and [coherent sheaf](/page/Coherent_sheaf) $\mathcal{F}$, each $H^i(\mathbb{P}^n, \mathcal{F})$ is a finite-dimensional $k$-[vector space](/page/Vector_space), and moreover, $H^i(\mathbb{P}^n, \mathcal{F}) = 0$ for all $i > n$. This follows from the structure of the Čech complex for the standard affine cover, which has length $n+1$ (yielding potential non-zero cohomology up to degree $n$), combined with the vanishing of higher [cohomology](/page/Cohomology) on the affine intersections; the argument proceeds by [induction](/page/Induction) on the [dimension](/page/Dimension) $n$, reducing to lower-dimensional projective spaces via exact sequences or spectral sequences associated to the cover. More generally, on any projective variety $X$ of [dimension](/page/Dimension) $d$ over $k$, Serre's theorem asserts that $H^i(X, \mathcal{F})$ is finite-dimensional for coherent $\mathcal{F}$ and vanishes for $i > d$, establishing the cohomological [dimension](/page/Dimension) bounded by the geometry of $X$. These properties distinguish projective varieties from affine ones, where higher [cohomology](/page/Cohomology) always vanishes, and enable inductive computations across [algebraic geometry](/page/Algebraic_geometry). Explicit computations of cohomology groups are available for important classes of sheaves on $\mathbb{P}^n$, such as the twisted differentials $\Omega^p(k) = \Omega^p_{\mathbb{P}^n} \otimes \mathcal{O}_{\mathbb{P}^n}(k)$, where $\Omega^p_{\mathbb{P}^n}$ is the sheaf of holomorphic (or algebraic) $p$-forms. The Bott formula provides the precise dimensions: for $k = 0$, $\dim H^q(\mathbb{P}^n, \Omega^p) = 1$ if $p = q$ and 0 otherwise; for $k > 0$, $H^q(\mathbb{P}^n, \Omega^p(k)) = 0$ unless $q = 0$ and $k > p$, in which case $\dim H^0(\mathbb{P}^n, \Omega^p(k)) = \binom{n + k - p}{k} \binom{k - 1}{p}$; for $k < 0$, $H^q(\mathbb{P}^n, \Omega^p(k)) = 0$ unless $q = n$ and $k < p - n$, in which case $\dim H^n(\mathbb{P}^n, \Omega^p(k)) = \binom{-k + p - 1}{-k} \binom{-k - 1}{n - p}$. This formula, derived using the Euler sequence and induction on $p$, highlights strong vanishing phenomena, such as $H^q(\mathbb{P}^n, \Omega^p(k)) = 0$ for $q > 0$ and $k \gg 0$, and is crucial for studying deformations and obstructions in moduli problems.[](https://arxiv.org/pdf/math/9904110) The long exact sequence in cohomology arises from short exact sequences of sheaves and is instrumental for relating cohomology groups across extensions. Given a short exact sequence $0 \to \mathcal{A} \to \mathcal{B} \to \mathcal{C} \to 0$ of coherent sheaves on $\mathbb{P}^n$, the associated long exact sequence is \cdots \to H^{i-1}(\mathbb{P}^n, \mathcal{C}) \to H^i(\mathbb{P}^n, \mathcal{A}) \to H^i(\mathbb{P}^n, \mathcal{B}) \to H^i(\mathbb{P}^n, \mathcal{C}) \to H^{i+1}(\mathbb{P}^n, \mathcal{A}) \to \cdots, which preserves exactness and finiteness. This sequence applies to study extensions of sheaves: the extensions of $\mathcal{C}$ by $\mathcal{A}$ (i.e., short exact sequences with middle term varying) are classified up to isomorphism by the cohomology group $H^1(\mathbb{P}^n, \mathcal{E}xt^0(\mathcal{C}, \mathcal{A}))$, where $\mathcal{E}xt^0(\mathcal{C}, \mathcal{A}) = \underline{\Hom}(\mathcal{C}, \mathcal{A})$ is the sheaf Hom; for locally free sheaves (vector bundles), this simplifies to $H^1(\mathbb{P}^n, \mathcal{C}^\vee \otimes \mathcal{A})$. Such applications allow inductive determination of cohomology for more complex sheaves from simpler ones, as in resolving a coherent sheaf via a finite free resolution and using the sequence repeatedly. ### Ring of Global Sections The ring of global sections $ H^0(X, \mathcal{O}_X(d)) $ for a projective variety $ X \subset \mathbb{P}^N $ over [an algebraically closed field](/page/Algebraically_closed_field) forms the degree-$ d $ component of the homogeneous coordinate [ring](/page/Ring) $ S(X) = \bigoplus_{d=0}^\infty H^0(X, \mathcal{O}_X(d)) $, which encodes the algebraic structure of $ X $ as a subvariety.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) This [graded ring](/page/Graded_ring) is finitely generated as [an algebra](/page/Algebra) over the [polynomial ring](/page/Polynomial_ring) $ k[x_0, \dots, x_N] $, with the irrelevant [ideal](/page/Ideal) defining $ X $ as its Proj.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) When $ X $ is embedded via a very ample line bundle $ \mathcal{L} $, identified with $ \mathcal{O}_X(1) $, the global sections $ H^0(X, \mathcal{L}^{\otimes d}) $ generate the [symmetric algebra](/page/Symmetric_algebra) $ \mathrm{Sym}(H^0(X, \mathcal{L})) $ in degree $ d $, providing the relations that cut out $ X $ in the [projective space](/page/Projective_space) $ \mathbb{P}(H^0(X, \mathcal{L})^*) $.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) For [normal](/page/Normal) projective varieties, Serre's [theorem](/page/Theorem) asserts that there exists an integer $ d_0 $ such that for all $ d \geq d_0 $, the sheaf $ \mathcal{O}_X(d) $ is generated by its global sections $ H^0(X, \mathcal{O}_X(d)) $ at every point, meaning the evaluation map $ \mathcal{O}_X \otimes H^0(X, \mathcal{O}_X(d)) \to \mathcal{O}_X(d) $ is surjective.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) Moreover, the higher [cohomology](/page/Cohomology) groups vanish: $ H^i(X, \mathcal{O}_X(d)) = 0 $ for all $ i > 0 $ and $ d \gg 0 $, ensuring that the global sections fully capture the sheaf without cohomological obstructions.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) This vanishing of higher cohomology for large twists parallels later analytic results like Kodaira's vanishing theorem in the [complex](/page/Complex) case. The generation property implies that for sufficiently large $ d $, the complete [linear system](/page/Linear_system) $ |\mathcal{O}_X(d)| $ is very ample, yielding an [embedding](/page/Embedding) of $ X $ into [projective space](/page/Projective_space) that realizes $ X $ as a closed subvariety defined by the relations among the sections.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) In the context of singularities, the [normalization](/page/Normalization) $ \tilde{X} $ of a projective variety $ X $ is itself projective, as the normalization morphism is finite, preserving projectivity; the global sections of $ \tilde{\mathcal{O}}_{\tilde{X}}(d) $, pulled back from ample bundles on $ X $, generate an [embedding](/page/Embedding) of the resolved variety, allowing singularities to be addressed while maintaining the projective structure.[](https://stacks.math.columbia.edu/tag/0GK4) ## Advanced Theorems for Smooth Varieties ### Serre Duality Serre duality establishes a profound connection between the cohomology groups of coherent sheaves on a [smooth](/page/Smooth) projective variety and those of their duals twisted by the canonical sheaf. This theorem, originally proved by [Jean-Pierre Serre](/page/Jean-Pierre_Serre) in 1955, applies to a [smooth](/page/Smooth) projective variety $X$ of dimension $n$ over an [algebraically closed field](/page/Algebraically_closed_field) $k$, and a [coherent sheaf](/page/Coherent_sheaf) $\mathcal{F}$ on $X$. The duality asserts that there is a natural [isomorphism](/page/Isomorphism) of [vector space](/page/Vector_space)s H^i(X, \mathcal{F})^\vee \cong H^{n-i}(X, \mathcal{H}om(\mathcal{F}, \mathcal{O}_X) \otimes \omega_X), where the superscript $\vee$ on the left denotes the $k$-dual vector space, $\mathcal{H}om(\mathcal{F}, \mathcal{O}_X)$ is the sheaf Hom to the structure sheaf, and $\omega_X$ is the canonical sheaf of $X$.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) This isomorphism is functorial in $\mathcal{F}$ and compatible with the cup-product structure on [cohomology](/page/Cohomology).[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) The canonical sheaf $\omega_X$ is defined as the [line bundle](/page/Line_bundle) $\det \Omega_X^1 = \bigwedge^n \Omega_X^1$, where $\Omega_X^1$ is the cotangent sheaf of $X$; locally, it is generated by a nowhere-vanishing $n$-form serving as a [volume element](/page/Volume_element).[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) A sketch of the proof relies on [Čech cohomology](/page/Čech_cohomology) computed with respect to an open affine cover of $X$. The key construction involves a non-degenerate bilinear [pairing](/page/Pairing) between $H^i(X, \mathcal{F})$ and $H^{n-i}(X, \mathcal{H}om(\mathcal{F}, \mathcal{O}_X) \otimes \omega_X)$, induced by residue maps. These residue maps are defined locally using a system of parameters and Thom-like residues along coordinate hyperplanes, extending to global [cohomology](/page/Cohomology) via the trace map from $H^n(X, \omega_X)$ to $k$. The pairing's non-degeneracy follows from the vanishing of cohomology on affines and the projective embedding properties, yielding the desired [isomorphism](/page/Isomorphism).[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) One immediate application arises in the study of [canonical](/page/Canonical) divisors: Serre duality implies that $H^n(X, \omega_X) \cong k$, identifying the top [cohomology](/page/Cohomology) of the [canonical](/page/Canonical) sheaf with the [ground](/page/Ground) [field](/page/Field), which reflects the [existence](/page/Existence) of a global [volume form](/page/Volume_form) up to scalar.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) For smooth hypersurfaces, the [adjunction formula](/page/Adjunction_formula) computes the [canonical](/page/Canonical) sheaf explicitly: if $Y \subset X$ is a smooth effective [Cartier](/page/Cartier) divisor on a smooth projective variety $X$, then $\omega_Y \cong (\omega_X \otimes \mathcal{O}_X(Y))|_Y$. This relation derives from the short [exact sequence](/page/Exact_sequence) of cotangent sheaves $0 \to \mathcal{O}_X(-Y) \to \Omega_X^1|_Y \to \Omega_Y^1 \to 0$, taking determinants and using the [identification](/page/Identification) $\det \mathcal{O}_X(-Y) = \mathcal{O}_X(-Y)$.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) In the case of a smooth [hypersurface](/page/Hypersurface) of degree $d$ in $\mathbb{P}^{n+1}$, this yields $\omega_Y \cong \mathcal{O}_Y(d - n - 2)$, highlighting how duality constrains the geometry of embeddings.[](https://www.college-de-france.fr/media/jean-pierre-serre/UPL5435398796951750634_Serre_FAC.pdf) ### Riemann-Roch Theorem The Hirzebruch-Riemann-Roch theorem provides a formula for the Euler characteristic of a coherent sheaf on a smooth projective variety, generalizing the classical Riemann-Roch theorem from dimension one to arbitrary dimensions. For a smooth projective variety $X$ over the complex numbers and a coherent sheaf $\mathcal{F}$ on $X$, the theorem states that \chi(X, \mathcal{F}) = \int_X \operatorname{ch}(\mathcal{F}) \operatorname{td}(T_X), where $\chi(X, \mathcal{F})$ is the Euler characteristic, defined via Serre duality as the alternating sum of dimensions of cohomology groups $\sum_i (-1)^i h^i(X, \mathcal{F})$, $\operatorname{ch}(\mathcal{F})$ is the Chern character of $\mathcal{F}$, $T_X$ is the tangent sheaf of $X$, and $\operatorname{td}(T_X)$ is the Todd class of $T_X$. This integral is taken in the Chow ring or cohomology ring of $X$, yielding a rational number equal to the holomorphic Euler characteristic. The theorem was proved by Friedrich Hirzebruch using topological methods involving the $\hat{A}$-genus and cobordism theory. In the special case where $X$ is a smooth projective curve of genus $g$ and $\mathcal{F} = \mathcal{L}$ is a line bundle of degree $d$, the formula simplifies to \chi(X, \mathcal{L}) = d + 1 - g. This classical form, originally established by [Bernhard Riemann](/page/Bernhard_Riemann) and Gustav Roch, relates the topology of the curve (via [genus](/page/Genus)) to the arithmetic of [line bundle](/page/Line_bundle)s and underpins much of the theory of divisors on curves. Hirzebruch's proof relies on characteristic classes in [topology](/page/Topology): it reduces the problem to computing the index of the Dolbeault operator on vector bundles using the Todd genus, which agrees with the $\hat{A}$-genus for complex manifolds via the Hirzebruch signature theorem. An alternative analytic proof follows from the Atiyah-Singer index theorem, which equates the analytic index of an elliptic operator (such as the $\bar{\partial}$-complex for $\mathcal{F}$) to a topological index expressed via the Chern character and Todd class. For singular projective varieties, the theorem extends by replacing the tangent sheaf with a suitable notion of tangent complex or using [homology](/page/Homology) theories; the [Euler characteristic](/page/Euler_characteristic) is then given by integration against a [homology](/page/Homology) Todd class defined via blow-ups or [resolution of singularities](/page/Resolution_of_singularities). This generalization, which preserves the multiplicative structure over proper morphisms, was developed using [intersection theory](/page/Intersection_theory) on singular schemes. ### Kodaira Vanishing Theorem The Kodaira vanishing theorem provides a fundamental result on the [cohomology](/page/Cohomology) of coherent sheaves twisted by [ample line bundle](/page/Ample_line_bundle)s on smooth projective varieties over the complex numbers. Specifically, let $X$ be a smooth projective variety over $\mathbb{C}$ and $L$ an [ample line bundle](/page/Ample_line_bundle) on $X$. Then $H^i(X, \omega_X \otimes L) = 0$ for all $i > 0$.[](https://math.uchicago.edu/~amathew/kodaira.pdf) This vanishing implies, in particular, that the global sections of $\omega_X \otimes L$ generate the sheaf, facilitating embeddings and computations in [algebraic geometry](/page/Algebraic_geometry).[](https://people.math.harvard.edu/~mpopa/483-3/notes.pdf) The proof of the theorem relies on the analytic structure of complex projective varieties, which are Kähler manifolds, and employs [Hodge theory](/page/Hodge_theory) to analyze [Dolbeault cohomology](/page/Dolbeault_cohomology). One identifies the sheaf cohomology $H^i(X, \omega_X \otimes L)$ with the [Dolbeault cohomology](/page/Dolbeault_cohomology) $H^{n,i}(X, L)$, resolved via the $\bar{\partial}$-complex. The key step involves the $\bar{\partial}$-Laplacian operator $\Delta_{\bar{\partial}} = \bar{\partial} \bar{\partial}^* + \bar{\partial}^* \bar{\partial}$, whose positivity is established using the Chern connection on $L$ and the Kähler metric. Since $L$ is ample, its curvature form is a positive (1,1)-form, ensuring that the [commutator](/page/Commutator) $[\nabla^2, \Lambda]$ (where $\Lambda$ is the adjoint of the Lefschetz operator) contributes a positive term via the Kodaira-Nakano identity $\Delta_{\bar{\partial}} = \Delta'' + i [\bar{\partial}, \bar{\partial}^*] + \sqrt{-1} [\nabla^2, \Lambda]$. This positivity implies no nonzero [harmonic](/page/Harmonic) forms exist in degrees $i > 0$, hence the vanishing.[](https://math.uchicago.edu/~amathew/kodaira.pdf) The argument extends to higher powers $L^k$ for $k \geq 1$ directly due to ampleness.[](https://people.math.harvard.edu/~mpopa/papers/oxford.pdf) A significant generalization is the Akizuki–Nakano vanishing theorem, which refines the result for bundles of holomorphic forms. For the same $X$ and ample $L$, the cohomology satisfies $H^q(X, \Omega^p_X \otimes L) = 0$ whenever $p + q > \dim X$.[](https://math.uchicago.edu/~amathew/kodaira.pdf) This follows from a similar Hodge-theoretic argument, applying the positivity of the Laplacian to (p,q)-forms with values in $\Omega^p_X \otimes L$, and leverages the full Kähler identities to control the degrees. The theorem plays a crucial role in computations involving the [cotangent bundle](/page/Cotangent_bundle) and deformations of varieties.[](https://arxiv.org/pdf/1912.12074) While the theorem holds over $\mathbb{C}$, it does not generalize to arbitrary fields. In characteristic $p > 0$, counterexamples exist; for instance, Raynaud constructed smooth projective surfaces $X$ in characteristic $p$ with an [ample line bundle](/page/Ample_line_bundle) $L$ such that $H^1(X, \omega_X \otimes L) \neq 0$, violating the expected vanishing.[](https://www.ias.ac.in/article/fulltext/pmsc/107/01/0021-0025) These examples, often involving ruled surfaces over curves with specific Frobenius actions, highlight the role of transcendental methods in the original proof, which fail algebraically in positive characteristic.[](https://www.math.utah.edu/~schwede/frob/Notes12-7-2010.pdf) ## Moduli and Parameter Spaces ### Hilbert Schemes The [Hilbert scheme](/page/Hilbert_scheme) $\Hilb^P(\mathbb{P}^n)$, where $P$ is a fixed [polynomial](/page/Polynomial) with integer coefficients, is the [moduli space](/page/Moduli_space) parametrizing closed subschemes of $\mathbb{P}^n$ whose structure sheaves have [Hilbert polynomial](/page/Hilbert_scheme) $P$. It represents the contravariant [functor](/page/Functor) from the [category](/page/Category) of [schemes](/page/Scheme) to sets that associates to any [scheme](/page/Scheme) $S$ the set of $S$-flat families of closed subschemes $Z \hookrightarrow \mathbb{P}^n \times_S S$ such that for every point $s \in S$, the fiber $Z_s$ is a closed subscheme of $\mathbb{P}^n$ with [Hilbert polynomial](/page/Hilbert_scheme) $P$. This [functor](/page/Functor) is representable by a projective [scheme](/page/Scheme) over the base field, as shown by Grothendieck in his construction using the theory of quot schemes and flattening stratifications.[](https://www.numdam.org/article/PMIHES_1966__29__5_0.pdf) A point $[Z]$ in $\Hilb^P(\mathbb{P}^n)$ corresponds to a closed subscheme $Z \subset \mathbb{P}^n$ with $\chi(\mathcal{O}_Z(m)) = P(m)$ for all sufficiently large $m$. The Zariski tangent space at $[Z]$ is $\Ext^1_{\mathcal{O}_{\mathbb{P}^n}}(I_Z, \mathcal{O}_Z)$, or equivalently $H^0(\mathcal{N}_{Z/\mathbb{P}^n})$ for smooth locally complete intersection $Z$, which governs first-order infinitesimal deformations of $Z$.[](https://gauss.math.yale.edu/~il282/Benjamin_S16.pdf) The Hilbert scheme is smooth (hence unobstructed) at $[Z]$ if the dimension of this tangent space equals the expected dimension (e.g., $\chi(\mathcal{N}_{Z/\mathbb{P}^n})$ for smooth $Z$) and if the obstruction space $\Ext^2_{\mathcal{O}_{\mathbb{P}^n}}(I_Z, \mathcal{O}_Z)$ (or $H^1(\mathcal{N}_{Z/\mathbb{P}^n})$) vanishes.[](https://gauss.math.yale.edu/~il282/Benjamin_S16.pdf) A prominent example is the Hilbert scheme $\Hilb^n(X)$ of $n$ points on a smooth projective surface $X \subset \mathbb{P}^N$. This scheme is a smooth, irreducible, projective variety of dimension $2n$, providing a desingularization of the symmetric product $X^{(n)}$ via the Hilbert-Chow morphism.[](https://www.jstor.org/stable/2373541) The Hilbert scheme $\Hilb^P(\mathbb{P}^n)$ relates to the Chow variety parametrizing effective cycles of class determined by $P$ through the Hilbert-Chow morphism, which sends each subscheme $Z$ to its associated cycle class $[Z]$ in the Chow group.[](https://link.springer.com/chapter/10.1007/978-3-662-03276-3_2) ### Chow Varieties and Cycles The Chow variety $\operatorname{Chow}^k(\mathbb{P}^n)$ is a projective algebraic variety that parametrizes effective algebraic $k$-cycles of fixed degree $d$ on the projective space $\mathbb{P}^n$ over an algebraically closed field, such as $\mathbb{C}$. More precisely, for each $d \geq 1$, the component $\operatorname{Chow}^{k,d}(\mathbb{P}^n)$ parametrizes effective $k$-dimensional cycles of degree $d$, including irreducible subvarieties with multiplicity one, with the full space obtained by taking the disjoint union over degrees and including multiplicities for non-reduced cycles. This construction relies on Chow forms, which are homogeneous polynomials encoding the geometry of a cycle via its intersections with generic linear subspaces; for a $k$-cycle $\nu = \sum m_i [V_i]$, the Chow form is $F_\nu(u) = \prod F_{V_i}(u)^{m_i}$, where each $F_{V_i}$ is of degree $\deg(V_i)$. The variety is projective and of finite type. The points of the Chow variety correspond to effective cycles, and there is a natural cycle class map from these cycles to the Chow groups $A^k(\mathbb{P}^n)$, which are the free abelian groups generated by irreducible $k$-dimensional subvarieties modulo rational equivalence. Rational equivalence identifies two cycles if their difference is the boundary of a rational family of cycles, formally defined via divisors on curves: a cycle $\partial f = \sum_{\eta \in \mathbb{P}^1} (f_* [Z_\eta] - [Z_{\eta_\infty}])$ for a rational map $f: Z \to \mathbb{P}^n$ from a variety $Z$ with a distinguished point at infinity. This quotient yields $A^k(\mathbb{P}^n) \cong \mathbb{Z}$, generated by the class of a linear subspace of dimension $k$, reflecting the rigidity of projective space. The map preserves degrees and induces the structure on intersection products in the Chow ring. In [intersection theory](/page/Intersection_theory), the Chow variety facilitates basic computations via the moving lemma, which asserts that any two cycles on a projective variety can be deformed—while preserving their classes in the Chow group—into proper intersection position with a given cycle, ensuring transverse intersections generically. This lemma underpins degree computations, such as [Bézout's theorem](/page/Bézout's_theorem): the degree of the intersection of two cycles of complementary dimensions in $\mathbb{P}^n$ equals the product of their degrees, computable as the pushforward to a point in $A^0(\operatorname{pt}) \cong \mathbb{Z}$. For example, two curves of degrees $d_1$ and $d_2$ in $\mathbb{P}^3$ intersect in $d_1 d_2$ points, counted with multiplicity. These tools enable [enumerative geometry](/page/Enumerative_geometry) without resolving singularities. The Chow variety relates to the Hilbert scheme $\Hilb^P(\mathbb{P}^n)$ (where $P$ corresponds to $k$-dimensional degree $d$ subschemes), via the Hilbert-Chow morphism, a proper map that resolves the singularities of the Chow variety. For reduced cycles—those supported on reduced subvarieties—this morphism provides a dense [embedding](/page/Embedding) of the open locus of reduced points in the Chow variety into the [Hilbert scheme](/page/Hilbert_scheme), where it is an [isomorphism](/page/Isomorphism) over smooth points; non-reduced cycles map to singular loci. This [embedding](/page/Embedding) highlights the Chow variety's role in studying algebraic equivalence classes within the more refined [Hilbert scheme](/page/Hilbert_scheme).[](https://link.springer.com/chapter/10.1007/978-3-662-03276-3_2) ## Analytic Aspects over the Complex Numbers ### Relation to Kähler Manifolds A complex projective variety $X \subset \mathbb{CP}^n$ over $\mathbb{C}$, when smooth, is a compact complex submanifold that inherits a Kähler structure from the ambient projective space. The Fubini-Study metric on $\mathbb{CP}^n$, defined via the quotient of the standard Hermitian metric on $\mathbb{C}^{n+1}$ by the $\mathbb{C}^*$-action, induces a Hermitian metric on $X$ whose associated Kähler form $\omega_{FS}$ is the restriction of the closed positive (1,1)-form on $\mathbb{CP}^n$. This form is given locally by $\omega = i \partial \bar{\partial} \log \det(1 + |w|^2)$ in homogeneous coordinates, ensuring $X$ is a Kähler manifold with constant holomorphic sectional curvature inherited from the ambient space.[](https://people.maths.ox.ac.uk/joyce/Nairobi2019/BallmannKahlerManifolds.pdf) Positive line bundles on such varieties play a central role in connecting algebraic and [analytic geometry](/page/Analytic_geometry) through their correspondence to Kähler forms. A holomorphic [line bundle](/page/Line_bundle) $L$ on $X$ is positive if it admits a Hermitian metric $h$ whose curvature form $\frac{i}{2\pi} \Theta_h(L)$ is a positive definite Kähler form, representing the first [Chern class](/page/Chern_class) $c_1(L) > 0$. In pluripotential theory, this positivity relates to the existence of plurisubharmonic potentials whose Monge-Ampère measures encode volume forms compatible with the Kähler structure; for ample $L = \mathcal{O}(d)|_X$ with $d > 0$, high powers $L^{\otimes k}$ yield sections generating embeddings, as per Kodaira's criterion. This framework underpins vanishing theorems and embedding results, distinguishing projective varieties among compact Kähler manifolds.[](https://people.maths.ox.ac.uk/~joyce/KahlerGeom2022/KG9%2B10.pdf)[](http://mcs.unife.it/alex.massarenti/files/Kodaira.pdf) The Hard Lefschetz theorem manifests differently in algebraic and analytic contexts for these varieties, yet aligns via [Hodge theory](/page/Hodge_theory). In the analytic setting, for a [Kähler manifold](/page/Kähler_manifold) $X$ of dimension $n$ with Kähler class $[\omega]$, the operator $L: H^{k}(X, \mathbb{C}) \to H^{k+2}(X, \mathbb{C})$ given by wedging with $[\omega]$ induces isomorphisms $L^{n-k}: H^{k}(X, \mathbb{C}) \to H^{2n-k}(X, \mathbb{C})$ for $k \leq n$, polarizing primitive cohomology via the [Hodge-Riemann bilinear form](/page/Bilinear_form). The algebraic Hard Lefschetz theorem provides an analogous statement: for a smooth projective variety $X$ of dimension $n$ and [ample line bundle](/page/Ample_line_bundle) $L$, the Lefschetz operator $L: H^k(X, \mathbb{C}) \to H^{k+2}(X, \mathbb{C})$ given by [cup product](/page/Cup_product) with $c_1(L)$ induces isomorphisms $L^{n-k}: H^k(X, \mathbb{C}) \to H^{2n-k}(X, \mathbb{C})$ for $k \leq n$, aligning with the analytic version via [Hodge theory](/page/Hodge_theory) and [GAGA](/page/Gaga) principles.[](https://www.numdam.org/article/ASENS_2002_4_35_5_759_0.pdf) Hodge structures on the cohomology of projective varieties incorporate periods and the transcendental lattice to capture transcendental aspects beyond algebraic cycles. The period map sends families of varieties to the [classifying space](/page/Classifying_space) of polarized Hodge structures, where periods are integrals $\int_\gamma \omega$ over cycles $\gamma \in H_k(X, \mathbb{Z})$ and holomorphic forms $\omega \in H^{p,q}(X)$, parametrizing the Hodge filtration via the Gauss-Manin connection. The transcendental lattice $T(X) \subset H^k(X, \mathbb{Z})$ is the [orthogonal complement](/page/Orthogonal_complement) to the algebraically trivial classes ([NS](/page/NS) lattice for $k=2$), forming a sublattice whose [Hodge structure](/page/Hodge_structure) encodes non-algebraic invariants; for example, in K3 surfaces, it determines the moduli via the [period](/page/Period) domain, preserving integrality under [monodromy](/page/Monodromy).[](https://cmsa.fas.harvard.edu/media/Voisin_HSTAVharvardrevision.pdf) ### GAGA Principles and Chow's Theorem The GAGA principles, established by Jean-Pierre Serre, provide a foundational bridge between algebraic geometry and complex analytic geometry for projective varieties over the complex numbers. Specifically, these principles assert an equivalence between the categories of coherent algebraic sheaves and coherent analytic sheaves on a projective algebraic variety $X \subset \mathbb{P}^n$ and its associated analytic space $X_h$. Under this equivalence, the natural map from global sections $\Gamma(X, \mathcal{F})$ to $\Gamma(X_h, \mathcal{F}_h)$ is an isomorphism for any coherent algebraic sheaf $\mathcal{F}$ on $X$, and cohomology groups satisfy $H^q(X, \mathcal{F}) \cong H^q(X_h, \mathcal{F}_h)$ for all $q \geq 0$.[](http://faculty.bicmr.pku.edu.cn/~qizheng/files/hodge/GAGA_en.pdf) Furthermore, every coherent analytic sheaf on $X_h$ is the analytification of a unique coherent algebraic sheaf on $X$, and morphisms between analytified sheaves lift uniquely to algebraic morphisms.[](http://faculty.bicmr.pku.edu.cn/~qizheng/files/hodge/GAGA_en.pdf) A key consequence of the [GAGA](/page/Gaga) principles is an analytic analogue of [Hilbert's Nullstellensatz](/page/Hilbert's_Nullstellensatz). In the algebraic setting, [Hilbert's Nullstellensatz](/page/Hilbert's_Nullstellensatz) describes the radical of an ideal in terms of the zero set of its generators in affine or [projective space](/page/Projective_space). Analytically, for a coherent ideal sheaf $\mathcal{I}$ on the analytification $X_h$ of a projective variety $X$, the zero locus $Z(\mathcal{I})$ coincides with the zero locus of the associated algebraic ideal, ensuring that zeros of coherent analytic sections define algebraic subvarieties.[](http://faculty.bicmr.pku.edu.cn/~qizheng/files/hodge/GAGA_en.pdf) This correspondence implies that coherent analytic ideals on projective embeddings are generated by algebraic polynomials, mirroring the algebraic Nullstellensatz structure.[](http://faculty.bicmr.pku.edu.cn/~qizheng/files/hodge/GAGA_en.pdf) Chow's theorem extends this analytic-algebraic correspondence to subsets of analytic spaces. In particular, for closed analytic subsets of [projective space](/page/Projective_space) $\mathbb{P}^n$, Chow's [theorem](/page/Theorem) asserts that they are precisely the algebraic subvarieties defined by homogeneous ideals, as the preimage under the quotient map from $\mathbb{C}^{n+1} \setminus \{0\}$ yields a cone whose ideal is finitely generated by homogeneous polynomials.[](https://www.math.stonybrook.edu/~cschnell/mat545/class26.pdf)[](https://scispace.com/pdf/on-compact-complex-analytic-varieties-2t5677vqez.pdf) These principles and theorems have significant applications in uniformization and embedding problems for projective varieties. The GAGA correspondence enables uniformization results by ensuring that analytic uniformizations of projective varieties admit algebraic counterparts, preserving the projective structure.[](http://faculty.bicmr.pku.edu.cn/~qizheng/files/hodge/GAGA_en.pdf) Similarly, embedding theorems benefit, as analytic embeddings into [projective space](/page/Projective_space) can be algebraized, allowing compact complex manifolds satisfying certain [coherence](/page/Coherence) conditions to be realized as projective algebraic varieties.[](http://faculty.bicmr.pku.edu.cn/~qizheng/files/hodge/GAGA_en.pdf) ### Complex Tori versus Abelian Varieties A [complex torus](/page/Complex_torus) of [dimension](/page/Dimension) $g$ is defined as the [quotient](/page/Quotient) [space](/page/Space) $\mathbb{C}^g / \Lambda$, where $\Lambda \subset \mathbb{C}^g$ is a [discrete](/page/Discrete) [subgroup](/page/Subgroup) isomorphic to $\mathbb{Z}^{2g}$, known as a [lattice](/page/Lattice).[](https://people.reed.edu/~jerry/311/tori.pdf) This construction endows the torus with the structure of a compact [complex](/page/Complex) [Lie group](/page/Lie_group), inheriting a flat Kähler [metric](/page/Metric) from the standard [Euclidean](/page/Euclidean) structure on $\mathbb{C}^g$.[](https://people.reed.edu/~jerry/311/tori.pdf) Unlike [projective varieties](/page/Projective_variety), [complex tori](/page/Complex_torus) are not necessarily algebraic or embeddable into [projective space](/page/Projective_space), as their complex structure is parameterized by the period matrix of the [lattice](/page/Lattice), which may not satisfy conditions for projectivity.[](https://math.uchicago.edu/~amathew/287y.pdf) The projectivity of a complex torus hinges on the existence of an ample line bundle. A line bundle on the torus admits an ample structure if and only if its associated Riemann form—a Hermitian form on $\mathbb{C}^g$ that is integer-valued on the lattice—has positive definite imaginary part.[](https://wstein.org/edu/Fall2003/252/lectures/10-08-03/10-08-03-Riemann_form.pdf) Equivalently, the period matrix $\Omega$, whose columns form a basis for the lattice together with the standard basis, must lie in the Siegel upper half-space, where the imaginary part $\operatorname{Im} \Omega$ is positive definite; this ensures the torus carries a Kähler metric compatible with an embedding into projective space.[](https://math.uchicago.edu/~amathew/287y.pdf) Without this positivity condition, the torus remains a purely analytic object without algebraic structure. Poincaré's theorem establishes that a [complex torus](/page/Complex_torus) is projective—and thus an [abelian variety](/page/Abelian_variety)—precisely when it admits a nondegenerate Riemann form with positive definite imaginary part.[](https://www.math.stonybrook.edu/~cschnell/pdf/notes/abelian-varieties.pdf) This criterion links the [analytic geometry](/page/Analytic_geometry) of tori to [algebraic geometry](/page/Algebraic_geometry), showing that only those tori with a suitable [polarization](/page/Polarization) can be realized as projective varieties over $\mathbb{C}$.[](https://wstein.org/edu/Fall2003/252/lectures/10-08-03/10-08-03-Riemann_form.pdf) In dimension $g=1$, every complex torus is projective and isomorphic to an elliptic curve, as the moduli space reduces to the upper half-plane, where all points yield elliptic curves via the Weierstrass embedding.[](https://people.dm.unipi.it/lombardo/Teaching/VarietaAbeliane1718/Notes.pdf) However, for $g \geq 2$, most complex tori are non-projective; for instance, in dimension 2, the 3-dimensional moduli space of tori includes regions where no positive definite Riemann form exists, yielding non-algebraic examples like certain quotients $\mathbb{C}^2 / \Lambda$ with period matrices outside the principally polarized locus.[](https://www.math.stonybrook.edu/~sunscorch/examples/Complex_Geo_Examples.pdf) These non-algebraic tori highlight the distinction between analytic complex structures and algebraic varieties in higher dimensions.