In algebraic geometry, local cohomology is an algebraic analogue of relative cohomology. Alexander Grothendieck introduced it in seminars in Harvard in 1961 written up by Hartshorne (1967), and in 1961-2 at IHES written up as SGA2 - Grothendieck (1968), republished as Grothendieck (2005). Given a function (more generally, a section of a quasicoherent sheaf) defined on an open subset of an algebraic variety (or scheme), local cohomology measures the obstruction to extending that function to a larger domain. The rational function , for example, is defined only on the complement of on the affine line over a field , and cannot be extended to a function on the entire space. The local cohomology module (where is the coordinate ring of ) detects this in the nonvanishing of a cohomology class . In a similar manner, is defined away from the and axes in the affine plane, but cannot be extended to either the complement of the -axis or the complement of the -axis alone (nor can it be expressed as a sum of such functions); this obstruction corresponds precisely to a nonzero class in the local cohomology module .[1]

Outside of algebraic geometry, local cohomology has found applications in commutative algebra,[2][3][4] combinatorics,[5][6][7] and certain kinds of partial differential equations.[8]

Definition edit

In the most general geometric form of the theory, sections   are considered of a sheaf   of abelian groups, on a topological space  , with support in a closed subset  , The derived functors of   form local cohomology groups

 

In the theory's algebraic form, the space X is the spectrum Spec(R) of a commutative ring R (assumed to be Noetherian throughout this article) and the sheaf F is the quasicoherent sheaf associated to an R-module M, denoted by  . The closed subscheme Y is defined by an ideal I. In this situation, the functor ΓY(F) corresponds to the I-torsion functor, a union of annihilators

 

i.e., the elements of M which are annihilated by some power of I. As a right derived functor, the ith local cohomology module with respect to I is the ith cohomology group   of the chain complex   obtained from taking the I-torsion part   of an injective resolution   of the module  .[9] Because   consists of R-modules and R-module homomorphisms, the local cohomology groups each have the natural structure of an R-module.

The I-torsion part   may alternatively be described as

 

and for this reason, the local cohomology of an R-module M agrees[10] with a direct limit of Ext modules,

 

It follows from either of these definitions that   would be unchanged if   were replaced by another ideal having the same radical.[11] It also follows that local cohomology does not depend on any choice of generators for I, a fact which becomes relevant in the following definition involving the Čech complex.

Using Koszul and Čech complexes edit

The derived functor definition of local cohomology requires an injective resolution of the module  , which can make it inaccessible for use in explicit computations. The Čech complex is seen as more practical in certain contexts. Iyengar et al. (2007), for example, state that they "essentially ignore" the "problem of actually producing any one of these [injective] kinds of resolutions for a given module"[12] prior to presenting the Čech complex definition of local cohomology, and Hartshorne (1977) describes Čech cohomology as "giv[ing] a practical method for computing cohomology of quasi-coherent sheaves on a scheme."[13] and as being "well suited for computations."[14]

The Čech complex can be defined as a colimit of Koszul complexes   where   generate  . The local cohomology modules can be described[15] as:

 

Koszul complexes have the property that multiplication by   induces a chain complex morphism   that is homotopic to zero,[16] meaning   is annihilated by the  . A non-zero map in the colimit of the   sets contains maps from the all but finitely many Koszul complexes, and which are not annihilated by some element in the ideal.

This colimit of Koszul complexes is isomorphic to[17] the Čech complex, denoted  , below.

 

where the ith local cohomology module of   with respect to   is isomorphic to[18] the ith cohomology group of the above chain complex,

 

The broader issue of computing local cohomology modules (in characteristic zero) is discussed in Leykin (2002) and Iyengar et al. (2007, Lecture 23).

Basic properties edit

Since local cohomology is defined as derived functor, for any short exact sequence of R-modules  , there is, by definition, a natural long exact sequence in local cohomology

 

There is also a long exact sequence of sheaf cohomology linking the ordinary sheaf cohomology of X and of the open set U = X \Y, with the local cohomology modules. For a quasicoherent sheaf F defined on X, this has the form

 

In the setting where X is an affine scheme   and Y is the vanishing set of an ideal I, the cohomology groups   vanish for  .[19] If  , this leads to an exact sequence

 

where the middle map is the restriction of sections. The target of this restriction map is also referred to as the ideal transform. For n ≥ 1, there are isomorphisms

 

Because of the above isomorphism with sheaf cohomology, local cohomology can be used to express a number of meaningful topological constructions on the scheme   in purely algebraic terms. For example, there is a natural analogue in local cohomology of the Mayer–Vietoris sequence with respect to a pair of open sets U and V in X, given by the complements of the closed subschemes corresponding to a pair of ideal I and J, respectively.[20] This sequence has the form

 

for any  -module  .

The vanishing of local cohomology can be used to bound the least number of equations (referred to as the arithmetic rank) needed to (set theoretically) define the algebraic set   in  . If   has the same radical as  , and is generated by   elements, then the Čech complex on the generators of   has no terms in degree  . The least number of generators among all ideals   such that   is the arithmetic rank of  , denoted  .[21] Since the local cohomology with respect to   may be computed using any such ideal, it follows that   for  .[22]

Graded local cohomology and projective geometry edit

When   is graded by  ,   is generated by homogeneous elements, and   is a graded module, there is a natural grading on the local cohomology module   that is compatible with the gradings of   and  .[23] All of the basic properties of local cohomology expressed in this article are compatible with the graded structure.[24] If   is finitely generated and   is the ideal generated by the elements of   having positive degree, then the graded components   are finitely generated over   and vanish for sufficiently large  .[25]

The case where   is the ideal generated by all elements of positive degree (sometimes called the irrelevant ideal) is particularly special, due to its relationship with projective geometry.[26] In this case, there is an isomorphism

 

where   is the projective scheme associated to  , and   denotes the Serre twist. This isomorphism is graded, giving

 

in all degrees  .[27]

This isomorphism relates local cohomology with the global cohomology of projective schemes. For example, the Castelnuovo–Mumford regularity can be formulated using local cohomology[28] as

 

where   denotes the highest degree   such that  . Local cohomology can be used to prove certain upper bound results concerning the regularity.[29]

Examples edit

Top local cohomology edit

Using the Čech complex, if   the local cohomology module   is generated over   by the images of the formal fractions

 

for   and  .[30] This fraction corresponds to a nonzero element of   if and only if there is no   such that  .[31] For example, if  , then

 
  • If   is a field and   is a polynomial ring over   in   variables, then the local cohomology module   may be regarded as a vector space over   with basis given by (the Čech cohomology classes of) the inverse monomials   for  .[32] As an  -module, multiplication by   lowers   by 1, subject to the condition   Because the powers   cannot be increased by multiplying with elements of  , the module   is not finitely generated.

Examples of H1 edit

If   is known (where  ), the module   can sometimes be computed explicitly using the sequence

 

In the following examples,   is any field.

  • If   and  , then   and as a vector space over  , the first local cohomology module   is  , a 1-dimensional   vector space generated by  .[33]
  • If   and  , then   and  , so   is an infinite-dimensional   vector space with basis  [34]

Relation to invariants of modules edit

The dimension dimR(M) of a module (defined as the Krull dimension of its support) provides an upper bound for local cohomology modules:[35]

 

If R is local and M finitely generated, then this bound is sharp, i.e.,  .

The depth (defined as the maximal length of a regular M-sequence; also referred to as the grade of M) provides a sharp lower bound, i.e., it is the smallest integer n such that[36]

 

These two bounds together yield a characterisation of Cohen–Macaulay modules over local rings: they are precisely those modules where   vanishes for all but one n.

Local duality edit

The local duality theorem is a local analogue of Serre duality. For a Cohen-Macaulay local ring   of dimension   that is a homomorphic image of a Gorenstein local ring[37] (for example, if   is complete[38]), it states that the natural pairing

 

is a perfect pairing, where   is a dualizing module for  .[39] In terms of the Matlis duality functor  , the local duality theorem may be expressed as the following isomorphism.[40]

 

The statement is simpler when  , which is equivalent[41] to the hypothesis that   is Gorenstein. This is the case, for example, if   is regular.

Applications edit

The initial applications were to analogues of the Lefschetz hyperplane theorems. In general such theorems state that homology or cohomology is supported on a hyperplane section of an algebraic variety, except for some 'loss' that can be controlled. These results applied to the algebraic fundamental group and to the Picard group.

Another type of application are connectedness theorems such as Grothendieck's connectedness theorem (a local analogue of the Bertini theorem) or the Fulton–Hansen connectedness theorem due to Fulton & Hansen (1979) and Faltings (1979). The latter asserts that for two projective varieties V and W in Pr over an algebraically closed field, the connectedness dimension of Z = VW (i.e., the minimal dimension of a closed subset T of Z that has to be removed from Z so that the complement Z \ T is disconnected) is bound by

c(Z) ≥ dim V + dim Wr − 1.

For example, Z is connected if dim V + dim W > r.[42]

In polyhedral geometry, a key ingredient of Stanley’s 1975 proof of the simplicial form of McMullen’s Upper bound theorem involves showing that the Stanley-Reisner ring of the corresponding simplicial complex is Cohen-Macaulay, and local cohomology is an important tool in this computation, via Hochster’s formula.[43][6][44]

See also edit

Notes edit

  1. ^ Hartshorne (1977, Exercise 4.3)
  2. ^ Eisenbud (2005, Chapter 4, Castelnuovo-Mumford Regularity)
  3. ^ Brodmann & Sharp (1998, Chapter 17, Hilbert Polynomials)
  4. ^ Brodmann & Sharp (1998, Chapter 18, Applications to reductions of ideals)
  5. ^ Huang (2002, Chapter 10, Residue Methods in Combinatorial Analysis)
  6. ^ a b Stanley, Richard (1996). Combinatorics and commutative algebra. Boston, MA: Birkhäuser Boston, Inc. p. 164. ISBN 0-8176-3836-9.
  7. ^ Iyengar et al. (2007, Lecture 16, Polyhedral Geometry)
  8. ^ Iyengar et al. (2007, Lecture 24, Holonomic Rank and Hypergeometric Systems)
  9. ^ Brodmann & Sharp (1998, 1.2.2)
  10. ^ Brodmann & Sharp (1998, Theorem 1.3.8)
  11. ^ Brodmann & Sharp (1998, Remark 1.2.3)
  12. ^ Iyengar et al. (2007)
  13. ^ Hartshorne (1977, p. 218)
  14. ^ Hartshorne (1977, p. 219)
  15. ^ Brodmann & Sharp (1998, Theorem 5.2.9)
  16. ^ "Lemma 15.28.6 (0663)—The Stacks project". stacks.math.columbia.edu. Retrieved 2020-05-01.
  17. ^ "Lemma 15.28.13 (0913)—The Stacks project". stacks.math.columbia.edu. Retrieved 2020-05-01.
  18. ^ Brodmann & Sharp (1998, Theorem 5.1.19)
  19. ^ Hartshorne (1977, Theorem 3.7)
  20. ^ Brodmann & Sharp (1998, Theorem 3.2.3)
  21. ^ Brodmann & Sharp (1998, Definition 3.3.2)
  22. ^ Brodmann & Sharp (1998, Remark 5.1.20)
  23. ^ Brodmann & Sharp (1998, Corollary 12.3.3)
  24. ^ Brodmann & Sharp (1998, Chapter 13)
  25. ^ Brodmann & Sharp (1998, Proposition 15.1.5)
  26. ^ Eisenbud (1995, §A.4)
  27. ^ Brodmann & Sharp (1998, Theorem 20.4.4)
  28. ^ Brodmann & Sharp (1998, Definition 15.2.9)
  29. ^ Brodmann & Sharp (1998, Chapter 16)
  30. ^ Iyengar et al. (2007, Corollary 7.14)
  31. ^ Brodmann & Sharp (1998, Exercise 5.1.21)
  32. ^ Iyengar et al. (2007, Exercise 7.16)
  33. ^ Brodmann & Sharp (1998, Exercise 2.3.6(v))
  34. ^ Eisenbud (2005, Example A1.10)
  35. ^ Brodmann & Sharp (1998, Theorem 6.1.2)
  36. ^ Hartshorne (1967, Theorem 3.8), Brodmann & Sharp (1998, Theorem 6.2.7), M is finitely generated, IMM
  37. ^ Bruns & Herzog (1998, Theorem 3.3.6)
  38. ^ Bruns & Herzog (1998, Corollary 3.3.8)
  39. ^ Hartshorne (1967, Theorem 6.7)
  40. ^ Brodmann & Sharp (1998, Theorem 11.2.8)
  41. ^ Bruns & Herzog (1998, Theorem 3.3.7)
  42. ^ Brodmann & Sharp (1998, §19.6)
  43. ^ Stanley, Richard (2014). "How the Upper Bound Conjecture Was Proved". Annals of Combinatorics. 18 (3): 533–539. doi:10.1007/s00026-014-0238-5. S2CID 253585250.
  44. ^ Iyengar et al. (2007, Lecture 16)

Introductory Reference edit

References edit