Schubert calculus

Summary

In mathematics, Schubert calculus[1] is a branch of algebraic geometry introduced in the nineteenth century by Hermann Schubert in order to solve various counting problems of projective geometry and, as such, is viewed as part of enumerative geometry. Giving it a more rigorous foundation was the aim of Hilbert's 15th problem. It is related to several more modern concepts, such as characteristic classes, and both its algorithmic aspects and applications remain of current interest. The term Schubert calculus is sometimes used to mean the enumerative geometry of linear subspaces of a vector space, which is roughly equivalent to describing the cohomology ring of Grassmannians. Sometimes it is used to mean the more general enumerative geometry of algebraic varieties that are homogenous spaces of simple Lie groups. Even more generally, Schubert calculus is sometimes understood as encompassing the study of analogous questions in generalized cohomology theories.

The objects introduced by Schubert are the Schubert cells,[2] which are locally closed sets in a Grassmannian defined by conditions of incidence of a linear subspace in projective space with a given flag. For further details see Schubert variety.

The intersection theory[3] of these cells, which can be seen as the product structure in the cohomology ring of the Grassmannian, consisting of associated cohomology classes, allows in particular the determination of cases in which the intersections of cells results in a finite set of points. A key result is that the Schubert cells (or rather, the classes of their Zariski closures, the Schubert cycles or Schubert varieties) span the whole cohomology ring.

The combinatorial aspects mainly arise in relation to computing intersections of Schubert cycles. Lifted from the Grassmannian, which is a homogeneous space, to the general linear group that acts on it, similar questions are involved in the Bruhat decomposition and classification of parabolic subgroups (as block triangular matrices).


Construction edit

Schubert calculus can be constructed using the Chow ring [3] of the Grassmannian, where the generating cycles are represented by geometrically defined data.[4] Denote the Grassmannian of  -planes in a fixed  -dimensional vector space   as  , and its Chow ring as  . (Note that the Grassmannian is sometimes denoted   if the vector space isn't explicitly given or as   if the ambient space   and its  -dimensional subspaces are replaced by their projectizations.) Choosing an (arbitrary) complete flag

 

to each weakly decreasing  -tuple of integers  , where

 

i.e., to each partition of weight

 

whose Young diagram fits into the   rectangular one for the partition  , we associate a Schubert variety[1][2] (or Schubert cycle)  , defined as

 

This is the closure, in the Zariski topology, of the Schubert cell[1][2]

 

which is used when considering cellular homology instead of the Chow ring. The latter are disjoint affine spaces, of dimension  , whose union is  .

An equivalent characterization of the Schubert cell   may be given in terms of the dual complete flag

 

where

 

Then   consists of those  -dimensional subspaces   that have a basis   consisting of elements

 

of the subspaces  

Since the homology class  , called a Schubert class, does not depend on the choice of complete flag  , it can be written as

 

It can be shown that these classes are linearly independent and generate the Chow ring as their linear span. The associated intersection theory is called Schubert calculus. For a given sequence   with   the Schubert class   is usually just denoted  . The Schubert classes given by a single integer  , (i.e., a horizontal partition), are called special classes. Using the Giambelli formula below, all the Schubert classes can be generated from these special classes.

Other notational conventions edit

In some sources,[1][2] the Schubert cells   and Schubert varieties   are labelled differently, as   and  , respectively, where   is the complementary partition to   with parts

 ,

whose Young diagram is the complement of the one for   within the   rectangular one (reversed, both horizontally and vertically).

Another labelling convention for   and   is   and  , respectively, where   is the multi-index defined by

 

The integers   are the pivot locations of the representations of elements of   in reduced matricial echelon form.

Explanation edit

In order to explain the definition, consider a generic  -plane  . It will have only a zero intersection with   for  , whereas

  for  

For example, in  , a  -plane   is the solution space of a system of five independent homogeneous linear equations. These equations will generically span when restricted to a subspace   with  , in which case the solution space (the intersection of   with  ) will consist only of the zero vector. However, if  ,   and   will necessarily have nonzero intersection. For example, the expected dimension of intersection of   and   is  , the intersection of   and   has expected dimension  , and so on.

The definition of a Schubert variety states that the first value of   with   is generically smaller than the expected value   by the parameter  . The  -planes   given by these constraints then define special subvarieties of  .[4]

Properties edit

Inclusion edit

There is a partial ordering on all  -tuples where   if   for every  . This gives the inclusion of Schubert varieties

 

showing an increase of the indices corresponds to an even greater specialization of subvarieties.

Dimension formula edit

A Schubert variety   has dimension equal to the weight

 

of the partition  . Alternatively, in the notational convention   indicated above, its codimension in   is the weight

 

of the complementary partition   in the   dimensional rectangular Young diagram.

This is stable under inclusions of Grassmannians. That is, the inclusion

 

defined, for  , by

 

has the property

 

and the inclusion

 

defined by adding the extra basis element   to each  -plane, giving a  -plane,

 

does as well

 

Thus, if   and   are a cell and a subvariety in the Grassmannian  , they may also be viewed as a cell   and a subvariety   within the Grassmannian   for any pair   with   and  .

Intersection product edit

The intersection product was first established using the Pieri and Giambelli formulas.

Pieri formula edit

In the special case  , there is an explicit formula of the product of   with an arbitrary Schubert class   given by

 

where  ,   are the weights of the partitions. This is called the Pieri formula, and can be used to determine the intersection product of any two Schubert classes when combined with the Giambelli formula. For example,

 

and

 

Giambelli formula edit

Schubert classes   for partitions of any length   can be expressed as the determinant of a   matrix having the special classes as entries.

 

This is known as the Giambelli formula. It has the same form as the first Jacobi-Trudi identity, expressing arbitrary Schur functions   as determinants in terms of the complete symmetric functions  .

For example,

 

and

 

General case edit

The intersection product between any pair of Schubert classes   is given by

 

where   are the Littlewood-Richardson coefficients.[5] The Pieri formula is a special case of this, when   has length  .

Relation with Chern classes edit

There is an easy description of the cohomology ring, or the Chow ring, of the Grassmannian   using the Chern classes of two natural vector bundles over  . We have the exact sequence of vector bundles over  

 

where   is the tautological bundle whose fiber, over any element   is the subspace   itself,   is the trivial vector bundle of rank  , with   as fiber and   is the quotient vector bundle of rank  , with   as fiber. The Chern classes of the bundles   and   are

 

where   is the partition whose Young diagram consists of a single column of length   and

 

The tautological sequence then gives the presentation of the Chow ring as

 

edit

One of the classical examples analyzed is the Grassmannian   since it parameterizes lines in  . Using the Chow ring  , Schubert calculus can be used to compute the number of lines on a cubic surface.[4]

Chow ring edit

The Chow ring has the presentation

 

and as a graded Abelian group[6] it is given by

 

Lines on a cubic surface edit

Recall that a line in   gives a dimension   subspace of  , hence an element of  . Also, the equation of a line can be given as a section of  . Since a cubic surface   is given as a generic homogeneous cubic polynomial, this is given as a generic section  . A line   is a subvariety of   if and only if the section vanishes on  . Therefore, the Euler class of   can be integrated over   to get the number of points where the generic section vanishes on  . In order to get the Euler class, the total Chern class of   must be computed, which is given as

 

The splitting formula then reads as the formal equation

 

where   and   for formal line bundles  . The splitting equation gives the relations

  and  .

Since   can be viewed as the direct sum of formal line bundles

 

whose total Chern class is

 

it follows that

 

using the fact that

  and  

Since   is the top class, the integral is then

 

Therefore, there are   lines on a cubic surface.

See also edit

References edit

  1. ^ a b c d Kleiman, S.L.; Laksov, Dan (1972). "Schubert Calculus". American Mathematical Monthly. 79 (10). American Mathematical Society: 1061–1082. doi:10.1080/00029890.1972.11993188. ISSN 0377-9017.
  2. ^ a b c d Fulton, William (1997). Young Tableaux. With Applications to Representation Theory and Geometry, Chapt. 9.4. London Mathematical Society Student Texts. Vol. 35. Cambridge, U.K.: Cambridge University Press. doi:10.1017/CBO9780511626241. ISBN 9780521567244.
  3. ^ a b Fulton, William (1998). Intersection Theory. Berlin, New York: Springer-Verlag. ISBN 978-0-387-98549-7. MR 1644323.
  4. ^ a b c 3264 and All That (PDF). pp. 132, section 4.1, 200, section 6.2.1.
  5. ^ Fulton, William (1997). Young Tableaux. With Applications to Representation Theory and Geometry, Chapt. 5. London Mathematical Society Student Texts. Vol. 35. Cambridge, U.K.: Cambridge University Press. doi:10.1017/CBO9780511626241. ISBN 9780521567244.
  6. ^ Katz, Sheldon. Enumerative Geometry and String Theory. p. 96.