Base change theorems

Last updated

In mathematics, the base change theorems relate the direct image and the inverse image of sheaves. More precisely, they are about the base change map, given by the following natural transformation of sheaves:

Contents

where

is a Cartesian square of topological spaces and is a sheaf on X.

Such theorems exist in different branches of geometry: for (essentially arbitrary) topological spaces and proper maps f, in algebraic geometry for (quasi-)coherent sheaves and f proper or g flat, similarly in analytic geometry, but also for étale sheaves for f proper or g smooth.

Introduction

A simple base change phenomenon arises in commutative algebra when A is a commutative ring and B and A' are two A-algebras. Let . In this situation, given a B-module M, there is an isomorphism (of A' -modules):

Here the subscript indicates the forgetful functor, i.e., is M, but regarded as an A-module. Indeed, such an isomorphism is obtained by observing

Thus, the two operations, namely forgetful functors and tensor products commute in the sense of the above isomorphism. The base change theorems discussed below are statements of a similar kind.

Definition of the base change map

The base change theorems presented below all assert that (for different types of sheaves, and under various assumptions on the maps involved), that the following base change map

is an isomorphism, where

are continuous maps between topological spaces that form a Cartesian square and is a sheaf on X. [1] Here denotes the higher direct image of under f, i.e., the derived functor of the direct image (also known as pushforward) functor .

This map exists without any assumptions on the maps f and g. It is constructed as follows: since is left adjoint to , there is a natural map (called unit map)

and so

The Grothendieck spectral sequence then gives the first map and the last map (they are edge maps) in:

Combining this with the above yields

Using the adjointness of and finally yields the desired map.

The above-mentioned introductory example is a special case of this, namely for the affine schemes and, consequently, , and the quasi-coherent sheaf associated to the B-module M.

It is conceptually convenient to organize the above base change maps, which only involve only a single higher direct image functor, into one which encodes all at a time. In fact, similar arguments as above yield a map in the derived category of sheaves on S':

where denotes the (total) derived functor of .

General topology

Proper base change

If X is a Hausdorff topological space, S is a locally compact Hausdorff space and f is universally closed (i.e., is a closed map for any continuous map ), then the base change map

is an isomorphism. [2] Indeed, we have: for ,

and so for

To encode all individual higher derived functors of into one entity, the above statement may equivalently be rephrased by saying that the base change map

is a quasi-isomorphism.

The assumptions that the involved spaces be Hausdorff have been weakened by Schnürer & Soergel (2016).

Lurie (2009) has extended the above theorem to non-abelian sheaf cohomology, i.e., sheaves taking values in simplicial sets (as opposed to abelian groups). [3]

Direct image with compact support

If the map f is not closed, the base change map need not be an isomorphism, as the following example shows (the maps are the standard inclusions) :

One the one hand is always zero, but if is a local system on corresponding to a representation of the fundamental group (which is isomorphic to Z), then can be computed as the invariants of the monodromy action of on the stalk (for any ), which need not vanish.

To obtain a base-change result, the functor (or its derived functor) has to be replaced by the direct image with compact support . For example, if is the inclusion of an open subset, such as in the above example, is the extension by zero, i.e., its stalks are given by

In general, there is a map , which is a quasi-isomorphism if f is proper, but not in general. The proper base change theorem mentioned above has the following generalization: there is a quasi-isomorphism [4]

Base change for quasi-coherent sheaves

Proper base change

Proper base change theorems for quasi-coherent sheaves apply in the following situation: is a proper morphism between noetherian schemes, and is a coherent sheaf which is flat over S (i.e., is flat over ). In this situation, the following statements hold: [5]

is an isomorphism for all .
Furthermore, if these conditions hold, then the natural map
is an isomorphism for all .
is an isomorphism for all .

As the stalk of the sheaf is closely related to the cohomology of the fiber of the point under f, this statement is paraphrased by saying that "cohomology commutes with base extension". [6]

These statements are proved using the following fact, where in addition to the above assumptions : there is a finite complex of finitely generated projective A-modules and a natural isomorphism of functors

on the category of -algebras.

Flat base change

The base change map

is an isomorphism for a quasi-coherent sheaf (on ), provided that the map is flat (together with a number of technical conditions: f needs to be a separated morphism of finite type, the schemes involved need to be Noetherian). [7]

Flat base change in the derived category

A far reaching extension of flat base change is possible when considering the base change map

in the derived category of sheaves on S', similarly as mentioned above. Here is the (total) derived functor of the pullback of -modules (because involves a tensor product, is not exact when g is not flat and therefore is not equal to its derived functor ). This map is a quasi-isomorphism provided that the following conditions are satisfied: [8]

  • is quasi-compact and is quasi-compact and quasi-separated,
  • is an object in , the bounded derived category of -modules, and its cohomology sheaves are quasi-coherent (for example, could be a bounded complex of quasi-coherent sheaves)
  • and are Tor-independent over , meaning that if and satisfy , then for all integers ,
.
  • One of the following conditions is satisfied:
    • has finite flat amplitude relative to , meaning that it is quasi-isomorphic in to a complex such that is -flat for all outside some bounded interval ; equivalently, there exists an interval such that for any complex in , one has for all outside ; or
    • has finite Tor-dimension, meaning that has finite flat amplitude relative to .

One advantage of this formulation is that the flatness hypothesis has been weakened. However, making concrete computations of the cohomology of the left- and right-hand sides now requires the Grothendieck spectral sequence.

Base change in derived algebraic geometry

Derived algebraic geometry provides a means to drop the flatness assumption, provided that the pullback is replaced by the homotopy pullback. In the easiest case when X, S, and are affine (with the notation as above), the homotopy pullback is given by the derived tensor product

Then, assuming that the schemes (or, more generally, derived schemes) involved are quasi-compact and quasi-separated, the natural transformation

is a quasi-isomorphism for any quasi-coherent sheaf, or more generally a complex of quasi-coherent sheaves. [9] The afore-mentioned flat base change result is in fact a special case since for g flat the homotopy pullback (which is locally given by a derived tensor product) agrees with the ordinary pullback (locally given by the underived tensor product), and since the pullback along the flat maps g and g' are automatically derived (i.e., ). The auxiliary assumptions related to the Tor-independence or Tor-amplitude in the preceding base change theorem also become unnecessary.

In the above form, base change has been extended by Ben-Zvi, Francis & Nadler (2010) to the situation where X, S, and S' are (possibly derived) stacks, provided that the map f is a perfect map (which includes the case that f is a quasi-compact, quasi-separated map of schemes, but also includes more general stacks, such as the classifying stack BG of an algebraic group in characteristic zero).

Variants and applications

Proper base change also holds in the context of complex manifolds and complex analytic spaces. [10] The theorem on formal functions is a variant of the proper base change, where the pullback is replaced by a completion operation.

The see-saw principle and the theorem of the cube, which are foundational facts in the theory of abelian varieties, are a consequence of proper base change. [11]

A base-change also holds for D-modules: if X, S, X', and S' are smooth varieties (but f and g need not be flat or proper etc.), there is a quasi-isomorphism

where and denote the inverse and direct image functors for D-modules. [12]

Base change for étale sheaves

For étale torsion sheaves , there are two base change results referred to as proper and smooth base change, respectively: base change holds if is proper. [13] It also holds if g is smooth, provided that f is quasi-compact and provided that the torsion of is prime to the characteristic of the residue fields of X. [14]

Closely related to proper base change is the following fact (the two theorems are usually proved simultaneously): let X be a variety over a separably closed field and a constructible sheaf on . Then are finite in each of the following cases:

Under additional assumptions, Deninger (1988) extended the proper base change theorem to non-torsion étale sheaves.

Applications

In close analogy to the topological situation mentioned above, the base change map for an open immersion f,

is not usually an isomorphism. [15] Instead the extension by zero functor satisfies an isomorphism

This fact and the proper base change suggest to define the direct image functor with compact support for a map f by

where is a compactification of f, i.e., a factorization into an open immersion followed by a proper map. The proper base change theorem is needed to show that this is well-defined, i.e., independent (up to isomorphism) of the choice of the compactification. Moreover, again in analogy to the case of sheaves on a topological space, a base change formula for vs. does hold for non-proper maps f.

For the structural map of a scheme over a field k, the individual cohomologies of , denoted by referred to as cohomology with compact support. It is an important variant of usual étale cohomology.

Similar ideas are also used to construct an analogue of the functor in A1-homotopy theory. [16] [17]

See also

Further reading

Notes

  1. The roles of and are symmetric, and in some contexts (especially smooth base change) the more familiar formulation is the other one (dealing instead with the map for a sheaf on ). For consistency, the results in this article below are all stated for the same situation, namely the map ; but readers should be sure to check this against their expectations.
  2. Milne (2012 , Theorem 17.3)
  3. Lurie (2009 , Theorem 7.3.1.16)
  4. Iversen (1986), the four spaces are assumed to be locally compact and of finite dimension.
  5. Grothendieck (1963 , Section 7.7), Hartshorne (1977 , Theorem III.12.11), Vakil (2015 , Chapter 28 Cohomology and base change theorems)
  6. Hartshorne (1977 , p. 255)
  7. Hartshorne (1977 , Proposition III.9.3)
  8. Berthelot, Grothendieck & Illusie (1971 , SGA 6 IV, Proposition 3.1.0)
  9. Toën (2012 , Proposition 1.4)
  10. Grauert (1960)
  11. Mumford (2008)
  12. Hotta, Takeuchi & Tanisaki (2008 , Theorem 1.7.3)
  13. Artin, Grothendieck & Verdier (1972 , Exposé XII), Milne (1980 , section VI.2)
  14. Artin, Grothendieck & Verdier (1972 , Exposé XVI)
  15. Milne (2012 , Example 8.5)
  16. Ayoub, Joseph (2007), Les six opérations de Grothendieck et le formalisme des cycles évanescents dans le monde motivique. I., Société Mathématique de France, ISBN   978-2-85629-244-0, Zbl   1146.14001
  17. Cisinski, Denis-Charles; Déglise, Frédéric (2019), Triangulated Categories of Mixed Motives, Springer Monographs in Mathematics, arXiv: 0912.2110 , Bibcode:2009arXiv0912.2110C, doi:10.1007/978-3-030-33242-6, ISBN   978-3-030-33241-9, S2CID   115163824

Related Research Articles

In mathematics, a sheaf is a tool for systematically tracking data attached to the open sets of a topological space and defined locally with regard to them. For example, for each open set, the data could be the ring of continuous functions defined on that open set. Such data is well behaved in that it can be restricted to smaller open sets, and also the data assigned to an open set is equivalent to all collections of compatible data assigned to collections of smaller open sets covering the original open set.

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.

In mathematics, algebraic geometry and analytic geometry are two closely related subjects. While algebraic geometry studies algebraic varieties, analytic geometry deals with complex manifolds and the more general analytic spaces defined locally by the vanishing of analytic functions of several complex variables. The deep relation between these subjects has numerous applications in which algebraic techniques are applied to analytic spaces and analytic techniques to algebraic varieties.

In mathematics, the derived categoryD(A) of an abelian category A is a construction of homological algebra introduced to refine and in a certain sense to simplify the theory of derived functors defined on A. The construction proceeds on the basis that the objects of D(A) should be chain complexes in A, with two such chain complexes considered isomorphic when there is a chain map that induces an isomorphism on the level of homology of the chain complexes. Derived functors can then be defined for chain complexes, refining the concept of hypercohomology. The definitions lead to a significant simplification of formulas otherwise described (not completely faithfully) by complicated spectral sequences.

In mathematics, sheaf cohomology is the application of homological algebra to analyze the global sections of a sheaf on a topological space. Broadly speaking, sheaf cohomology describes the obstructions to solving a geometric problem globally when it can be solved locally. The central work for the study of sheaf cohomology is Grothendieck's 1957 Tôhoku paper.

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, a noetherian scheme is a scheme that admits a finite covering by open affine subsets , noetherian rings. More generally, a scheme is locally noetherian if it is covered by spectra of noetherian rings. Thus, a scheme is noetherian if and only if it is locally noetherian and quasi-compact. As with noetherian rings, the concept is named after Emmy Noether.

In mathematics, the direct image functor is a construction in sheaf theory that generalizes the global sections functor to the relative case. It is of fundamental importance in topology and algebraic geometry. Given a sheaf F defined on a topological space X and a continuous map f: XY, we can define a new sheaf fF on Y, called the direct image sheaf or the pushforward sheaf of F along f, such that the global sections of fF is given by the global sections of F. This assignment gives rise to a functor f from the category of sheaves on X to the category of sheaves on Y, which is known as the direct image functor. Similar constructions exist in many other algebraic and geometric contexts, including that of quasi-coherent sheaves and étale sheaves on a scheme.

In algebraic geometry, an étale morphism is a morphism of schemes that is formally étale and locally of finite presentation. This is an algebraic analogue of the notion of a local isomorphism in the complex analytic topology. They satisfy the hypotheses of the implicit function theorem, but because open sets in the Zariski topology are so large, they are not necessarily local isomorphisms. Despite this, étale maps retain many of the properties of local analytic isomorphisms, and are useful in defining the algebraic fundamental group and the étale topology.

In mathematics, an Azumaya algebra is a generalization of central simple algebras to R-algebras where R need not be a field. Such a notion was introduced in a 1951 paper of Goro Azumaya, for the case where R is a commutative local ring. The notion was developed further in ring theory, and in algebraic geometry, where Alexander Grothendieck made it the basis for his geometric theory of the Brauer group in Bourbaki seminars from 1964–65. There are now several points of access to the basic definitions.

In mathematics, the Leray spectral sequence was a pioneering example in homological algebra, introduced in 1946 by Jean Leray. It is usually seen nowadays as a special case of the Grothendieck spectral sequence.

In mathematics, Verdier duality is a cohomological duality in algebraic topology that generalizes Poincaré duality for manifolds. Verdier duality was introduced in 1965 by Jean-Louis Verdier (1995) as an analog for locally compact topological spaces of Alexander Grothendieck's theory of Poincaré duality in étale cohomology for schemes in algebraic geometry. It is thus one instance of Grothendieck's six operations formalism.

In mathematics a stack or 2-sheaf is, roughly speaking, a sheaf that takes values in categories rather than sets. Stacks are used to formalise some of the main constructions of descent theory, and to construct fine moduli stacks when fine moduli spaces do not exist.

In mathematics, especially in sheaf theory—a domain applied in areas such as topology, logic and algebraic geometry—there are four image functors for sheaves that belong together in various senses.

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.

This is a glossary of algebraic geometry.

In mathematics, Grothendieck's six operations, named after Alexander Grothendieck, is a formalism in homological algebra, also known as the six-functor formalism. It originally sprang from the relations in étale cohomology that arise from a morphism of schemes f : XY. The basic insight was that many of the elementary facts relating cohomology on X and Y were formal consequences of a small number of axioms. These axioms hold in many cases completely unrelated to the original context, and therefore the formal consequences also hold. The six operations formalism has since been shown to apply to contexts such as D-modules on algebraic varieties, sheaves on locally compact topological spaces, and motives.

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).

Faithfully flat descent is a technique from algebraic geometry, allowing one to draw conclusions about objects on the target of a faithfully flat morphism. Such morphisms, that are flat and surjective, are common, one example coming from an open cover.

References