An equivariant basis for the cohomology of Springer fibers

By Martha Precup and Edward Richmond

Abstract

Springer fibers are subvarieties of the flag variety that play an important role in combinatorics and geometric representation theory. In this paper, we analyze the equivariant cohomology of Springer fibers for using results of Kumar and Procesi that describe this equivariant cohomology as a quotient ring. We define a basis for the equivariant cohomology of a Springer fiber, generalizing a monomial basis of the ordinary cohomology defined by De Concini and Procesi and studied by Garsia and Procesi. Our construction yields a combinatorial framework with which to study the equivariant and ordinary cohomology rings of Springer fibers. As an application, we identify an explicit collection of (equivariant) Schubert classes whose images in the (equivariant) cohomology ring of a given Springer fiber form a basis.

1. Introduction

This paper analyzes the equivariant cohomology of Springer fibers in Lie type A. Springer fibers are fibers of a desingularization of the nilpotent cone in . Springer showed that the symmetric group acts on the cohomology of each Springer fiber, the top-dimensional cohomology is an irreducible representation, and each irreducible symmetric group representation can be obtained in this way Reference 35Reference 36. As a consequence, Springer fibers frequently arise in geometric representation theory and algebraic combinatorics; see Reference 14Reference 15Reference 18Reference 20Reference 32Reference 34 for just a few examples.

There is also an algebraic approach to the Springer representation for , as we now explain. Motivated by a conjecture of Kraft Reference 25, De Concini and Procesi Reference 8 gave a presentation for the cohomology of a type A Springer fiber as the quotient of a polynomial ring. Furthermore, this identification is -equivariant so Springer’s representation can also be constructed as the symmetric group action on the quotient of a polynomial ring. These results were generalized to the setting of other algebraic groups by Carrell in Reference 6.

The generators of the ideal defining the presentation of the cohomology of a type A Springer fiber were further simplified by Tanisaki Reference 37. Finally, Garsia and Procesi used the aforementioned results to study the graded character of the Springer representation in Reference 19. Their work gives a linear algebraic proof that this character is closely connected to the so-called -Kostka polynomials. As part of their analysis, Garsia and Procesi study a monomial basis for the cohomology ring, originally defined by De Concini and Procesi in Reference 8, with many amenable combinatorial and inductive properties. We refer to the collection of these monomials as the Springer monomial basis.

Let denote the algebraic group of invertible matrices with Lie algebra of matrices. Denote by the Borel subgroup of upper triangular matrices, and by its Lie algebra. Given a nilpotent matrix , let be the partition of determined by the sizes of the Jordan blocks of . The flag variety of is the quotient and the Springer fiber corresponding to is defined as the subvariety

Let denote the maximal torus of diagonal matrices in and be the Levi subgroup of block diagonal matrices with block sizes determined by the partition . We may assume without loss of generality that is in Jordan canonical form, and hence is regular in the Lie algebra of . Moreover, the subtorus acts on the Springer fiber . We consider the equivariant cohomology . The goal of this manuscript is provide a combinatorial framework to study this equivariant cohomology.

There is a known presentation for given by Kumar and Procesi Reference 26, and the equivariant Tanisaki ideal has been determined by Abe and Horiguchi Reference 1. Our work below initiates a study of which parallels the analysis of the ordinary cohomology by Garsia and Procesi in Reference 19. We define a collection of polynomials in using the combinatorics of row-strict tableaux. Since these polynomials map onto the Springer monomial basis under the natural projection map from equivariant to ordinary cohomology , we call them equivariant Springer monomials. We prove that a basis of equivariant Springer monomials exists for any Springer fiber, and provide a determinant formula (see Theorem 4.5 below) for the structure constants of any element of with respect to this basis.

As an application, we use the algebraic and combinatorial framework developed in this manuscript to study the images of Schubert classes in . Let denote the inclusion of varieties, and the induced map on ordinary cohomology. We prove that for every partition , there is a natural collection of Schubert classes whose images under form an additive basis of . This result appears as Theorem 5.9 in Section 5 below and Corollary 5.14 contains the equivariant version of the statement. Phrased in terms of the work of Harada and Tymoczko in Reference 22, the equivariant version of Theorem 5.9 says that there exists a successful game of Betti poset pinball for each type A Springer fiber. As a result, we can do computations in the (equivariant) cohomology ring more easily, as combinatorial properties of (double) Schubert polynomials are well-studied (c.f., for example, Reference 29). Bases of this kind have been used to do Schubert calculus style computations in the equivariant cohomology rings of other subvarieties of the flag variety Reference 10Reference 21; the authors will explore analogous computations for Springer fibers in future work.

Our Theorem 5.9 generalizes results of Harada–Tymoczko Reference 22 and Dewitt–Harada Reference 9 which address the case of and , respectively. The main difficulty in generalizing the methods used in those papers is that the equivariant cohomology classes in constructed via poset pinball may not satisfy upper triangular vanishing conditions (with respect to some partial ordering on the set of -fixed points of ). The methods used to prove Theorem 5.9 side-step this difficulty by making use of the equivariant Springer monomials. Combining our determinantal formula for the structure coefficients of this basis with known combinatorial properties of the Schubert polynomials yields the desired result.

Let denote the Schubert variety corresponding to a permutation . Recall that the Schubert polynomial represents the fundamental cohomology class of the Schubert variety where denotes the longest element in . That is, is a polynomial representative for the cohomology class defined uniquely by the property that . Here and denote the fundamental homology classes of and , respectively, and denotes the Poincaré duality isomorphism obtained by taking the cap product with the top fundamental class.

In this paper, we study the polynomials in from a combinatorial perspective. On the other hand, each is a polynomial representative for the cohomology class and it is natural to ask if these classes have geometric meaning. In the last section, we show that the classes play an analogous role with respect to the homology of as that played by the Schubert classes with respect to the homology of . More precisely, we prove in Proposition 6.1 below that

for generic . Here denotes capping with the top fundamental class . Since is typically not smooth, this map is not an isomorphism of groups.

The remainder of the paper is structured as follows. The next section covers the necessary background information and notation needed in later sections, including a presentation of the equivariant cohomology of the Springer fiber due to Kumar and Procesi. The third and fourth sections of this paper establish the combinatorial groundwork for our study of . We use row-strict composition tableaux to define an equivariant generalization of the Springer monomial basis in Section 3, called the equivariant Springer monomials, and develop the structural properties of these polynomials further in Section 4. In particular, we give a determinant formula for the structure coefficients of with respect to the basis of equivariant Springer monomials in Theorem 4.5 of Section 4. Finally, Section 5 uses the equivariant Springer monomials to study the images of monomials and Schubert polynomials in the cohomology of Springer fibers. Our main result in Section 5 is Theorem 5.9, which was discussed above. We conclude with an analysis of the geometric meaning of the classes in Section 6.

2. Background

As in the introduction, let and denote its Lie algebra. Denote by the maximal torus of diagonal matrices in and by the Borel subgroup of upper triangular matrices. Let denote the Lie algebra of . The Weyl group of is . We let denote the simple transposition exchanging and . Throughout this manuscript, denotes a (strong) composition of . We call the partition of obtained by sorting the parts of into weakly decreasing order the underlying partition shape of .

The composition uniquely determines a standard Levi subgroup in , namely the subgroup of block diagonal matrices such that the -th diagonal block has dimension . We denote the Weyl group for by . Let be a principal nilpotent element of , the Lie algebra of . Note that by construction, is a nilpotent matrix of Jordan type , where is the underlying partition shape of .

Let denote the flag variety. The Springer fiber of is defined to be

If two compositions have the same underlying partition shape, then the corresponding Springer fibers are isomorphic. However, taking different compositions corresponding to the same partition shape yields actions of different sub-tori of on the corresponding Springer fibers. This ultimately leads to the construction of different bases for the equivariant cohomology ring of .

Let denote the connected component of the centralizer of in containing the identity, so . Since , we get that centralizes and therefore acts on by left multiplication. The purpose of this manuscript is to study the equivariant cohomology ring . We begin by reviewing a presentation for due to Kumar and Procesi Reference 26.

2.1. A presentation of

Recall from the introduction that denotes the inclusion map of into the flag variety and consider the induced map on equivariant cohomology, . In this paper, we work with singular and equivariant cohomology with coefficients in . Note that naturally factors through ,

Let denote the Lie algebra of . The coordinate ring of is the polynomial ring

where is the ideal .

It is well known that the -action on by left multiplication is equivariantly formal, implying

Since we have that is a free -module. Recall that the Borel homomorphism,

is defined by , where is the -equivariant first Chern class of , the -th line bundle of the tautological filtration of sub-bundles on . In other words, the fiber of over a flag is the line . This map induces a surjective algebra homomorphism,

given by where . Following the work the Kumar in Procesi Reference 26, we define to be the composition of maps

Let denote the Lie algebra of and be the reduced closed subvariety of defined by

Note that we may also view as a subvariety of , and we use this perspective in our computations below. Let denote the vanishing ideal of . The coordinate ring

is naturally an -algebra via the projection onto the first factor. Moreover, the ring inherits a non-negatively graded structure from . We also define the graded -algebra

where is considered an -module under evaluation at 0. Note that if , then and . In this special case, we denote the corresponding coordinate ring by . The next theorem from Reference 26 gives a presentation of .

Theorem 2.1 (Kumar–Procesi).

The kernel of the map defined in Equation 2.2 is the ideal . In particular, induces a graded -algebra isomorphism

making the following diagram commute.

Furthermore, the map naturally descends to a -algebra isomorphism:

with the following commutative diagram.

Since the map is an isomorphism, we will use for the respective restriction maps and . In particular, if denotes a Schubert polynomial representing the class , then the polynomial represents the class .

Remark 2.2.

It is well-known that the cohomology is concentrated in even degrees Reference 34. Thus the equivariant cohomology is a free -module, and isomorphic to the tensor product,

The graded -algebra isomorphism of Theorem 2.1 implies is a free -module with rank equal to the number of -fixed points of , namely (c.f. Reference 26, Lemma 2.1).

2.2. Maps of polynomial rings

Recall that is the standard Levi subgroup associated to as above and is the connected component of the centralizer of in containing the identity. Since is a standard Levi subgroup, if and has coordinates , then the embedding of the subalgebra into is given by

where

This embedding induces a map If , then and have the same values on . This implies

where denotes the vanishing ideal of as a subvariety of . We let denote the canonical projection map. By a slight abuse of notation, we will also denote the quotient by ; the isomorphism of Equation 2.5 tells us that we may do so without loss of generality.

As in Equation 2.1, there are isomorphisms

where is the ideal . Note that

and we make this identification below whenever it is convenient (and similarly for ).

The ring inherits the graded structure of . In particular, the degree component of is and we denote its image under the canonical projection map by . Let

denote the positive degree and degree zero components of with respect to the grading of . It is easy to see that and There is a surjective map

given by evaluation at . More explicitly, if with and , then (then extend linearly to all of ). This map induces an evaluation map giving the commutative diagram:

where, by Theorem 2.1, and . Note that, under these identifications, the evaluation map is simply the usual restriction map from equivariant to ordinary cohomology. The following lemma relates bases of the ordinary and equivariant cohomology rings of .

Lemma 2.3.

Suppose is a homogeneous basis of and let be a set of homogeneous polynomials in such that . Then is an -module basis of .

Proof.

To begin, we prove that the -span of is . First note that if , then and for some . Let . It suffices to assume that for some . We proceed by induction on . Since is a basis of , we can write

for some and hence for some and for all . We now have

Observe that the second term of the above sum belongs to and is therefore of the form where each from some . By induction, each is a -linear combination of and hence so is .

We next prove that is -linearly independent. As noted in Remark 2.2, is a free -module of rank . Let denote the field of fractions of . Since is a free module, the extension of scalars is a free module of the same rank Reference 11, §10.4, Cor. 18. Furthermore, the polynomials must also span . Since the extension of scalars is an -dimensional vector space, are -linearly independent. Any non-trivial linear relation among with -coefficients would also be a non-trivial linear relation over , contradicting the previous sentence. We conclude is -linearly independent, as desired.

2.3. The Springer monomial basis

We now recall the monomial basis of defined by De Concini and Procesi in Reference 8 and further analyzed by Garsia and Procesi in Reference 19.

Let be a partition of with parts and be the partition of obtained from by decreasing the -th part by and sorting the resulting composition so that the parts are in weakly decreasing order. Define to be the collection of monomials constructed recursively as in Reference 19, §1 by

with initial condition for . Here denotes the set of monomials obtained by multiplying each monomial in by . Observe that as defined in Reference 19, the monomials in are in the variables . We define

where the action of the longest permutation on variables is given by . Hence the monomials in are in the variables . Since the ideal is invariant under the action of , it follows by results of De Concini and Procesi that, as graded vector spaces,

Here we use standard monomial notation

We refer to the basis of as the Springer monomial basis, and to its elements as Springer monomials. We adopt the convention throughout this manuscript that if , then we denote both and its image under the canonical projection by the same symbol.

Example 2.4.

Let and , then .

See Reference 19, §1 for a more detailed example.

Remark 2.5.

The Springer monomials have been generalized to study the cohomology rings of other subvarieties of the flag variety. In particular, Mbirika in Reference 30 constructs an analogous set of monomials for nilpotent Hessenberg varieties (which include Springer fibers). In a later paper, Mbirika and Tymoczko give an analogue of the Tanisaki ideal in the Hessenberg setting Reference 31.

3. Row-strict tableaux

In this section we develop a combinatorial framework to study the ring defined in Equation 2.3 using row-strict composition tableaux.

3.1. Row strict composition tableaux

For any integers , we let denote the interval . If , then we set . Given any , consider a weak composition of . The composition diagram of is an array of boxes with boxes in the -th row ordered from top to bottom (English notation).

A shifted row-strict composition tableau of shape is a labeling of the composition diagram with the integers such that the values decrease from left to right in each row. For simplicity of notation, let . Let denote the collection of all shifted-standard row-strict tableaux of composition shape with content . Observe that if (i.e. is a composition of ), then the content of is the full standard content ; in this case, we say that is a row-strict composition tableau.

Example 3.1.

Consider the composition with . In this case and . There are 12 row-strict composition tableaux in . Indeed, note that there are possible fillings of using the content . Furthermore, if we define two fillings to be equivalent up to the entries in each row, e.g.

then there are precisely two tableaux in each equivalence class and each class contains a unique row-strict composition tableau.

Given a composition , let be the strong composition obtained from by deleting any part equal to zero. By similar reasoning as in the example above, we have that

which is precisely the number of -fixed points in the Springer fiber . Notice that if , then each shifted row-strict composition tableau can be associated to a unique row-strict tableau in by the relabeling map . We use the “shifted” terminology since it simplifies the arguments below. Similarly, although we typically begin with a strong composition of , our inductive procedures require the generality of weak compositions.

We now define a map

where the union on the RHS is taken over all compositions obtained from by deleting one box from any nonzero row. Let be the composition tableau obtained by removing the box from which contains its smallest entry, namely . For example:

In this case, the disjoint union in Equation 3.1 is taken over . The map plays an important role in the inductive arguments below; note that is in fact a bijection.

Definition 3.2.

Let . We say that is an Springer inversion of if there exists in row such that and either:

(1)

appears above and in the same column, or

(2)

appears in a column strictly to the right of the column containing .

Denote the set of Springer inversions of by .

Example 3.3.

Let and . Consider with content :

The inversions of are .

Remark 3.4.

Note that the definition above is closely related to the notion of a Springer dimension pair considered by the first author and Tymoczko in Reference 32. In that paper, the convention is that the row-strict tableaux have increasing entries (from left to right), while our convention is that the entries are decreasing (from left to right). This change in conventions is routine; to convert from one to the other, apply the permutation such that for all . A Springer inversion from this paper corresponds to a unique Springer dimension pair as defined in Reference 32 (up to transformation under ). If is a Springer inversion then is a Springer dimension pair, where denotes the smallest element in row such that .

The following lemma is a simple, but important fact about inversions.

Lemma 3.5.

Let . Let denote the indices of the rows containing in and , respectively. Then exactly one of the following is true:

(1)

(2)

(3)

.

Proof.

Without loss of generality, suppose that and hence is contained in different rows of the tableaux and . Since is the smallest number in the content, it must lie at the end its respective row of and . Moreover, the content of the row indexed by in is strictly larger than and vice versa. If the size of row is at least the size of row , then . Otherwise, .

Lemma 3.5 induces a total ordering on the set as follows.

Definition 3.6.

Let and denote the indices of the rows containing in and , respectively. First suppose . We say if and if . Otherwise, if , then and have the same composition shape. In this case, we inductively say if .

Example 3.7.

Let and . The total order on tableaux in is displayed below.

In the next section we will associate a unique monomial to each element of . We will see that the total ordering on the shifted row-strict composition tableaux defined above corresponds to the lex ordering on these monomials.

3.2. Equivariant Springer monomials

In this section we define a collection of polynomials indexed by row-strict composition tableaux. The main purpose of defining these polynomials is to provide a combinatorial framework to study the cohomology ring in the following sections. Indeed, the polynomials defined below will serve as an equivariant generalization of the Springer monomial basis.

Definition 3.8.

Let be a composition of and . If , let be the polynomial of degree defined by,

If the inversion set of is empty, then define . We call the collection of polynomials obtained in this way equivariant Springer monomials.

While the are not monomials in the traditional sense, we use the term “monomial” since is a product of equivariant factors , a common generalization of monomials in ordinary cohomology. We adopt the convention throughout this manuscript that each equivariant Springer monomial and its image under the canonical projection map are denoted by the same symbol. This greatly simplifies the notation below.

Example 3.9.

Let , , and as in Example 3.3. Then

There is a simple inductive description of the equivariant Springer monomials, as explained in the next two paragraphs. Suppose labels a box in row of and recall that must label the last box in row . Since is the smallest label that appears, we have

Denote this set by . Note in particular that is uniquely determined by the value of .

Recall the map from Equation 3.1 defined by deleting the box labeled by in . The Springer inversions of decompose as

We obtain a corresponding decomposition formula for the polynomial given by

where

if and otherwise. The next lemma shows that the decomposition formula for the polynomials from Equation 3.3 is compatible the recursive formula defining the Springer monomials given in equation Equation 2.7.

Lemma 3.10.

Let be a composition of and denote its underlying partition shape. Then,

In particular, the set only depends on , the underlying partition shape of .

Proof.

First observe that if is a weak composition of , then determines a unique strong composition obtained by deleting the parts of equal to 0. If , then one obtains a unique element by upward justifying all rows. It is easy to see from the definitions that and have the same number of Springer inversions and that . Hence we may assume without loss of generality that is a strong composition of .

We now proceed by (reverse) induction on , the smallest value appearing in any . If then and . In this case, contains a single element, namely the row-strict composition tableau consisting of a single box labeled by . Therefore and , as desired.

Now suppose and has non-zero parts. Let denote the unique minimal length permutation of such that . In other words, is equal to the number of such that plus the number of such that and . Combining this notation with Equation 3.2 and Equation 3.3 implies that if with then .

Let be the composition of obtained from by decreasing by . Note that has content . Since the map from Equation 3.1 is a bijection, the decomposition of given in Equation 3.3 now gives us,

By the induction hypothesis, and our claim now follows directly from the recursive definition of given in Equation 2.7.

Example 3.11.

Consider the following tableaux in for and , respectively.

Here we have

While these polynomials are different, they correspond to the same Springer monomial, as .

The next theorem tells us that the collection of equivariant Springer monomials is an -module basis for the equivariant cohomology ring . We study the structure coefficients of with respect to this basis in the next section.

Theorem 3.12.

Let be a (strong) composition of . The collection of equivariant Springer monomials is an -module basis of .

Proof.

The polynomials are homogeneous elements of . Lemmas 2.3 and 3.10 now imply the desired result.

4. Localization and determinant formulas

In this section, we explore algebraic properties of the equivariant Springer monomials. The results of this section establish methods for computing the expansion of any as an -linear combination of the , . We begin by showing that the equivariant Springer monomials satisfy upper triangular vanishing relations with respect to the total ordering defined on row-strict composition tableaux defined in the previous section. We then use these vanishing properties to give a determinant formula for the structure coefficients in Theorem 4.5 below.

4.1. Localization formulas

Suppose is a strong composition of . Let be a regular element of , which we identify as a point in , by

The condition that be a regular element means that each of the are distinct. For every , there is a natural localization map,

given by . In other words, for any . Here acts on (and the coordinates of ) by permuting the entries; for example, if and then .

It is easy to see that if and only if for all . Hence any is uniquely determined by the collection of values . Recall that is the Levi subgroup of determined by the composition and denotes the Weyl group of . Since is standard, the parabolic subgroup is generated by a subset of simple reflections. Also, since acts trivially on (because ), it suffices to consider the maps where . Here denotes the set of minimal length coset representatives of . Recall that each permutation can be written uniquely as for and .

We now associate a coset representative to each by constructing a vector which is a particular permutation of the coordinates of . Specifically, if lies in the -th row of , then we require the -th coordinate of equal to . Let to be the unique permutation in such that . Observe that the map from to given by is a bijection.

Example 4.1.

Let and with given by:

Then with (in one-line notation). Note that in this case, and it easy to check that ; we have only to observe that and . Also, in this example we have since .

Our next proposition says that the equivariant Springer monomials satisfy upper triangular vanishing conditions with respect to the total order on row-strict composition tableaux defined in the previous section.

Proposition 4.2.

Let . Then the following are true:

(1)

, and

(2)

Proof.

Fix a regular element as in Equation 4.1. We first prove part (1) of the proposition. By definition, if is the -th coordinate of , then is contained in the -th row of . We have

Note that if , then cannot be contained in the -th row of . Hence for all and as claimed.

We now prove part (2). Indeed, we have

Since , there exists such that the content of -th row of contains . This implies that and hence . Since is an arbitrary regular element, we have in .

Remark 4.3.

A alternative proof of Theorem 3.12 from the previous section can be given using Proposition 4.2 as follows. Note that one can establish the fact that is an -linearly independent set by using the vanishing conditions of Proposition 4.2. Furthermore, the number of polynomials in of degree is precisely by Lemma 3.10. Thus is an -basis of by Proposition 18 of Reference 21.

We conclude with a detailed example.

Example 4.4.

Let and . A table of , and for all elements is displayed in Figure 1. The matrix written with respect to the total ordering on given in Example 3.7 is:

Proposition 4.2 implies this matrix is always upper triangular with respect to the total ordering in Definition 3.6 with non-vanishing polynomials in on the diagonal.

4.2. Structure constants for the equivariant Springer monomials

We now present a determinant formula for calculating the structure coefficients of the expansion of in the basis of equivariant Springer monomials. For these calculations, we work in the algebra

where the denotes the field of fractions of . We index the set

by the total ordering given in Definition 3.6 where .

For notational and computational simplicity, let and . Given any , we write

for some coefficients . Define vectors

and the matrix

Note that was computed for and in Example 4.4. Equation Equation 4.3 implies . Proposition 4.2 tells us that is an upper triangular matrix with nonzero diagonal entries, and is therefore invertible as a matrix with entries in . Hence

Our next theorem uses this equation to prove that each coefficient is the determinant of some matrix with entries determined by and . Normalize the polynomials by defining Note that this definition makes sense, since for all by Proposition 4.2.

Theorem 4.5.

Suppose and define

Write

Then . In particular, the coefficients for appearing in Equation 4.3 are

for all .

Proof.

Define the matrix and let denote the submatrix of obtained by removing the -th row and -th column. Applying Proposition 4.2, we observe that for all and if . This implies and

for . We prove the theorem by induction on . When , applying to both sides of Equation Equation 4.5 gives , as desired. Now suppose for all , we have that

We now apply the localization map to both sides of Equation Equation 4.5. Solving for and applying Equations Equation 4.6 and Equation 4.7 yields

proving the theorem.

This theorem provides us with the computational tools to expand any polynomial of in the basis of equivariant Springer monomials. It follows immediately that we can compute the expansion of any polynomial in in the Springer monomial basis by simply applying the evaluation map . We use these results in the next section to study the images of monomials and Schubert polynomials in .

Example 4.6.

Let and . The polynomials for are computed in Example 4.4 (see Figure 1 also). In this case and the total order on is as in Example 3.7, so the rows of the table in Figure 1 list the polynomials in order: , from top to bottom. We compute the expansion of

using the determinant formula of Theorem 4.5. The reader may note that is the image of the double Schubert polynomial under the map . The matrix from Theorem 4.5 is given by:

Where the first row is the vector with the rest of the matrix coming from first five rows of the matrix in Example 4.4 (normalized to ). If , then Theorem 4.5 says the coefficients are given by the upper-left minors (with a sign) yielding:

This implies

and hence Note that we can also compute by using the equation with

the inverse of the matrix from Example 4.4.

Remark 4.7.

If is the image of a double Schubert polynomial , then the vector can be computed directly using Billey’s localization formula (also called the Andersen–Jantzen–Soergel formula) given in Reference 3, Theorem 3.

5. Monomials and Schubert polynomials

In this section, we study the images of the Schubert polynomials under the map . We use Theorem 4.5 to identify an explicit collection of permutations for which the set is a basis of . This result is stated in Theorem 5.9. We obtain an analogous statement for equivariant cohomology in Corollary 5.14. Our analysis generalizes work of Harada–Tymoczko Reference 22 and Harada–Dewitt Reference 9 in the sense that Corollary 5.14 implies the existence of an explicit module basis for constructed by playing poset pinball.

We prove Theorem 5.9 in two steps. First, we use the expansion formula of Theorem 4.5 to prove that the Springer monomial basis of defined in Equation 2.7 above is upper-triangular in an appropriate sense. In particular, we study the expansion of any monomial in with respect to the Springer monomial basis. Since each Schubert polynomial is a sum of monomials, we are then able to leverage our results for monomials to prove the desired result for Schubert polynomials. More specifically, we prove that the transition matrix from to is invertible.

To begin, recall the commutative diagram from Equation 2.6. In particular, recall that and and the maps and denote the canonical projection maps.

5.1. Monomials

The first class of polynomials we study are monomials in the ring . Monomials in are indexed by weak compositions under the exponent identification

We impose the lexicographical total ordering on monomials. In other words, if and only if where denotes the smallest index where the entries of the compositions and differ.

If , then is a monomial in . Hence we define the notation

By Lemma 3.10 the set of all monomials obtained in this way is precisely the set of Springer monomials where is the underlying partition shape of . Recall that, by convention, since we also write to denote the image of the monomial in under . Observe that if is the associated exponent composition of , then is simply the number of inversions in whose first factor is . Hence we will call the composition the inversion vector of . For as in Example 3.3, the inversion vector is and . The next lemma follows immediately from the definition of the total order on .

Lemma 5.1.

Let . Then as row strict composition tableaux (c.f. Definition 3.6) if and only if as monomials in .

Lemma 5.1 implies that the vanishing property given in Proposition 4.2 is, in some way, compatible with the total ordering on all monomials. To make this compatibility precise, for each monomial we construct a polynomial such that . This polynomial serves as an analogue of for when .

Let be a composition of . If is the inversion vector for some , then set . Otherwise, by Lemma 5.1 there is a unique maximal such that . Let denote the inversion vector of . By definition of the total ordering on monomials, there exists an index such that if and . Let . We define the polynomial by

where denotes the index of the row containing in and the composition is defined by

The following example illustrates the construction.

Example 5.2.

Let and . Consider the monomial with . The maximal with is

with and . Note that the compositions and agree in the first four entries with so and in this case. We have , so

The next lemma is a technical result proving the key computational properties of .

Lemma 5.3.

Let . Then we obtain the following:

(1)

, and

(2)

for all such that , where is the localization map defined in Equation 4.2 above.

Proof.

It easy to see by construction that which proves (1). If is the inversion vector for some , then (2) is an immediate consequence of Proposition 4.2. Thus we have only to prove (2) in the case that is not the inversion vector for some row strict composition tableau. Let be the maximal element of such that .

First observe that if , then since the factor in evaluates to zero on any with . Now suppose . By definition of the total order on , there exists such that the content of the -th row of contains . Furthermore, we have that the numbers appear in the same rows (and the same exact position) of and . If , i.e. if , then . Otherwise, if then the tableaux and must contain in the same row. This implies due to the factor again evaluating to zero.

The following proposition tells us that the expansion of in the Springer monomial basis contains only monomials for such that . This is what we mean when we say that the Springer monomial basis is compatible with the total ordering on all monomials. Note that the proposition is also true if we impose the graded lexicographical order on monomials since is a graded map.

Proposition 5.4.

Let . Then

where the sum is over all such that . In other words, if , then .

Proof.

Let and note that if for some , then the proposition is trivial. We therefore assume that for any , i.e., that is not the inversion vector for any row strict composition tableaux of shape . Consider the polynomial as defined in equation Equation 5.1 and write

Let be the unique maximal tableau for which . Theorem 4.5 and Lemma 5.3 together imply for all . (Note that this fact also follows from equation Equation 4.4). Again by Lemma 5.3, we have and hence for all .

We demonstrate Proposition 5.4 with an example.

Example 5.5.

Let , , and . The and . The tableaux

is the unique maximal element of such that . In this case, we have . If we write , then the coefficients can be computed using Theorem 4.5. The nonzero coefficients are listed in the table appearing in Figure 2. From this information, we immediately get that

If we label with respect to the total order, then and the set of tableaux corresponding to nonzero coefficients are:

The underlined tableaux correspond to nonzero constant coefficients.

One immediate consequence of Proposition 5.4 is the following.

Corollary 5.6.

Let and let such that . Write

If and , then and .

5.2. Schubert polynomials

The set of Schubert polynomials in is an important collection of polynomials. Note that the map factors through where (see Theorem 2.1) and hence may be viewed as a polynomial in . It is widely known that Schubert polynomials are representatives for the Schubert classes in and form a basis of the cohomology ring. The main result of this section is Theorem 5.9 which states there is a natural subset such that the set of images form a basis for the cohomology of the Springer fiber . Corollary 5.14 in this section proves an equivariant version of this statement and generalizes results of Harada–Tymoczko Reference 22 and Harada–Dewitt Reference 9.

Given a permutation , we recall that the inversion set of is

Recall that the length of a permutation is . The Lehmer code of is defined as the sequence where denotes the number of inversions of of the form for some . Given any permutation and , it is clear that . On the other hand, given a sequence of nonnegative integers such that , the following well known lemma defines an explicit permutation with Lehmer code . See, for example, Reference 5, Ch. 2 for a proof.

Lemma 5.7.

Suppose is a sequence of nonnegative integers such that for . For each such , define

if , and if . Then has Lehmer code , and is unique with respect to this property.

We now describe the set . This subset is analogous to the set of Schubert points defined by the first author and Tymoczko in Reference 32, although our conventions differ, as discussed in Remark 3.4 above. To any we define to be the unique permutation (as defined in Lemma 5.7) such that the inversion vector of equals the Lehmer code of . Define

This collection of permutations has the property that the number of with Bruhat length is precisely . Thus the set is the output of a successful game of Betti pinball in the sense of Reference 22.

Example 5.8.

Let and . Take to be as in Example 5.2, and recall that has exponent vector . Applying Lemma 5.7 we have (where , , ). The Lehmer code of is .

We can now state the main theorem of this section.

Theorem 5.9.

The set forms an additive basis of .

Before we prove the theorem, we review the definition Schubert polynomials given by Lascoux and Schützenberger in Reference 27 and prove a key property about their monomial expansions. First recall Newton’s divided difference operator defined as:

where is the polynomial obtained by swapping the variables and in . The Schubert polynomials are defined recursively by first setting

where denotes the longest permutation in and then defining

if . Since the divided difference operators satisfy the braid relations on , Equation 5.2 is well defined. It was proved separately by Billey, Jockusch and Stanley in Reference 4, and Fomin and Stanley in Reference 13, that Schubert polynomials are nonnegative sums of monomials. For more details on Schubert polynomials and their properties, see Reference 28Reference 29.

Example 5.10.

Let and . We have

Note that is the minimal term appearing in the expansion above with respect to our monomial ordering, and where is the Lehmer code of .

As noted in the example above, the smallest monomial term (with respect to the lexicographical order) appearing in is the monomial , where is a Lehmer code of . We now prove that this property is true for all Schubert polynomials.

Lemma 5.11.

Let and denote the Lehmer code of . Then the Schubert polynomial has the expansion:

where implies that .

Proof.

We proceed by induction on the Lehmer code of , which we interpret as the exponent vector of a monomial. In particular, we induct on degree (i.e. the number of inversions of ) and use the converse of lexicographical order to induct on the Lehmer codes of a given degree. When then so and the desired expansion of holds trivially in this case.

We now assume and that there is an expansion of the form Equation 5.3 for every Schubert polynomial with or and such that the Lehmer code of is greater than that of .

For any , let denote the transposition which swaps and . Monk’s formula for Schubert polynomials implies that for any and , we have

Equation Equation 5.4 appears in Reference 28, Equation (4.15’) and in Reference 29, Exercise 2.7.3.

Let denote the smallest value for which and denote the unique inversion such that . Note that such an inversion exists as for all by our assumptions. In particular, this implies and . Define and (if , then we disregard ). In particular, note that and . Furthermore, our choice of implies that is the unique permutation such that with and . Applying Equation Equation 5.4 with and now gives us

where the sum is taken over all with , and .

Let and denote codes of and , respectively, and let denote the code of for some appearing in the sum. It is easy to check that

In particular, and .

We now have only to show that . To start, recall that with and . This implies . We furthermore know that , so . Thus . By construction, for all which implies for all . Since , we obtain

as desired. This proves for any with code appearing in the sum from Equation 5.5. The lemma now follows by induction.

Example 5.12.

In this example, we illustrate Equation Equation 5.5 from the proof of Lemma 5.11. Let . Then has Lehmer code . Using the notation in the proof of Lemma 5.11, and . We also have

In this case, the sum in Equation Equation 5.5 contains only one summand with yielding:

The codes of , and are respectively , and .

Remark 5.13.

Lemma 5.11 is analogous to Billey and Haiman’s Lemma 4.11 in Reference 2 which states that is the leading term (i.e. maximal monomial) in the expansion of when imposing reverse lexicographical order on the monomials. It should be noted that reverse lexicographical order is not the converse of lexicographical order, so Reference 2, Lemma 4.11 does not directly imply Lemma 5.11. However the proof of Lemma 5.11 given above is modeled after Mcdonald’s proof of Reference 2, Lemma 4.11 which appears in Reference 28, (4.16). The main difference in the proof Lemma 5.11 above is that we use the “smallest” inversion (i.e. in the proof of Lemma 5.11 we take to be the smallest value for which ) and not the “largest”. Observe that our argument using the “smallest” inversion, as seen in Equation Equation 5.5, does not yield a manifestly positive formula for the expansion of Schubert polynomials as a sum of monomials. However, the induction used to prove Reference 28, (4.16) does give a positive formula which is stated as a corollary in Reference 28, (4.19).

We can now prove our main theorem, which shows that the images of the Schubert polynomials corresponding to elements from form a basis of .

Proof of Theorem 5.9.

Let . Then there exists a unique for which . Let denote the Lehmer code of , which is also the inversion vector of . By Lemma 5.11, we can write

where the sum is over compositions and for all . We now have that

Since is the inversion vector of , we have . Furthermore, Proposition 5.4 implies

where the sum is over where and hence we can write

for some coefficients . This equation implies that the transition matrix from the set to the basis of is invertible. In fact, it is upper triangular with ’s on the diagonal. This proves the theorem.

Let denote the double Schubert polynomial indexed by ; see Reference 29 for the definition. As a corollary of Theorem 5.9, we obtain the corresponding statement for equivariant cohomology.

Corollary 5.14.

The set forms an -module basis of the equivariant cohomology .

Proof.

By Theorem 5.9, the polynomials form a homogeneous basis of . Since , the result now follows from Lemma 2.3.

Example 5.15.

Let and . We calculate the image of each Schubert polynomial under . We first recall the set and corresponding Springer monomial basis of ; this data is displayed in the table below (c.f. Example 4.4).

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{SVG} \begin{tabular}{c|cccccc} $\Upsilon$ & \begin{ytableau} 3 & 1\\ 4 & 2\\ \end{ytableau}& \begin{ytableau} 4 & 1\\ 3 & 2\\ \end{ytableau}& \begin{ytableau} 2 & 1\\ 4 & 3\\ \end{ytableau}& \begin{ytableau} 3 & 2\\ 4 & 1\\ \end{ytableau}& \begin{ytableau} 4 & 2\\ 3 & 1\\ \end{ytableau}& \begin{ytableau} 4 & 3\\ 2 & 1\\ \end{ytableau}\\ \\ \hline\\ $\mathbf{x}^{\Upsilon}$ & $1$ &$x_3$ & $x_2$&$x_1$ & $x_1x_3$ & $x_1x_2$\\ \end{tabular} \end{SVG}

By degree considerations, it suffices to calculate for (if then ). We obtain the following; note that the last column records whether or not is an element of .

yes
yes
yes
no
no
yes
yes
yes
no

For each Schubert polynomial , we have underlined the minimal monomial , so is the Lehmer code of as in Lemma 5.11.

We make two observations from Example 5.15. The first is that the set does not uniquely satisfy the basis property from Theorem 5.9. In particular, and hence replacing with in also corresponds to a basis of . The second is that each polynomial is a non-negative sum of Springer monomials. This motivates the following question.

Question 5.16.

Let be a composition of and and write

Do we have for all ?

Note that negative terms can appear in the expansion formula for the image of an individual monomial , as seen in Example 5.5. Looking more closely at the calculation of in Example 5.15 we find that

Observe that while the monomial is not an element of . This example shows that although the structure constants are nonnegative in many examples, there is typically some cancellation to take into account. The answer to Question 5.16 is known to be ‘yes’ in the special case that is one of , , or . Note that when , the Springer fiber is the full flag variety, and we obtain a positive answer to Question 5.16 using the formulas from Reference 2Reference 13Reference 28 .

6. Connections with the geometry of Springer fibers

It is well known that the Schubert polynomial is a polynomial representative for the fundamental cohomology class of the Schubert variety where denotes the longest element of . It is therefore natural to ask if the polynomials represent a fundamental cohomology class of a subvariety in the Springer fiber . Unfortunately, due the fact that Springer fibers are usually singular, the classical notion of a fundamental cohomology class of a subvariety using Poincaré duality is not defined. However, the notion of a fundamental homology class of a subvariety is well defined (see Reference 7 or Reference 17, Appendix B).

We briefly recall the connections between homology classes and Schubert polynomials for the flag variety , which is smooth. For any subvariety , let denote the corresponding fundamental homology class in . Since is smooth, the Poincaré duality isomorphism implies that for each class , there exists a unique cohomology class for which

Here denotes the cap product with the top fundamental class . For the Schubert variety, the class can be represented by the Schubert polynomial using the Borel presentation of .

Recall that the inclusion induces a surjective map . Since we do not consider equivariant cohomology in this section, we denote will denote by just . We now give a geometric interpretation of the classes , which are represented by the polynomials in . Note that the following proposition is true for any subvariety of the flag variety with inclusion map (not just Springer fibers).

Proposition 6.1.

Let denote the fundamental cohomology class of the Schubert variety . Then

for generic .

Proof.

Let denote a desingularization of the Springer fiber . Note that the fact that such a resolution exists is due to Hironaka Reference 23. Let and consider the diagram

Since is a surjective, birational morphism, we must have that . This implies

which is the left hand side of Equation Equation 6.1.

Kleiman’s transversality theorem Reference 24 implies that for generic , the preimage is generically transverse. By Reference 12, Theorem 1.23, we have that as elements in the Chow ring of (graded by codimension). Since smooth pullback commutes with the cycle map from the Chow ring to cohomology Reference 16, Corollary 19.2, it follows that in (Note that the maps are proper morphisms and is a proper morphism between smooth varieties. Hence proper pushforward and smooth pullback are well defined on Chow groups/rings). We now have that

Again, by Kleiman’s transversality theorem, the varieties and are both generically reduced and of the same codimension. Since is birational, the varieties and are also of the same dimension and hence

which completes the proof.

Observe that the variety is simply the intersection . If we let denote the open Schubert cell, then it is known that for carefully chosen , the collection of nonempty intersections is an affine paving of Reference 33Reference 38. In these cases, the corresponding nonzero homology classes form a basis of . We remark that the generic condition of in Proposition 6.1 typically excludes any such that is an affine paving of . Indeed, otherwise the map would be an isomorphism and imply that the Poincaré polynomials of Springer fibers are palindromic, which is false in most cases.

One immediate consequence of Proposition 6.1 is that linear relations among the classes in translate to linear relations on in .

Corollary 6.2.

Let be generic and suppose that in for some coefficients . Then

in . In particular, if , then .

The converse of this statement is not true since is usually not an isomorphism.

Note that can be computed explicitly by expanding its polynomial representative in terms of the Springer monomial basis using Theorem 4.5. Hence Corollary 6.2 gives a combinatorially sufficient condition to determine if the homology classes for generic .

Acknowledgments

The authors are grateful to Alex Woo, Jim Carrell, Dave Anderson, Anand Patel, Prakash Belkale, Jeff Mermin, and Vasu Tewari for helpful conversations and feedback.

Figures

Figure 1.

Equivariant Springer monomials for and .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{SVG} \begin{tabular}{cccc} $\Upsilon$& $P_\Upsilon$ & $w_\Upsilon$ & $h_\Upsilon$ \\ \hline\multirow{2}{*}{\begin{ytableau} 3 & 1\\ 4 & 2\\ \end{ytableau}}& \multirow{2}{*}{1} & \multirow{2}{*}{$[1,3,2,4]$} \,& \multirow{2}{*}{$(h_1, h_2, h_1, h_2)$}\\ & & & \\ \multirow{2}{*}{\begin{ytableau} 4 & 1\\ 3 & 2\\ \end{ytableau}}& \multirow{2}{*}{$x_3-z_1$} & \multirow{2}{*}{$[1,4,2,3]$} \, & \multirow{2}{*}{$(h_1, h_2, h_2, h_1)$}\\ & & & \\ \multirow{2}{*}{\begin{ytableau} 2 & 1\\ 4 & 3\\ \end{ytableau}}& \multirow{2}{*}{$x_2-z_2$} & \multirow{2}{*}{$[1,2,3,4]$} \, & \multirow{2}{*}{$(h_1, h_1, h_2, h_2)$}\\ & & & \\ \multirow{2}{*}{\begin{ytableau} 3 & 2\\ 4 & 1\\ \end{ytableau}}& \multirow{2}{*}{$x_1-z_1$} & \multirow{2}{*}{$[2,3,1,4]$} \,& \multirow{2}{*}{$(h_2, h_1, h_1, h_2)$}\\ & & & \\ \multirow{2}{*}{\begin{ytableau} 4 & 2\\ 3 & 1\\ \end{ytableau}} & \,\multirow{2}{*}{ $(x_1-z_1)(x_3-z_1)$} \, \, & \multirow{2}{*}{$[2,4,1,3]$} \,& \multirow{2}{*}{$(h_2, h_1, h_2, h_1)$}\\ & & & \\ \multirow{2}{*}{\begin{ytableau} 4 & 3\\ 2 & 1\\ \end{ytableau}}& \multirow{2}{*}{$(x_1-z_1)(x_2-z_1)$} & \multirow{2}{*}{$[3,4,1,2]$} \, & \multirow{2}{*}{$(h_2, h_2, h_1, h_1)$}\\ &&& \end{tabular} \end{SVG}
Figure 2.

Coefficients of for , , and .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{SVG} \begin{tabular}{c|ccccccc} $\Upsilon'$ & \begin{ytableau} 5 & 3 & 2\\ 6 & 4 & 1\\ \end{ytableau}& \begin{ytableau} 6 & 3 & 2\\ 5 & 4 & 1\\ \end{ytableau}& \begin{ytableau} 5 & 4 & 2\\ 6 & 3 & 1\\ \end{ytableau}& \begin{ytableau} 6 & 4 & 2\\ 5 & 3 & 1\\ \end{ytableau}& \begin{ytableau} 5 & 4 & 3\\ 6 & 2 & 1\\ \end{ytableau}& \begin{ytableau} 6 & 4 & 3\\ 5 & 2 & 1\\ \end{ytableau}& \begin{ytableau} 6 & 5 & 4\\ 3 & 2 & 1\\ \end{ytableau}\\ \\ \hline\\ $\mathbf{x}^{\Upsilon'}$& $x_1$& $x_1x_5$ & $x_1x_3$ & $x_1x_3x_5$ & $x_1x_2$ & $x_1x_2x_5$ & $x_1x_2x_3$ \\ \\ $C_{\Upsilon'}$ & $-(z_2-z_1)^2$& $(z_2-z_1)$ & $(z_2-z_1)$ & $-1$ & $(z_2-z_1)$ & $-1$ & $-1$ \\ \end{tabular} \end{SVG}

Mathematical Fragments

Equation (2.1)
Equation (2.2)
Equation (2.3)
Theorem 2.1 (Kumar–Procesi).

The kernel of the map defined in Equation 2.2 is the ideal . In particular, induces a graded -algebra isomorphism

making the following diagram commute.

Furthermore, the map naturally descends to a -algebra isomorphism:

with the following commutative diagram.

Remark 2.2.

It is well-known that the cohomology is concentrated in even degrees Reference 34. Thus the equivariant cohomology is a free -module, and isomorphic to the tensor product,

The graded -algebra isomorphism of Theorem 2.1 implies is a free -module with rank equal to the number of -fixed points of , namely (c.f. Reference 26, Lemma 2.1).

Equation (2.5)
Equation (2.6)
Lemma 2.3.

Suppose is a homogeneous basis of and let be a set of homogeneous polynomials in such that . Then is an -module basis of .

Equation (2.7)
Equation (3.1)
Example 3.3.

Let and . Consider with content :

The inversions of are .

Remark 3.4.

Note that the definition above is closely related to the notion of a Springer dimension pair considered by the first author and Tymoczko in Reference 32. In that paper, the convention is that the row-strict tableaux have increasing entries (from left to right), while our convention is that the entries are decreasing (from left to right). This change in conventions is routine; to convert from one to the other, apply the permutation such that for all . A Springer inversion from this paper corresponds to a unique Springer dimension pair as defined in Reference 32 (up to transformation under ). If is a Springer inversion then is a Springer dimension pair, where denotes the smallest element in row such that .

Lemma 3.5.

Let . Let denote the indices of the rows containing in and , respectively. Then exactly one of the following is true:

(1)

(2)

(3)

.

Definition 3.6.

Let and denote the indices of the rows containing in and , respectively. First suppose . We say if and if . Otherwise, if , then and have the same composition shape. In this case, we inductively say if .

Example 3.7.

Let and . The total order on tableaux in is displayed below.

Equation (3.2)
Equation (3.3)
Lemma 3.10.

Let be a composition of and denote its underlying partition shape. Then,

In particular, the set only depends on , the underlying partition shape of .

Theorem 3.12.

Let be a (strong) composition of . The collection of equivariant Springer monomials is an -module basis of .

Equation (4.1)
Equation (4.2)
Proposition 4.2.

Let . Then the following are true:

(1)

, and

(2)

Example 4.4.

Let and . A table of , and for all elements is displayed in Figure 1. The matrix written with respect to the total ordering on given in Example 3.7 is:

Proposition 4.2 implies this matrix is always upper triangular with respect to the total ordering in Definition 3.6 with non-vanishing polynomials in on the diagonal.

Equation (4.3)
Equation (4.4)
Theorem 4.5.

Suppose and define

Write

Then . In particular, the coefficients for appearing in Equation 4.3 are

for all .

Equation (4.6)
Equation (4.7)
Lemma 5.1.

Let . Then as row strict composition tableaux (c.f. Definition 3.6) if and only if as monomials in .

Equation (5.1)
Example 5.2.

Let and . Consider the monomial with . The maximal with is

with and . Note that the compositions and agree in the first four entries with so and in this case. We have , so

Lemma 5.3.

Let . Then we obtain the following:

(1)

, and

(2)

for all such that , where is the localization map defined in Equation 4.2 above.

Proposition 5.4.

Let . Then

where the sum is over all such that . In other words, if , then .

Example 5.5.

Let , , and . The and . The tableaux

is the unique maximal element of such that . In this case, we have . If we write , then the coefficients can be computed using Theorem 4.5. The nonzero coefficients are listed in the table appearing in Figure 2. From this information, we immediately get that

If we label with respect to the total order, then and the set of tableaux corresponding to nonzero coefficients are:

The underlined tableaux correspond to nonzero constant coefficients.

Lemma 5.7.

Suppose is a sequence of nonnegative integers such that for . For each such , define

if , and if . Then has Lehmer code , and is unique with respect to this property.

Theorem 5.9.

The set forms an additive basis of .

Equation (5.2)
Lemma 5.11.

Let and denote the Lehmer code of . Then the Schubert polynomial has the expansion:

where implies that .

Equation (5.4)
Equation (5.5)
Corollary 5.14.

The set forms an -module basis of the equivariant cohomology .

Example 5.15.

Let and . We calculate the image of each Schubert polynomial under . We first recall the set and corresponding Springer monomial basis of ; this data is displayed in the table below (c.f. Example 4.4).

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{SVG} \begin{tabular}{c|cccccc} $\Upsilon$ & \begin{ytableau} 3 & 1\\ 4 & 2\\ \end{ytableau}& \begin{ytableau} 4 & 1\\ 3 & 2\\ \end{ytableau}& \begin{ytableau} 2 & 1\\ 4 & 3\\ \end{ytableau}& \begin{ytableau} 3 & 2\\ 4 & 1\\ \end{ytableau}& \begin{ytableau} 4 & 2\\ 3 & 1\\ \end{ytableau}& \begin{ytableau} 4 & 3\\ 2 & 1\\ \end{ytableau}\\ \\ \hline\\ $\mathbf{x}^{\Upsilon}$ & $1$ &$x_3$ & $x_2$&$x_1$ & $x_1x_3$ & $x_1x_2$\\ \end{tabular} \end{SVG}

By degree considerations, it suffices to calculate for (if then ). We obtain the following; note that the last column records whether or not is an element of .

yes
yes
yes
no
no
yes
yes
yes
no

For each Schubert polynomial , we have underlined the minimal monomial , so is the Lehmer code of as in Lemma 5.11.

Question 5.16.

Let be a composition of and and write

Do we have for all ?

Proposition 6.1.

Let denote the fundamental cohomology class of the Schubert variety . Then

for generic .

Corollary 6.2.

Let be generic and suppose that in for some coefficients . Then

in . In particular, if , then .

References

Reference [1]
Hiraku Abe and Tatsuya Horiguchi, The torus equivariant cohomology rings of Springer varieties, Topology Appl. 208 (2016), 143–159, DOI 10.1016/j.topol.2016.05.004. MR3506975,
Show rawAMSref \bib{AH2016}{article}{ author={Abe, Hiraku}, author={Horiguchi, Tatsuya}, title={The torus equivariant cohomology rings of Springer varieties}, journal={Topology Appl.}, volume={208}, date={2016}, pages={143--159}, issn={0166-8641}, review={\MR {3506975}}, doi={10.1016/j.topol.2016.05.004}, }
Reference [2]
Sara Billey and Mark Haiman, Schubert polynomials for the classical groups, J. Amer. Math. Soc. 8 (1995), no. 2, 443–482, DOI 10.2307/2152823. MR1290232,
Show rawAMSref \bib{Billey-Haiman1995}{article}{ author={Billey, Sara}, author={Haiman, Mark}, title={Schubert polynomials for the classical groups}, journal={J. Amer. Math. Soc.}, volume={8}, date={1995}, number={2}, pages={443--482}, issn={0894-0347}, review={\MR {1290232}}, doi={10.2307/2152823}, }
Reference [3]
Sara C. Billey, Kostant polynomials and the cohomology ring for , Duke Math. J. 96 (1999), no. 1, 205–224, DOI 10.1215/S0012-7094-99-09606-0. MR1663931,
Show rawAMSref \bib{Bi96}{article}{ author={Billey, Sara C.}, title={Kostant polynomials and the cohomology ring for $G/B$}, journal={Duke Math. J.}, volume={96}, date={1999}, number={1}, pages={205--224}, issn={0012-7094}, review={\MR {1663931}}, doi={10.1215/S0012-7094-99-09606-0}, }
Reference [4]
Sara C. Billey, William Jockusch, and Richard P. Stanley, Some combinatorial properties of Schubert polynomials, J. Algebraic Combin. 2 (1993), no. 4, 345–374, DOI 10.1023/A:1022419800503. MR1241505,
Show rawAMSref \bib{BJS1993}{article}{ author={Billey, Sara C.}, author={Jockusch, William}, author={Stanley, Richard P.}, title={Some combinatorial properties of Schubert polynomials}, journal={J. Algebraic Combin.}, volume={2}, date={1993}, number={4}, pages={345--374}, issn={0925-9899}, review={\MR {1241505}}, doi={10.1023/A:1022419800503}, }
Reference [5]
Anders Björner and Francesco Brenti, Combinatorics of Coxeter groups, Graduate Texts in Mathematics, vol. 231, Springer, New York, 2005. MR2133266,
Show rawAMSref \bib{Bjorner-Brenti}{book}{ author={Bj\"{o}rner, Anders}, author={Brenti, Francesco}, title={Combinatorics of Coxeter groups}, series={Graduate Texts in Mathematics}, volume={231}, publisher={Springer, New York}, date={2005}, pages={xiv+363}, isbn={978-3540-442387}, isbn={3-540-44238-3}, review={\MR {2133266}}, }
Reference [6]
James B. Carrell, Orbits of the Weyl group and a theorem of DeConcini and Procesi, Compositio Math. 60 (1986), no. 1, 45–52. MR867955,
Show rawAMSref \bib{Ca86}{article}{ author={Carrell, James B.}, title={Orbits of the Weyl group and a theorem of DeConcini and Procesi}, journal={Compositio Math.}, volume={60}, date={1986}, number={1}, pages={45--52}, issn={0010-437X}, review={\MR {867955}}, }
Reference [7]
C. De Concini, G. Lusztig, and C. Procesi, Homology of the zero-set of a nilpotent vector field on a flag manifold, J. Amer. Math. Soc. 1 (1988), no. 1, 15–34, DOI 10.2307/1990965. MR924700,
Show rawAMSref \bib{DLP1988}{article}{ author={De Concini, C.}, author={Lusztig, G.}, author={Procesi, C.}, title={Homology of the zero-set of a nilpotent vector field on a flag manifold}, journal={J. Amer. Math. Soc.}, volume={1}, date={1988}, number={1}, pages={15--34}, issn={0894-0347}, review={\MR {924700}}, doi={10.2307/1990965}, }
Reference [8]
Corrado De Concini and Claudio Procesi, Symmetric functions, conjugacy classes and the flag variety, Invent. Math. 64 (1981), no. 2, 203–219, DOI 10.1007/BF01389168. MR629470,
Show rawAMSref \bib{DP81}{article}{ author={De Concini, Corrado}, author={Procesi, Claudio}, title={Symmetric functions, conjugacy classes and the flag variety}, journal={Invent. Math.}, volume={64}, date={1981}, number={2}, pages={203--219}, issn={0020-9910}, review={\MR {629470}}, doi={10.1007/BF01389168}, }
Reference [9]
Barry Dewitt and Megumi Harada, Poset pinball, highest forms, and Springer varieties, Electron. J. Combin. 19 (2012), no. 1, Paper 56, 35. MR2900431,
Show rawAMSref \bib{DH}{article}{ author={Dewitt, Barry}, author={Harada, Megumi}, title={Poset pinball, highest forms, and $(n-2,2)$ Springer varieties}, journal={Electron. J. Combin.}, volume={19}, date={2012}, number={1}, pages={Paper 56, 35}, review={\MR {2900431}}, }
Reference [10]
Elizabeth Drellich, Monk’s rule and Giambelli’s formula for Peterson varieties of all Lie types, J. Algebraic Combin. 41 (2015), no. 2, 539–575, DOI 10.1007/s10801-014-0545-2. MR3306081,
Show rawAMSref \bib{Drellich2015}{article}{ author={Drellich, Elizabeth}, title={Monk's rule and Giambelli's formula for Peterson varieties of all Lie types}, journal={J. Algebraic Combin.}, volume={41}, date={2015}, number={2}, pages={539--575}, issn={0925-9899}, review={\MR {3306081}}, doi={10.1007/s10801-014-0545-2}, }
Reference [11]
David S. Dummit and Richard M. Foote, Abstract algebra, 3rd ed., John Wiley & Sons, Inc., Hoboken, NJ, 2004. MR2286236,
Show rawAMSref \bib{DF}{book}{ author={Dummit, David S.}, author={Foote, Richard M.}, title={Abstract algebra}, edition={3}, publisher={John Wiley \& Sons, Inc., Hoboken, NJ}, date={2004}, pages={xii+932}, isbn={0-471-43334-9}, review={\MR {2286236}}, }
Reference [12]
David Eisenbud and Joe Harris, 3264 and all that—a second course in algebraic geometry, Cambridge University Press, Cambridge, 2016, DOI 10.1017/CBO9781139062046. MR3617981,
Show rawAMSref \bib{EH2016}{book}{ author={Eisenbud, David}, author={Harris, Joe}, title={3264 and all that---a second course in algebraic geometry}, publisher={Cambridge University Press, Cambridge}, date={2016}, pages={xiv+616}, isbn={978-1-107-60272-4}, isbn={978-1-107-01708-5}, review={\MR {3617981}}, doi={10.1017/CBO9781139062046}, }
Reference [13]
Sergey Fomin and Richard P. Stanley, Schubert polynomials and the nil-Coxeter algebra, Adv. Math. 103 (1994), no. 2, 196–207, DOI 10.1006/aima.1994.1009. MR1265793,
Show rawAMSref \bib{FoSt94}{article}{ author={Fomin, Sergey}, author={Stanley, Richard P.}, title={Schubert polynomials and the nil-Coxeter algebra}, journal={Adv. Math.}, volume={103}, date={1994}, number={2}, pages={196--207}, issn={0001-8708}, review={\MR {1265793}}, doi={10.1006/aima.1994.1009}, }
Reference [14]
Lucas Fresse, Betti numbers of Springer fibers in type , J. Algebra 322 (2009), no. 7, 2566–2579, DOI 10.1016/j.jalgebra.2009.07.008. MR2553695,
Show rawAMSref \bib{Fresse2009}{article}{ author={Fresse, Lucas}, title={Betti numbers of Springer fibers in type $A$}, journal={J. Algebra}, volume={322}, date={2009}, number={7}, pages={2566--2579}, issn={0021-8693}, review={\MR {2553695}}, doi={10.1016/j.jalgebra.2009.07.008}, }
Reference [15]
Lucas Fresse, Singular components of Springer fibers in the two-column case (English, with English and French summaries), Ann. Inst. Fourier (Grenoble) 59 (2009), no. 6, 2429–2444. MR2640925,
Show rawAMSref \bib{Fresse2009-2}{article}{ author={Fresse, Lucas}, title={Singular components of Springer fibers in the two-column case}, language={English, with English and French summaries}, journal={Ann. Inst. Fourier (Grenoble)}, volume={59}, date={2009}, number={6}, pages={2429--2444}, issn={0373-0956}, review={\MR {2640925}}, }
Reference [16]
William Fulton, Intersection theory, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 2, Springer-Verlag, Berlin, 1984, DOI 10.1007/978-3-662-02421-8. MR732620,
Show rawAMSref \bib{Fulton1984}{book}{ author={Fulton, William}, title={Intersection theory}, series={Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]}, volume={2}, publisher={Springer-Verlag, Berlin}, date={1984}, pages={xi+470}, isbn={3-540-12176-5}, review={\MR {732620}}, doi={10.1007/978-3-662-02421-8}, }
Reference [17]
William Fulton, Young tableaux: With applications to representation theory and geometry, London Mathematical Society Student Texts, vol. 35, Cambridge University Press, Cambridge, 1997. MR1464693,
Show rawAMSref \bib{Fulton1997}{book}{ author={Fulton, William}, title={Young tableaux}, subtitle={With applications to representation theory and geometry}, series={London Mathematical Society Student Texts}, volume={35}, publisher={Cambridge University Press, Cambridge}, date={1997}, pages={x+260}, isbn={0-521-56144-2}, isbn={0-521-56724-6}, review={\MR {1464693}}, }
Reference [18]
Francis Y. C. Fung, On the topology of components of some Springer fibers and their relation to Kazhdan-Lusztig theory, Adv. Math. 178 (2003), no. 2, 244–276, DOI 10.1016/S0001-8708(02)00072-5. MR1994220,
Show rawAMSref \bib{Fung2003}{article}{ author={Fung, Francis Y. C.}, title={On the topology of components of some Springer fibers and their relation to Kazhdan-Lusztig theory}, journal={Adv. Math.}, volume={178}, date={2003}, number={2}, pages={244--276}, issn={0001-8708}, review={\MR {1994220}}, doi={10.1016/S0001-8708(02)00072-5}, }
Reference [19]
A. M. Garsia and C. Procesi, On certain graded -modules and the -Kostka polynomials, Adv. Math. 94 (1992), no. 1, 82–138, DOI 10.1016/0001-8708(92)90034-I. MR1168926,
Show rawAMSref \bib{GP92}{article}{ author={Garsia, A. M.}, author={Procesi, C.}, title={On certain graded $S_n$-modules and the $q$-Kostka polynomials}, journal={Adv. Math.}, volume={94}, date={1992}, number={1}, pages={82--138}, issn={0001-8708}, review={\MR {1168926}}, doi={10.1016/0001-8708(92)90034-I}, }
Reference [20]
William Graham and R. Zierau, Smooth components of Springer fibers (English, with English and French summaries), Ann. Inst. Fourier (Grenoble) 61 (2011), no. 5, 2139–2182 (2012), DOI 10.5802/aif.2669. MR2961851,
Show rawAMSref \bib{Graham-Zierau2011}{article}{ author={Graham, William}, author={Zierau, R.}, title={Smooth components of Springer fibers}, language={English, with English and French summaries}, journal={Ann. Inst. Fourier (Grenoble)}, volume={61}, date={2011}, number={5}, pages={2139--2182 (2012)}, issn={0373-0956}, review={\MR {2961851}}, doi={10.5802/aif.2669}, }
Reference [21]
Megumi Harada and Julianna Tymoczko, A positive Monk formula in the -equivariant cohomology of type Peterson varieties, Proc. Lond. Math. Soc. (3) 103 (2011), no. 1, 40–72, DOI 10.1112/plms/pdq038. MR2812501,
Show rawAMSref \bib{HT11}{article}{ author={Harada, Megumi}, author={Tymoczko, Julianna}, title={A positive Monk formula in the $S^1$-equivariant cohomology of type $A$ Peterson varieties}, journal={Proc. Lond. Math. Soc. (3)}, volume={103}, date={2011}, number={1}, pages={40--72}, issn={0024-6115}, review={\MR {2812501}}, doi={10.1112/plms/pdq038}, }
Reference [22]
Megumi Harada and Julianna Tymoczko, Poset pinball, GKM-compatible subspaces, and Hessenberg varieties, J. Math. Soc. Japan 69 (2017), no. 3, 945–994, DOI 10.2969/jmsj/06930945. MR3685032,
Show rawAMSref \bib{HT17}{article}{ author={Harada, Megumi}, author={Tymoczko, Julianna}, title={Poset pinball, GKM-compatible subspaces, and Hessenberg varieties}, journal={J. Math. Soc. Japan}, volume={69}, date={2017}, number={3}, pages={945--994}, issn={0025-5645}, review={\MR {3685032}}, doi={10.2969/jmsj/06930945}, }
Reference [23]
Heisuke Hironaka, Resolution of singularities of an algebraic variety over a field of characteristic zero. I, II, Ann. of Math. (2) 79 (1964), 109–203; ibid. (2) 79 (1964), 205–326, DOI 10.2307/1970547. MR0199184,
Show rawAMSref \bib{Hironaka1964}{article}{ author={Hironaka, Heisuke}, title={Resolution of singularities of an algebraic variety over a field of characteristic zero. I, II}, journal={Ann. of Math. (2) {\bf 79} (1964), 109--203; ibid. (2)}, volume={79}, date={1964}, pages={205--326}, issn={0003-486X}, review={\MR {0199184}}, doi={10.2307/1970547}, }
Reference [24]
Steven L. Kleiman, The transversality of a general translate, Compositio Math. 28 (1974), 287–297. MR360616,
Show rawAMSref \bib{Kleiman1974}{article}{ author={Kleiman, Steven L.}, title={The transversality of a general translate}, journal={Compositio Math.}, volume={28}, date={1974}, pages={287--297}, issn={0010-437X}, review={\MR {360616}}, }
Reference [25]
Hanspeter Kraft, Conjugacy classes and Weyl group representations, Young tableaux and Schur functors in algebra and geometry (Toruń, 1980), Astérisque, vol. 87, Soc. Math. France, Paris, 1981, pp. 191–205. MR646820,
Show rawAMSref \bib{Kraft1981}{article}{ author={Kraft, Hanspeter}, title={Conjugacy classes and Weyl group representations}, conference={ title={Young tableaux and Schur functors in algebra and geometry (Toru\'{n}, 1980)}, }, book={ series={Ast\'{e}risque}, volume={87}, publisher={Soc. Math. France, Paris}, }, date={1981}, pages={191--205}, review={\MR {646820}}, }
Reference [26]
Shrawan Kumar and Claudio Procesi, An algebro-geometric realization of equivariant cohomology of some Springer fibers, J. Algebra 368 (2012), 70–74, DOI 10.1016/j.jalgebra.2012.06.019. MR2955222,
Show rawAMSref \bib{KP12}{article}{ author={Kumar, Shrawan}, author={Procesi, Claudio}, title={An algebro-geometric realization of equivariant cohomology of some Springer fibers}, journal={J. Algebra}, volume={368}, date={2012}, pages={70--74}, issn={0021-8693}, review={\MR {2955222}}, doi={10.1016/j.jalgebra.2012.06.019}, }
Reference [27]
Alain Lascoux and Marcel-Paul Schützenberger, Polynômes de Schubert (French, with English summary), C. R. Acad. Sci. Paris Sér. I Math. 294 (1982), no. 13, 447–450. MR660739,
Show rawAMSref \bib{LaSc81}{article}{ author={Lascoux, Alain}, author={Sch\"{u}tzenberger, Marcel-Paul}, title={Polyn\^{o}mes de Schubert}, language={French, with English summary}, journal={C. R. Acad. Sci. Paris S\'{e}r. I Math.}, volume={294}, date={1982}, number={13}, pages={447--450}, issn={0249-6291}, review={\MR {660739}}, }
Reference [28]
I. G. Macdonald, Schubert polynomials, Surveys in combinatorics, 1991 (Guildford, 1991), London Math. Soc. Lecture Note Ser., vol. 166, Cambridge Univ. Press, Cambridge, 1991, pp. 73–99. MR1161461,
Show rawAMSref \bib{McD91}{article}{ author={Macdonald, I. G.}, title={Schubert polynomials}, conference={ title={Surveys in combinatorics, 1991}, address={Guildford}, date={1991}, }, book={ series={London Math. Soc. Lecture Note Ser.}, volume={166}, publisher={Cambridge Univ. Press, Cambridge}, }, date={1991}, pages={73--99}, review={\MR {1161461}}, }
Reference [29]
Laurent Manivel, Symmetric functions, Schubert polynomials and degeneracy loci, SMF/AMS Texts and Monographs, vol. 6, American Mathematical Society, Providence, RI; Société Mathématique de France, Paris, 2001. Translated from the 1998 French original by John R. Swallow; Cours Spécialisés [Specialized Courses], 3. MR1852463,
Show rawAMSref \bib{Manivel}{book}{ author={Manivel, Laurent}, title={Symmetric functions, Schubert polynomials and degeneracy loci}, series={SMF/AMS Texts and Monographs}, volume={6}, note={Translated from the 1998 French original by John R. Swallow; Cours Sp\'{e}cialis\'{e}s [Specialized Courses], 3}, publisher={American Mathematical Society, Providence, RI; Soci\'{e}t\'{e} Math\'{e}matique de France, Paris}, date={2001}, pages={viii+167}, isbn={0-8218-2154-7}, review={\MR {1852463}}, }
Reference [30]
Aba Mbirika, A Hessenberg generalization of the Garsia-Procesi basis for the cohomology ring of Springer varieties, Electron. J. Combin. 17 (2010), no. 1, Research Paper 153, 29. MR2745706,
Show rawAMSref \bib{Mb10}{article}{ author={Mbirika, Aba}, title={A Hessenberg generalization of the Garsia-Procesi basis for the cohomology ring of Springer varieties}, journal={Electron. J. Combin.}, volume={17}, date={2010}, number={1}, pages={Research Paper 153, 29}, review={\MR {2745706}}, }
Reference [31]
Aba Mbirika and Julianna Tymoczko, Generalizing Tanisaki’s ideal via ideals of truncated symmetric functions, J. Algebraic Combin. 37 (2013), no. 1, 167–199, DOI 10.1007/s10801-012-0372-2. MR3016306,
Show rawAMSref \bib{MT13}{article}{ author={Mbirika, Aba}, author={Tymoczko, Julianna}, title={Generalizing Tanisaki's ideal via ideals of truncated symmetric functions}, journal={J. Algebraic Combin.}, volume={37}, date={2013}, number={1}, pages={167--199}, issn={0925-9899}, review={\MR {3016306}}, doi={10.1007/s10801-012-0372-2}, }
Reference [32]
Martha Precup and Julianna Tymoczko, Springer fibers and Schubert points, European J. Combin. 76 (2019), 10–26, DOI 10.1016/j.ejc.2018.08.010. MR3886508,
Show rawAMSref \bib{PT19}{article}{ author={Precup, Martha}, author={Tymoczko, Julianna}, title={Springer fibers and Schubert points}, journal={European J. Combin.}, volume={76}, date={2019}, pages={10--26}, issn={0195-6698}, review={\MR {3886508}}, doi={10.1016/j.ejc.2018.08.010}, }
Reference [33]
Naohisa Shimomura, A theorem on the fixed point set of a unipotent transformation on the flag manifold, J. Math. Soc. Japan 32 (1980), no. 1, 55–64, DOI 10.2969/jmsj/03210055. MR554515,
Show rawAMSref \bib{Shimomura1980}{article}{ author={Shimomura, Naohisa}, title={A theorem on the fixed point set of a unipotent transformation on the flag manifold}, journal={J. Math. Soc. Japan}, volume={32}, date={1980}, number={1}, pages={55--64}, issn={0025-5645}, review={\MR {554515}}, doi={10.2969/jmsj/03210055}, }
Reference [34]
N. Spaltenstein, The fixed point set of a unipotent transformation on the flag manifold, Nederl. Akad. Wetensch. Proc. Ser. A 79=Indag. Math. 38 (1976), no. 5, 452–456. MR0485901,
Show rawAMSref \bib{Spaltenstein1976}{article}{ author={Spaltenstein, N.}, title={The fixed point set of a unipotent transformation on the flag manifold}, journal={Nederl. Akad. Wetensch. Proc. Ser. A {\bf 79}=Indag. Math.}, volume={38}, date={1976}, number={5}, pages={452--456}, review={\MR {0485901}}, }
Reference [35]
T. A. Springer, Trigonometric sums, Green functions of finite groups and representations of Weyl groups, Invent. Math. 36 (1976), 173–207, DOI 10.1007/BF01390009. MR442103,
Show rawAMSref \bib{Springer1976}{article}{ author={Springer, T. A.}, title={Trigonometric sums, Green functions of finite groups and representations of Weyl groups}, journal={Invent. Math.}, volume={36}, date={1976}, pages={173--207}, issn={0020-9910}, review={\MR {442103}}, doi={10.1007/BF01390009}, }
Reference [36]
T. A. Springer, A construction of representations of Weyl groups, Invent. Math. 44 (1978), no. 3, 279–293, DOI 10.1007/BF01403165. MR491988,
Show rawAMSref \bib{Springer1978}{article}{ author={Springer, T. A.}, title={A construction of representations of Weyl groups}, journal={Invent. Math.}, volume={44}, date={1978}, number={3}, pages={279--293}, issn={0020-9910}, review={\MR {491988}}, doi={10.1007/BF01403165}, }
Reference [37]
Toshiyuki Tanisaki, Defining ideals of the closures of the conjugacy classes and representations of the Weyl groups, Tohoku Math. J. (2) 34 (1982), no. 4, 575–585, DOI 10.2748/tmj/1178229158. MR685425,
Show rawAMSref \bib{Tanisaki1982}{article}{ author={Tanisaki, Toshiyuki}, title={Defining ideals of the closures of the conjugacy classes and representations of the Weyl groups}, journal={Tohoku Math. J. (2)}, volume={34}, date={1982}, number={4}, pages={575--585}, issn={0040-8735}, review={\MR {685425}}, doi={10.2748/tmj/1178229158}, }
Reference [38]
Julianna S. Tymoczko, Linear conditions imposed on flag varieties, Amer. J. Math. 128 (2006), no. 6, 1587–1604. MR2275912,
Show rawAMSref \bib{Tymoczko2006}{article}{ author={Tymoczko, Julianna S.}, title={Linear conditions imposed on flag varieties}, journal={Amer. J. Math.}, volume={128}, date={2006}, number={6}, pages={1587--1604}, issn={0002-9327}, review={\MR {2275912}}, }

Article Information

MSC 2020
Primary: 05E10 (Combinatorial aspects of representation theory), 14M15 (Grassmannians, Schubert varieties, flag manifolds)
Secondary: 14N15 (Classical problems, Schubert calculus)
Author Information
Martha Precup
Department of Mathematics and Statistics, Washington University in St. Louis, St. Louis, Missouri 63130
martha.precup@wustl.edu
MathSciNet
Edward Richmond
Department of Mathematics, Oklahoma State University, Stillwater, Oklahoma , 74078
edward.richmond@okstate.edu
MathSciNet
Additional Notes

The first author was supported by an Oklahoma State University CAS summer research grant. The second author was partially supported by an AWM-NSF travel grant and NSF grant DMS 1954001 during the course of this research.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 8, Issue 17, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2021 by the authors under Creative Commons Attribution 3.0 License (CC BY 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/57
  • MathSciNet Review: 4273195
  • Show rawAMSref \bib{4273195}{article}{ author={Precup, Martha}, author={Richmond, Edward}, title={An equivariant basis for the cohomology of Springer fibers}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={8}, number={17}, date={2021}, pages={481-509}, issn={2330-0000}, review={4273195}, doi={10.1090/btran/57}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.