To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
Live Statistics
English Articles
Improved in 24 Hours
Added in 24 Hours
Languages
Recent
Show all languages
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Kähler differential

From Wikipedia, the free encyclopedia

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.

YouTube Encyclopedic

  • 1/5
    Views:
    2 132
    688
    3 235
    1 266
    754
  • Sheaf of Differential Forms
  • Riemannian Geometry,Kahler Manifold with the help of Almost Hermite Manifold
  • Complex Differential Geometry | Dr Vamsi Pingali 1
  • Seminário: Elie Cartan, differential systems, transformation groups and differential... (2020)
  • Finding solitons in differential geometry

Transcription

Definition

Let R and S be commutative rings and φ : RS be a ring homomorphism. An important example is for R a field and S a unital algebra over R (such as the coordinate ring of an affine variety). Kähler differentials formalize the observation that the derivatives of polynomials are again polynomial. In this sense, differentiation is a notion which can be expressed in purely algebraic terms. This observation can be turned into a definition of the module

of differentials in different, but equivalent ways.

Definition using derivations

An R-linear derivation on S is an R-module homomorphism to an S-module M satisfying the Leibniz rule (it automatically follows from this definition that the image of R is in the kernel of d [1]). The module of Kähler differentials is defined as the S-module for which there is a universal derivation . As with other universal properties, this means that d is the best possible derivation in the sense that any other derivation may be obtained from it by composition with an S-module homomorphism. In other words, the composition with d provides, for every S-module M, an S-module isomorphism

One construction of ΩS/R and d proceeds by constructing a free S-module with one formal generator ds for each s in S, and imposing the relations

  • dr = 0,
  • d(s + t) = ds + dt,
  • d(st) = s dt + t ds,

for all r in R and all s and t in S. The universal derivation sends s to ds. The relations imply that the universal derivation is a homomorphism of R-modules.

Definition using the augmentation ideal

Another construction proceeds by letting I be the ideal in the tensor product defined as the kernel of the multiplication map

Then the module of Kähler differentials of S can be equivalently defined by[2]

and the universal derivation is the homomorphism d defined by

This construction is equivalent to the previous one because I is the kernel of the projection

Thus we have:

Then may be identified with I by the map induced by the complementary projection

This identifies I with the S-module generated by the formal generators ds for s in S, subject to d being a homomorphism of R-modules which sends each element of R to zero. Taking the quotient by I2 precisely imposes the Leibniz rule.

Examples and basic facts

For any commutative ring R, the Kähler differentials of the polynomial ring are a free S-module of rank n generated by the differentials of the variables:

Kähler differentials are compatible with extension of scalars, in the sense that for a second R-algebra R and for , there is an isomorphism

As a particular case of this, Kähler differentials are compatible with localizations, meaning that if W is a multiplicative set in S, then there is an isomorphism

Given two ring homomorphisms , there is a short exact sequence of T-modules

If for some ideal I, the term vanishes and the sequence can be continued at the left as follows:

A generalization of these two short exact sequences is provided by the cotangent complex.

The latter sequence and the above computation for the polynomial ring allows the computation of the Kähler differentials of finitely generated R-algebras . Briefly, these are generated by the differentials of the variables and have relations coming from the differentials of the equations. For example, for a single polynomial in a single variable,

Kähler differentials for schemes

Because Kähler differentials are compatible with localization, they may be constructed on a general scheme by performing either of the two definitions above on affine open subschemes and gluing. However, the second definition has a geometric interpretation that globalizes immediately. In this interpretation, I represents the ideal defining the diagonal in the fiber product of Spec(S) with itself over Spec(S) → Spec(R). This construction therefore has a more geometric flavor, in the sense that the notion of first infinitesimal neighbourhood of the diagonal is thereby captured, via functions vanishing modulo functions vanishing at least to second order (see cotangent space for related notions). Moreover, it extends to a general morphism of schemes by setting to be the ideal of the diagonal in the fiber product . The cotangent sheaf , together with the derivation defined analogously to before, is universal among -linear derivations of -modules. If U is an open affine subscheme of X whose image in Y is contained in an open affine subscheme V, then the cotangent sheaf restricts to a sheaf on U which is similarly universal. It is therefore the sheaf associated to the module of Kähler differentials for the rings underlying U and V.

Similar to the commutative algebra case, there exist exact sequences associated to morphisms of schemes. Given morphisms and of schemes there is an exact sequence of sheaves on

Also, if is a closed subscheme given by the ideal sheaf , then and there is an exact sequence of sheaves on

Examples

Finite separable field extensions

If is a finite field extension, then if and only if is separable. Consequently, if is a finite separable field extension and is a smooth variety (or scheme), then the relative cotangent sequence

proves .

Cotangent modules of a projective variety

Given a projective scheme , its cotangent sheaf can be computed from the sheafification of the cotangent module on the underlying graded algebra. For example, consider the complex curve

then we can compute the cotangent module as

Then,

Morphisms of schemes

Consider the morphism

in . Then, using the first sequence we see that

hence

Higher differential forms and algebraic de Rham cohomology

de Rham complex

As before, fix a map . Differential forms of higher degree are defined as the exterior powers (over ),

The derivation extends in a natural way to a sequence of maps

satisfying This is a cochain complex known as the de Rham complex.

The de Rham complex enjoys an additional multiplicative structure, the wedge product

This turns the de Rham complex into a commutative differential graded algebra. It also has a coalgebra structure inherited from the one on the exterior algebra.[3]

de Rham cohomology

The hypercohomology of the de Rham complex of sheaves is called the algebraic de Rham cohomology of X over Y and is denoted by or just if Y is clear from the context. (In many situations, Y is the spectrum of a field of characteristic zero.) Algebraic de Rham cohomology was introduced by Grothendieck (1966a). It is closely related to crystalline cohomology.

As is familiar from coherent cohomology of other quasi-coherent sheaves, the computation of de Rham cohomology is simplified when X = Spec S and Y = Spec R are affine schemes. In this case, because affine schemes have no higher cohomology, can be computed as the cohomology of the complex of abelian groups

which is, termwise, the global sections of the sheaves .

To take a very particular example, suppose that is the multiplicative group over Because this is an affine scheme, hypercohomology reduces to ordinary cohomology. The algebraic de Rham complex is

The differential d obeys the usual rules of calculus, meaning The kernel and cokernel compute algebraic de Rham cohomology, so

and all other algebraic de Rham cohomology groups are zero. By way of comparison, the algebraic de Rham cohomology groups of are much larger, namely,

Since the Betti numbers of these cohomology groups are not what is expected, crystalline cohomology was developed to remedy this issue; it defines a Weil cohomology theory over finite fields.

Grothendieck's comparison theorem

If X is a smooth complex algebraic variety, there is a natural comparison map of complexes of sheaves

between the algebraic de Rham complex and the smooth de Rham complex defined in terms of (complex-valued) differential forms on , the complex manifold associated to X. Here, denotes the complex analytification functor. This map is far from being an isomorphism. Nonetheless, Grothendieck (1966a) showed that the comparison map induces an isomorphism

from algebraic to smooth de Rham cohomology (and thus to singular cohomology by de Rham's theorem). In particular, if X is a smooth affine algebraic variety embedded in , then the inclusion of the subcomplex of algebraic differential forms into that of all smooth forms on X is a quasi-isomorphism. For example, if

,

then as shown above, the computation of algebraic de Rham cohomology gives explicit generators for and , respectively, while all other cohomology groups vanish. Since X is homotopy equivalent to a circle, this is as predicted by Grothendieck's theorem.

Counter-examples in the singular case can be found with non-Du Bois singularities such as the graded ring with where and .[4] Other counterexamples can be found in algebraic plane curves with isolated singularities whose Milnor and Tjurina numbers are non-equal.[5]

A proof of Grothendieck's theorem using the concept of a mixed Weil cohomology theory was given by Cisinski & Déglise (2013).

Applications

Canonical divisor

If X is a smooth variety over a field k,[clarification needed] then is a vector bundle (i.e., a locally free -module) of rank equal to the dimension of X. This implies, in particular, that

is a line bundle or, equivalently, a divisor. It is referred to as the canonical divisor. The canonical divisor is, as it turns out, a dualizing complex and therefore appears in various important theorems in algebraic geometry such as Serre duality or Verdier duality.

Classification of algebraic curves

The geometric genus of a smooth algebraic variety X of dimension d over a field k is defined as the dimension

For curves, this purely algebraic definition agrees with the topological definition (for ) as the "number of handles" of the Riemann surface associated to X. There is a rather sharp trichotomy of geometric and arithmetic properties depending on the genus of a curve, for g being 0 (rational curves), 1 (elliptic curves), and greater than 1 (hyperbolic Riemann surfaces, including hyperelliptic curves), respectively.

Tangent bundle and Riemann–Roch theorem

The tangent bundle of a smooth variety X is, by definition, the dual of the cotangent sheaf . The Riemann–Roch theorem and its far-reaching generalization, the Grothendieck–Riemann–Roch theorem, contain as a crucial ingredient the Todd class of the tangent bundle.

Unramified and smooth morphisms

The sheaf of differentials is related to various algebro-geometric notions. A morphism of schemes is unramified if and only if is zero.[6] A special case of this assertion is that for a field k, is separable over k iff , which can also be read off the above computation.

A morphism f of finite type is a smooth morphism if it is flat and if is a locally free -module of appropriate rank. The computation of above shows that the projection from affine space is smooth.

Periods

Periods are, broadly speaking, integrals of certain arithmetically defined differential forms.[7] The simplest example of a period is , which arises as

Algebraic de Rham cohomology is used to construct periods as follows:[8] For an algebraic variety X defined over the above-mentioned compatibility with base-change yields a natural isomorphism

On the other hand, the right hand cohomology group is isomorphic to de Rham cohomology of the complex manifold associated to X, denoted here Yet another classical result, de Rham's theorem, asserts an isomorphism of the latter cohomology group with singular cohomology (or sheaf cohomology) with complex coefficients, , which by the universal coefficient theorem is in its turn isomorphic to Composing these isomorphisms yields two rational vector spaces which, after tensoring with become isomorphic. Choosing bases of these rational subspaces (also called lattices), the determinant of the base-change matrix is a complex number, well defined up to multiplication by a rational number. Such numbers are periods.

Algebraic number theory

In algebraic number theory, Kähler differentials may be used to study the ramification in an extension of algebraic number fields. If L / K is a finite extension with rings of integers R and S respectively then the different ideal δL / K, which encodes the ramification data, is the annihilator of the R-module ΩR/S:[9]

Related notions

Hochschild homology is a homology theory for associative rings that turns out to be closely related to Kähler differentials. This is because of the Hochschild-Kostant-Rosenberg theorem which states that the Hochschild homology of an algebra of a smooth variety is isomorphic to the de-Rham complex for a field of characteristic . A  derived  enhancement of this theorem states that the Hochschild homology of a differential graded algebra is isomorphic to the derived de-Rham complex.

The de Rham–Witt complex is, in very rough terms, an enhancement of the de Rham complex for the ring of Witt vectors.

Notes

  1. ^ "Stacks Project". Retrieved 2022-11-21.
  2. ^ Hartshorne (1977, p. 172)
  3. ^ Laurent-Gengoux, C.; Pichereau, A.; Vanhaecke, P. (2013), Poisson structures, §3.2.3: Springer, ISBN 978-3-642-31090-4{{citation}}: CS1 maint: location (link)
  4. ^ "algebraic de Rham cohomology of singular varieties", mathoverflow.net
  5. ^ Arapura, Donu; Kang, Su-Jeong (2011), "Kähler-de Rham cohomology and Chern classes" (PDF), Communications in Algebra, 39 (4): 1153–1167, doi:10.1080/00927871003610320, MR 2782596, S2CID 15924437, archived from the original (PDF) on 2015-11-12
  6. ^ Milne, James, Etale cohomology, Proposition I.3.5{{citation}}: CS1 maint: location (link); the map f is supposed to be locally of finite type for this statement.
  7. ^ André, Yves (2004), Une introduction aux motifs, Partie III: Société Mathématique de France
  8. ^ Periods and Nori Motives (PDF), Elementary examples
  9. ^ Neukirch (1999, p. 201)

References

External links

This page was last edited on 20 September 2023, at 08:13
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.