Ample line bundle

Last updated

In mathematics, a distinctive feature of algebraic geometry is that some line bundles on a projective variety can be considered "positive", while others are "negative" (or a mixture of the two). The most important notion of positivity is that of an ample line bundle, although there are several related classes of line bundles. Roughly speaking, positivity properties of a line bundle are related to having many global sections. Understanding the ample line bundles on a given variety X amounts to understanding the different ways of mapping X into projective space. In view of the correspondence between line bundles and divisors (built from codimension-1 subvarieties), there is an equivalent notion of an ample divisor.

Contents

In more detail, a line bundle is called basepoint-free if it has enough sections to give a morphism to projective space. A line bundle is semi-ample if some positive power of it is basepoint-free; semi-ampleness is a kind of "nonnegativity". More strongly, a line bundle on a complete variety X is very ample if it has enough sections to give a closed immersion (or "embedding") of X into projective space. A line bundle is ample if some positive power is very ample.

An ample line bundle on a projective variety X has positive degree on every curve in X. The converse is not quite true, but there are corrected versions of the converse, the Nakai–Moishezon and Kleiman criteria for ampleness.

Introduction

Pullback of a line bundle and hyperplane divisors

Given a morphism of schemes, a vector bundle E on Y (or more generally a coherent sheaf on Y) has a pullback to X, (see Sheaf of modules#Operations). The pullback of a vector bundle is a vector bundle of the same rank. In particular, the pullback of a line bundle is a line bundle. (Briefly, the fiber of at a point x in X is the fiber of E at f(x).)

The notions described in this article are related to this construction in the case of a morphism to projective space

with E = O(1) the line bundle on projective space whose global sections are the homogeneous polynomials of degree 1 (that is, linear functions) in variables . The line bundle O(1) can also be described as the line bundle associated to a hyperplane in (because the zero set of a section of O(1) is a hyperplane). If f is a closed immersion, for example, it follows that the pullback is the line bundle on X associated to a hyperplane section (the intersection of X with a hyperplane in ).

Basepoint-free line bundles

Let X be a scheme over a field k (for example, an algebraic variety) with a line bundle L. (A line bundle may also be called an invertible sheaf.) Let be elements of the k-vector space of global sections of L. The zero set of each section is a closed subset of X; let U be the open subset of points at which at least one of is not zero. Then these sections define a morphism

In more detail: for each point x of U, the fiber of L over x is a 1-dimensional vector space over the residue field k(x). Choosing a basis for this fiber makes into a sequence of n+1 numbers, not all zero, and hence a point in projective space. Changing the choice of basis scales all the numbers by the same nonzero constant, and so the point in projective space is independent of the choice.

Moreover, this morphism has the property that the restriction of L to U is isomorphic to the pullback . [1]

The base locus of a line bundle L on a scheme X is the intersection of the zero sets of all global sections of L. A line bundle L is called basepoint-free if its base locus is empty. That is, for every point x of X there is a global section of L which is nonzero at x. If X is proper over a field k, then the vector space of global sections has finite dimension; the dimension is called . [2] So a basepoint-free line bundle L determines a morphism over k, where , given by choosing a basis for . Without making a choice, this can be described as the morphism

from X to the space of hyperplanes in , canonically associated to the basepoint-free line bundle L. This morphism has the property that L is the pullback .

Conversely, for any morphism f from a scheme X to projective space over k, the pullback line bundle is basepoint-free. Indeed, O(1) is basepoint-free on , because for every point y in there is a hyperplane not containing y. Therefore, for every point x in X, there is a section s of O(1) over that is not zero at f(x), and the pullback of s is a global section of that is not zero at x. In short, basepoint-free line bundles are exactly those that can be expressed as the pullback of O(1) by some morphism to projective space.

Nef, globally generated, semi-ample

The degree of a line bundle L on a proper curve C over k is defined as the degree of the divisor (s) of any nonzero rational section s of L. The coefficients of this divisor are positive at points where s vanishes and negative where s has a pole. Therefore, any line bundle L on a curve C such that has nonnegative degree (because sections of L over C, as opposed to rational sections, have no poles). [3] In particular, every basepoint-free line bundle on a curve has nonnegative degree. As a result, a basepoint-free line bundle L on any proper scheme X over a field is nef , meaning that L has nonnegative degree on every (irreducible) curve in X. [4]

More generally, a sheaf F of -modules on a scheme X is said to be globally generated if there is a set I of global sections such that the corresponding morphism

of sheaves is surjective. [5] A line bundle is globally generated if and only if it is basepoint-free.

For example, every quasi-coherent sheaf on an affine scheme is globally generated. [6] Analogously, in complex geometry, Cartan's theorem A says that every coherent sheaf on a Stein manifold is globally generated.

A line bundle L on a proper scheme over a field is semi-ample if there is a positive integer r such that the tensor power is basepoint-free. A semi-ample line bundle is nef (by the corresponding fact for basepoint-free line bundles). [7]

Very ample line bundles

A line bundle L on a proper scheme X over a field k is said to be very ample if it is basepoint-free and the associated morphism

is a closed immersion. Here . Equivalently, L is very ample if X can be embedded into projective space of some dimension over k in such a way that L is the restriction of the line bundle O(1) to X. [8] The latter definition is used to define very ampleness for a line bundle on a proper scheme over any commutative ring. [9]

The name "very ample" was introduced by Alexander Grothendieck in 1961. [10] Various names had been used earlier in the context of linear systems of divisors.

For a very ample line bundle L on a proper scheme X over a field with associated morphism f, the degree of L on a curve C in X is the degree of f(C) as a curve in . So L has positive degree on every curve in X (because every subvariety of projective space has positive degree). [11]

Definitions

Ample invertible sheaves on quasi-compact schemes

Ample line bundles are used most often on proper schemes, but they can be defined in much wider generality.

Let X be a scheme, and let be an invertible sheaf on X. For each , let denote the ideal sheaf of the reduced subscheme supported only at x. For , define

Equivalently, if denotes the residue field at x (considered as a skyscraper sheaf supported at x), then

where is the image of s in the tensor product.

Fix . For every s, the restriction is a free -module trivialized by the restriction of s, meaning the multiplication-by-s morphism is an isomorphism. The set is always open, and the inclusion morphism is an affine morphism. Despite this, need not be an affine scheme. For example, if , then is open in itself and affine over itself but generally not affine.

Assume X is quasi-compact. Then is ample if, for every , there exists an and an such that and is an affine scheme. [12] For example, the trivial line bundle is ample if and only if X is quasi-affine. [13]

In general, it is not true that every is affine. For example, if for some point O, and if is the restriction of to X, then and have the same global sections, and the non-vanishing locus of a section of is affine if and only if the corresponding section of contains O.

It is necessary to allow powers of in the definition. In fact, for every N, it is possible that is non-affine for every with . Indeed, suppose Z is a finite set of points in , , and . The vanishing loci of the sections of are plane curves of degree N. By taking Z to be a sufficiently large set of points in general position, we may ensure that no plane curve of degree N (and hence any lower degree) contains all the points of Z. In particular their non-vanishing loci are all non-affine.

Define . Let denote the structural morphism. There is a natural isomorphism between -algebra homomorphisms and endomorphisms of the graded ring S. The identity endomorphism of S corresponds to a homomorphism . Applying the functor produces a morphism from an open subscheme of X, denoted , to .

The basic characterization of ample invertible sheaves states that if X is a quasi-compact quasi-separated scheme and is an invertible sheaf on X, then the following assertions are equivalent: [14]

  1. is ample.
  2. The open sets , where and , form a basis for the topology of X.
  3. The open sets with the property of being affine, where and , form a basis for the topology of X.
  4. and the morphism is a dominant open immersion.
  5. and the morphism is a homeomorphism of the underlying topological space of X with its image.
  6. For every quasi-coherent sheaf on X, the canonical map is surjective.
  7. For every quasi-coherent sheaf of ideals on X, the canonical map is surjective.
  8. For every quasi-coherent sheaf of ideals on X, the canonical map is surjective.
  9. For every quasi-coherent sheaf of finite type on X, there exists an integer such that for , is generated by its global sections.
  10. For every quasi-coherent sheaf of finite type on X, there exists integers and such that is isomorphic to a quotient of .
  11. For every quasi-coherent sheaf of ideals of finite type on X, there exists integers and such that is isomorphic to a quotient of .

On proper schemes

When X is separated and finite type over an affine scheme, an invertible sheaf is ample if and only if there exists a positive integer r such that the tensor power is very ample. [15] [16] In particular, a proper scheme over R has an ample line bundle if and only if it is projective over R. Often, this characterization is taken as the definition of ampleness.

The rest of this article will concentrate on ampleness on proper schemes over a field, as this is the most important case. An ample line bundle on a proper scheme X over a field has positive degree on every curve in X, by the corresponding statement for very ample line bundles.

A Cartier divisor D on a proper scheme X over a field k is said to be ample if the corresponding line bundle O(D) is ample. (For example, if X is smooth over k, then a Cartier divisor can be identified with a finite linear combination of closed codimension-1 subvarieties of X with integer coefficients.)

Weakening the notion of "very ample" to "ample" gives a flexible concept with a wide variety of different characterizations. A first point is that tensoring high powers of an ample line bundle with any coherent sheaf whatsoever gives a sheaf with many global sections. More precisely, a line bundle L on a proper scheme X over a field (or more generally over a Noetherian ring) is ample if and only if for every coherent sheaf F on X, there is an integer s such that the sheaf is globally generated for all . Here s may depend on F. [17] [18]

Another characterization of ampleness, known as the CartanSerreGrothendieck theorem, is in terms of coherent sheaf cohomology. Namely, a line bundle L on a proper scheme X over a field (or more generally over a Noetherian ring) is ample if and only if for every coherent sheaf F on X, there is an integer s such that

for all and all . [19] [18] In particular, high powers of an ample line bundle kill cohomology in positive degrees. This implication is called the Serre vanishing theorem, proved by Jean-Pierre Serre in his 1955 paper Faisceaux algébriques cohérents.

Examples/Non-examples

by
This is a closed immersion for , with image a rational normal curve of degree d in . Therefore, O(d) is basepoint-free if and only if , and very ample if and only if . It follows that O(d) is ample if and only if .

Criteria for ampleness of line bundles

Intersection theory

To determine whether a given line bundle on a projective variety X is ample, the following numerical criteria (in terms of intersection numbers) are often the most useful. It is equivalent to ask when a Cartier divisor D on X is ample, meaning that the associated line bundle O(D) is ample. The intersection number can be defined as the degree of the line bundle O(D) restricted to C. In the other direction, for a line bundle L on a projective variety, the first Chern class means the associated Cartier divisor (defined up to linear equivalence), the divisor of any nonzero rational section of L.

On a smooth projective curve X over an algebraically closed field k, a line bundle L is very ample if and only if for all k-rational points x,y in X. [23] Let g be the genus of X. By the Riemann–Roch theorem, every line bundle of degree at least 2g + 1 satisfies this condition and hence is very ample. As a result, a line bundle on a curve is ample if and only if it has positive degree. [24]

For example, the canonical bundle of a curve X has degree 2g  2, and so it is ample if and only if . The curves with ample canonical bundle form an important class; for example, over the complex numbers, these are the curves with a metric of negative curvature. The canonical bundle is very ample if and only if and the curve is not hyperelliptic. [25]

The Nakai–Moishezon criterion (named for Yoshikazu Nakai (1963) and Boris Moishezon (1964)) states that a line bundle L on a proper scheme X over a field is ample if and only if for every (irreducible) closed subvariety Y of X (Y is not allowed to be a point). [26] In terms of divisors, a Cartier divisor D is ample if and only if for every (nonzero-dimensional) subvariety Y of X. For X a curve, this says that a divisor is ample if and only if it has positive degree. For X a surface, the criterion says that a divisor D is ample if and only if its self-intersection number is positive and every curve C on X has .

Kleiman's criterion

To state Kleiman's criterion (1966), let X be a projective scheme over a field. Let be the real vector space of 1-cycles (real linear combinations of curves in X) modulo numerical equivalence, meaning that two 1-cycles A and B are equal in if and only if every line bundle has the same degree on A and on B. By the Néron–Severi theorem, the real vector space has finite dimension. Kleiman's criterion states that a line bundle L on X is ample if and only if L has positive degree on every nonzero element C of the closure of the cone of curves NE(X) in . (This is slightly stronger than saying that L has positive degree on every curve.) Equivalently, a line bundle is ample if and only if its class in the dual vector space is in the interior of the nef cone. [27]

Kleiman's criterion fails in general for proper (rather than projective) schemes X over a field, although it holds if X is smooth or more generally Q-factorial. [28]

A line bundle on a projective variety is called strictly nef if it has positive degree on every curve Nagata (1959). and David Mumford constructed line bundles on smooth projective surfaces that are strictly nef but not ample. This shows that the condition cannot be omitted in the Nakai–Moishezon criterion, and it is necessary to use the closure of NE(X) rather than NE(X) in Kleiman's criterion. [29] Every nef line bundle on a surface has , and Nagata and Mumford's examples have .

C. S. Seshadri showed that a line bundle L on a proper scheme over an algebraically closed field is ample if and only if there is a positive real number ε such that deg(L|C) ≥ εm(C) for all (irreducible) curves C in X, where m(C) is the maximum of the multiplicities at the points of C. [30]

Several characterizations of ampleness hold more generally for line bundles on a proper algebraic space over a field k. In particular, the Nakai-Moishezon criterion is valid in that generality. [31] The Cartan-Serre-Grothendieck criterion holds even more generally, for a proper algebraic space over a Noetherian ring R. [32] (If a proper algebraic space over R has an ample line bundle, then it is in fact a projective scheme over R.) Kleiman's criterion fails for proper algebraic spaces X over a field, even if X is smooth. [33]

Openness of ampleness

On a projective scheme X over a field, Kleiman's criterion implies that ampleness is an open condition on the class of an R-divisor (an R-linear combination of Cartier divisors) in , with its topology based on the topology of the real numbers. (An R-divisor is defined to be ample if it can be written as a positive linear combination of ample Cartier divisors. [34] ) An elementary special case is: for an ample divisor H and any divisor E, there is a positive real number b such that is ample for all real numbers a of absolute value less than b. In terms of divisors with integer coefficients (or line bundles), this means that nH + E is ample for all sufficiently large positive integers n.

Ampleness is also an open condition in a quite different sense, when the variety or line bundle is varied in an algebraic family. Namely, let be a proper morphism of schemes, and let L be a line bundle on X. Then the set of points y in Y such that L is ample on the fiber is open (in the Zariski topology). More strongly, if L is ample on one fiber , then there is an affine open neighborhood U of y such that L is ample on over U. [35]

Kleiman's other characterizations of ampleness

Kleiman also proved the following characterizations of ampleness, which can be viewed as intermediate steps between the definition of ampleness and numerical criteria. Namely, for a line bundle L on a proper scheme X over a field, the following are equivalent: [36]

as .

Generalizations

Ample vector bundles

Robin Hartshorne defined a vector bundle F on a projective scheme X over a field to be ample if the line bundle on the space of hyperplanes in F is ample. [37]

Several properties of ample line bundles extend to ample vector bundles. For example, a vector bundle F is ample if and only if high symmetric powers of F kill the cohomology of coherent sheaves for all . [38] Also, the Chern class of an ample vector bundle has positive degree on every r-dimensional subvariety of X, for . [39]

Big line bundles

A useful weakening of ampleness, notably in birational geometry, is the notion of a big line bundle. A line bundle L on a projective variety X of dimension n over a field is said to be big if there is a positive real number a and a positive integer such that for all . This is the maximum possible growth rate for the spaces of sections of powers of L, in the sense that for every line bundle L on X there is a positive number b with for all j > 0. [40]

There are several other characterizations of big line bundles. First, a line bundle is big if and only if there is a positive integer r such that the rational map from X to given by the sections of is birational onto its image. [41] Also, a line bundle L is big if and only if it has a positive tensor power which is the tensor product of an ample line bundle A and an effective line bundle B (meaning that ). [42] Finally, a line bundle is big if and only if its class in is in the interior of the cone of effective divisors. [43]

Bigness can be viewed as a birationally invariant analog of ampleness. For example, if is a dominant rational map between smooth projective varieties of the same dimension, then the pullback of a big line bundle on Y is big on X. (At first sight, the pullback is only a line bundle on the open subset of X where f is a morphism, but this extends uniquely to a line bundle on all of X.) For ample line bundles, one can only say that the pullback of an ample line bundle by a finite morphism is ample. [20]

Example: Let X be the blow-up of the projective plane at a point over the complex numbers. Let H be the pullback to X of a line on , and let E be the exceptional curve of the blow-up . Then the divisor H + E is big but not ample (or even nef) on X, because

This negativity also implies that the base locus of H + E (or of any positive multiple) contains the curve E. In fact, this base locus is equal to E.

Relative ampleness

Given a quasi-compact morphism of schemes , an invertible sheaf L on X is said to be ample relative to f or f-ample if the following equivalent conditions are met: [44] [45]

  1. For each open affine subset , the restriction of L to is ample (in the usual sense).
  2. f is quasi-separated and there is an open immersion induced by the adjunction map:
    .
  3. The condition 2. without "open".

The condition 2 says (roughly) that X can be openly compactified to a projective scheme with (not just to a proper scheme).

See also

General algebraic geometry

Ampleness in complex geometry

Notes

  1. Hartshorne (1977), Theorem II.7.1.
  2. Hartshorne (1977), Theorem III.5.2; ( tag 02O6 ).
  3. Hartshorne (1977), Lemma IV.1.2.
  4. Lazarsfeld (2004), Example 1.4.5.
  5. tag 01AM.
  6. Hartshorne (1977), Example II.5.16.2.
  7. Lazarsfeld (2004), Definition 2.1.26.
  8. Hartshorne (1977), section II.5.
  9. tag 02NP.
  10. Grothendieck, EGA II, Definition 4.2.2.
  11. Hartshorne (1977), Proposition I.7.6 and Example IV.3.3.2.
  12. tag 01PS.
  13. tag 01QE.
  14. EGA II, Théorème 4.5.2 and Proposition 4.5.5.
  15. EGA II, Proposition 4.5.10.
  16. tag 01VU.
  17. Hartshorne (1977), Theorem II.7.6
  18. 1 2 Lazarsfeld (2004), Theorem 1.2.6.
  19. Hartshorne (1977), Proposition III.5.3
  20. 1 2 Lazarsfeld (2004), Theorem 1.2.13.
  21. Hartshorne (1977), Example II.7.6.3.
  22. Hartshorne (1977), Exercise IV.3.2(b).
  23. Hartshorne (1977), Proposition IV.3.1.
  24. Hartshorne (1977), Corollary IV.3.3.
  25. Hartshorne (1977), Proposition IV.5.2.
  26. Lazarsfeld (2004), Theorem 1.2.23, Remark 1.2.29; Kleiman (1966), Theorem III.1.
  27. Lazarsfeld (2004), Theorems 1.4.23 and 1.4.29; Kleiman (1966), Theorem IV.1.
  28. Fujino (2005), Corollary 3.3; Lazarsfeld (2004), Remark 1.4.24.
  29. Lazarsfeld (2004), Example 1.5.2.
  30. Lazarsfeld (2004), Theorem 1.4.13; Hartshorne (1970), Theorem I.7.1.
  31. Kollár (1990), Theorem 3.11.
  32. tag 0D38.
  33. Kollár (1996), Chapter VI, Appendix, Exercise 2.19.3.
  34. Lazarsfeld (2004), Definition 1.3.11.
  35. Lazarsfeld (2004), Theorem 1.2.17 and its proof.
  36. Lazarsfeld (2004), Example 1.2.32; Kleiman (1966), Theorem III.1.
  37. Lazarsfeld (2004), Definition 6.1.1.
  38. Lazarsfeld (2004), Theorem 6.1.10.
  39. Lazarsfeld (2004), Theorem 8.2.2.
  40. Lazarsfeld (2004), Corollary 2.1.38.
  41. Lazarsfeld (2004), section 2.2.A.
  42. Lazarsfeld (2004), Corollary 2.2.7.
  43. Lazarsfeld (2004), Theorem 2.2.26.
  44. tag 01VG.
  45. Grothendieck & Dieudonné 1961, Proposition 4.6.3.

Sources

Related Research Articles

The Riemann–Roch theorem is an important theorem in mathematics, specifically in complex analysis and algebraic geometry, for the computation of the dimension of the space of meromorphic functions with prescribed zeros and allowed poles. It relates the complex analysis of a connected compact Riemann surface with the surface's purely topological genus g, in a way that can be carried over into purely algebraic settings.

<span class="mw-page-title-main">Projective variety</span>

In algebraic geometry, a projective variety over an algebraically closed field k is a subset of some projective n-space over k that is the zero-locus of some finite family of homogeneous polynomials of n + 1 variables with coefficients in k, that generate a prime ideal, the defining ideal of the variety. Equivalently, an algebraic variety is projective if it can be embedded as a Zariski closed subvariety of .

In algebraic geometry, a branch of mathematics, Serre duality is a duality for the coherent sheaf cohomology of algebraic varieties, proved by Jean-Pierre Serre. The basic version applies to vector bundles on a smooth projective variety, but Alexander Grothendieck found wide generalizations, for example to singular varieties. On an n-dimensional variety, the theorem says that a cohomology group is the dual space of another one, . Serre duality is the analog for coherent sheaf cohomology of Poincaré duality in topology, with the canonical line bundle replacing the orientation sheaf.

In mathematics, especially in algebraic geometry and the theory of complex manifolds, coherent sheaves are a class of sheaves closely linked to the geometric properties of the underlying space. The definition of coherent sheaves is made with reference to a sheaf of rings that codifies this geometric information.

In mathematics, Kähler differentials provide an adaptation of differential forms to arbitrary commutative rings or schemes. The notion was introduced by Erich Kähler in the 1930s. It was adopted as standard in commutative algebra and algebraic geometry somewhat later, once the need was felt to adapt methods from calculus and geometry over the complex numbers to contexts where such methods are not available.

<span class="mw-page-title-main">Linear system of divisors</span>

In algebraic geometry, a linear system of divisors is an algebraic generalization of the geometric notion of a family of curves; the dimension of the linear system corresponds to the number of parameters of the family.

<span class="mw-page-title-main">Grothendieck–Riemann–Roch theorem</span>

In mathematics, specifically in algebraic geometry, the Grothendieck–Riemann–Roch theorem is a far-reaching result on coherent cohomology. It is a generalisation of the Hirzebruch–Riemann–Roch theorem, about complex manifolds, which is itself a generalisation of the classical Riemann–Roch theorem for line bundles on compact Riemann surfaces.

In algebraic geometry, divisors are a generalization of codimension-1 subvarieties of algebraic varieties. Two different generalizations are in common use, Cartier divisors and Weil divisors. Both are derived from the notion of divisibility in the integers and algebraic number fields.

In mathematics, coherent duality is any of a number of generalisations of Serre duality, applying to coherent sheaves, in algebraic geometry and complex manifold theory, as well as some aspects of commutative algebra that are part of the 'local' theory.

In algebraic geometry, Proj is a construction analogous to the spectrum-of-a-ring construction of affine schemes, which produces objects with the typical properties of projective spaces and projective varieties. The construction, while not functorial, is a fundamental tool in scheme theory.

In mathematics, the Hirzebruch–Riemann–Roch theorem, named after Friedrich Hirzebruch, Bernhard Riemann, and Gustav Roch, is Hirzebruch's 1954 result generalizing the classical Riemann–Roch theorem on Riemann surfaces to all complex algebraic varieties of higher dimensions. The result paved the way for the Grothendieck–Hirzebruch–Riemann–Roch theorem proved about three years later.

In mathematics, especially in algebraic geometry and the theory of complex manifolds, the adjunction formula relates the canonical bundle of a variety and a hypersurface inside that variety. It is often used to deduce facts about varieties embedded in well-behaved spaces such as projective space or to prove theorems by induction.

In algebraic geometry, a line bundle on a projective variety is nef if it has nonnegative degree on every curve in the variety. The classes of nef line bundles are described by a convex cone, and the possible contractions of the variety correspond to certain faces of the nef cone. In view of the correspondence between line bundles and divisors, there is an equivalent notion of a nef divisor.

In mathematics, especially in algebraic geometry and the theory of complex manifolds, coherent sheaf cohomology is a technique for producing functions with specified properties. Many geometric questions can be formulated as questions about the existence of sections of line bundles or of more general coherent sheaves; such sections can be viewed as generalized functions. Cohomology provides computable tools for producing sections, or explaining why they do not exist. It also provides invariants to distinguish one algebraic variety from another.

The concept of a Projective space plays a central role in algebraic geometry. This article aims to define the notion in terms of abstract algebraic geometry and to describe some basic uses of projective spaces.

This is a glossary of algebraic geometry.

In algebraic geometry, the Quot scheme is a scheme parametrizing sheaves on a projective scheme. More specifically, if X is a projective scheme over a Noetherian scheme S and if F is a coherent sheaf on X, then there is a scheme whose set of T-points is the set of isomorphism classes of the quotients of that are flat over T. The notion was introduced by Alexander Grothendieck.

In mathematics, a sheaf of O-modules or simply an O-module over a ringed space (X, O) is a sheaf F such that, for any open subset U of X, F(U) is an O(U)-module and the restriction maps F(U) → F(V) are compatible with the restriction maps O(U) → O(V): the restriction of fs is the restriction of f times that of s for any f in O(U) and s in F(U).

In algebraic geometry, the dualizing sheaf on a proper scheme X of dimension n over a field k is a coherent sheaf together with a linear functional

In mathematics, derived noncommutative algebraic geometry, the derived version of noncommutative algebraic geometry, is the geometric study of derived categories and related constructions of triangulated categories using categorical tools. Some basic examples include the bounded derived category of coherent sheaves on a smooth variety, , called its derived category, or the derived category of perfect complexes on an algebraic variety, denoted . For instance, the derived category of coherent sheaves on a smooth projective variety can be used as an invariant of the underlying variety for many cases. Unfortunately, studying derived categories as geometric objects of themselves does not have a standardized name.