Skip to main content

Affine Schubert Calculus

  • Chapter
  • First Online:
  • 847 Accesses

Part of the book series: Fields Institute Monographs ((FIM,volume 33))

Abstract

This chapter discusses how k-Schur and dual k-Schur functions can be defined for all types. This is done via some combinatorial problems that come from the geometry of a very large family of generalized flag varieties. They apply to the expansion of products of Schur functions, k-Schur functions and their dual basis, and Schubert polynomials. Despite the geometric origin of these problems, concrete algebraic models will be given for the relevant cohomology rings and their Schubert bases.

This is a preview of subscription content, log in via an institution.

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   49.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD   59.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

Notes

  1. 1.

    The author was supported by NSF grants DMS–0652641, DMS–0652648, and DMS–1200804.

Bibliography

  1. H.H. Andersen, J.C. Jantzen, W. Soergel, Representations of quantum groups at a pth root of unity and of semisimple groups in characteristic p: independence of p. Astérisque 220, 321 (1994)

    Google Scholar 

  2. A. Arabia, Cycles de Schubert et cohomologie équivariante de KT. Invent. Math. 85(1), 39–52 (1986)

    Google Scholar 

  3. S. Billey, Kostant polynomials and the cohomology ring for GB. Duke Math. J. 96, 205–224 (1999)

    Google Scholar 

  4. S. Billey, M. Haiman, Schubert polynomials for the classical groups. J. Am. Math. Soc. 8(2), 443–482 (1995)

    Article  MATH  MathSciNet  Google Scholar 

  5. S. Fomin, S. Gelfand, A. Postnikov, Quantum Schubert polynomials. J. Am. Math. Soc. 10(3), 565–596 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  6. V. Ginzburg, Geometric methods in the representation theory of Hecke algebras and quantum groups. Notes by Vladimir Baranovsky, in Representation Theories and Algebraic Geometry, Montreal, PQ, 1997. NATO Advanced Science Institute Series C, Mathematical and Physical Sciences, vol. 514 (Kluwer, Dordrecht, 1998), pp. 127–183

    Google Scholar 

  7. M. Goresky, R. Kottwitz, R. MacPherson, Equivariant cohomology, Koszul duality, and the localization theorem. Invent. Math. 131(1), 25–83 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  8. M. Goresky, R. Kottwitz, R. MacPherson, Homology of affine Springer fibers in the unramified case. Duke Math. J. 121, 509–561 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  9. W. Graham, Positivity in equivariant Schubert calculus. Duke Math. J. 109(3), 599–614 (2001)

    Article  MATH  MathSciNet  Google Scholar 

  10. T. Ikeda, L. Mihalcea, H. Naruse, Double Schubert polynomials for the classical groups. Adv. Math. 226(1), 840–886 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  11. V.G. Kac, Infinite-Dimensional Lie Algebras, 3rd edn. (Cambridge University Press, Cambridge, 1990), pp. xxii+400. ISBN:0-521-37215-1

    Google Scholar 

  12. M. Kashiwara, The flag manifold of Kac-Moody Lie algebra, in Algebraic Analysis, Geometry, and Number Theory, Baltimore, 1988 (Johns Hopkins University Press, Baltimore, 1989), pp. 161–190

    Google Scholar 

  13. M. Kashiwara, M. Shimozono, Equivariant K-theory of affine flag manifolds and affine Grothendieck polynomials. Duke Math. J. 148(3), 501–538 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  14. B. Kostant, S. Kumar, The nil Hecke ring and cohomology of GP for a Kac–Moody group G. Adv. Math. 62(3), 187–237 (1986)

    Google Scholar 

  15. S. Kumar, Kac-Moody Groups, Their Flag Varieties and Representation Theory. Progress in Mathematics, vol. 204 (Birkhäuser, Boston, 2002), pp. xvi+606

    Google Scholar 

  16. T. Lam, Affine Stanley symmetric functions. Am. J. Math. 128(6), 1553–1586 (2006)

    Article  MATH  Google Scholar 

  17. T. Lam, Schubert polynomials for the affine Grassmannian. J. Am. Math. Soc. 21(1), 259–281 (2008)

    Article  MATH  Google Scholar 

  18. T. Lam, A. Schilling, M. Shimozono, Schubert polynomials for the affine Grassmannian of the symplectic group. Math. Z. 264, 765–811 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  19. T. Lam, M. Shimozono, Quantum cohomology of GP and homology of affine Grassmannian. Acta Math. 204, 49–90 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  20. L. Lapointe, J. Morse, A k-tableau characterization of k-Schur functions. Adv. Math. 213(1), 183–204 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  21. L. Lapointe, J. Morse, Quantum cohomology and the k-Schur basis. Trans. Am. Math. Soc. 360, 2021–2040 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  22. A. Lascoux, Classes de Chern des variétés de drapeaux. C. R. Acad. Sci. Paris Sér. I Math. 295(5), 393–398 (1982)

    MATH  MathSciNet  Google Scholar 

  23. A. Lascoux, M.-P. Schützenberger, Polynômes de Schubert. C. R. Acad. Sci. Paris Sér. I Math. 294(13), 447–450 (1982)

    MATH  Google Scholar 

  24. D. Peterson, Quantum cohomology of G/P. Lecture Notes (MIT, 1997)

    Google Scholar 

  25. S. Pon, Affine Stanley symmetric functions for classical types. J. Algebr. Comb. 36(4), 595–622 (2012). And Ph.D thesis, UC Davis, 2010

    Google Scholar 

  26. P. Pragacz, Algebro-geometric applications of Schur S- and Q-polynomials, in Topics in Invariant Theory, Paris, 1989/1990. Lecture Notes in Mathematics, vol. 1478 (Springer, Berlin, 1991), pp. 130–191

    Google Scholar 

  27. P. Pragacz, J. Ratajski, Formulas for Lagrangian and orthogonal degeneracy loci; \(\tilde{Q}\)-polynomial approach. Compositio Math. 107(1), 11–87 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  28. D. Quillen, unpublished

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

A Appendix: Proof of Coalgebra Properties

Let M and N be left F W -modules. Then M, N, MF N, and HomF(M, N) are left F-modules. We define an F W -module structure on MF N and HomF(M, N).

Proof of Proposition 3.14.

We first check the well-definedness of the formula for a ⋅ mn. Let a ∈ F W with \(\Delta (a) =\sum _{(a)}a_{(1)} \otimes a_{(2)}\). We further expand \(a_{(k)} =\sum _{w}a_{(k)}^{w}w\) for k = 1, 2 where \(a_{(k)}^{w} \in \mathrm{ F}\). The condition of membership in \(\mathrm{Im}(\Delta )\) of the right hand side of (3.33), is that only terms of the form vv survive:

$$\displaystyle{ \sum _{a}a_{(1)}^{v}a_{ (2)}^{w} = 0\qquad \mbox{ for all $v\neq w$.} }$$
(A.1)

We take a typical generator of the relations in M Q N: qmnmqn. We compute the componentwise action of \(\Delta (a)\) on qmn and mqn.

$$\displaystyle\begin{array}{rcl} \Delta (a) \cdot (qm \otimes n)& =& \sum _{a}a_{(1)} \cdot qm \otimes a_{(2)} \cdot n {}\\ & =& \sum _{a,v,w}a_{(1)}^{v}a_{ (2)}^{w}v \cdot qm \otimes w \cdot n {}\\ & =& \sum _{a,v,w}a_{(1)}^{v}a_{ (2)}^{w}(v \cdot q)(v \cdot m) \otimes w \cdot n {}\\ & =& \sum _{a,v,w}a_{(1)}^{v}a_{ (2)}^{w}(v \cdot q)(v \cdot m \otimes w \cdot n). {}\\ \end{array}$$

Similarly

$$\displaystyle\begin{array}{rcl} \Delta (a) \cdot (m \otimes qn) =\sum _{a,v,w}a_{(1)}^{v}a_{ (2)}^{w}(w \cdot q)(v \cdot m \otimes w \cdot n).& & {}\\ \end{array}$$

The difference of these two expressions is

$$\displaystyle\begin{array}{rcl} & & \quad \,\,\sum _{a}\sum _{v\neq w}a_{(1)}^{v}a_{ (2)}^{w}(v \cdot q - w \cdot q)(v \cdot m \otimes w \cdot n) {}\\ & & =\sum _{v\neq w}(v \cdot q - w \cdot q)(v \cdot m \otimes w \cdot n)\sum _{a}a_{(1)}^{v}a_{ (2)}^{w} {}\\ & & = 0. {}\\ \end{array}$$

Thus the formula for a ⋅ (mn) is well-defined.

Applying this to the special case of the action of F W on \(\mathrm{F}_{W} \otimes _{\mathrm{F}}\mathrm{F}_{W}\), we recover part (1), including (3.32). For a, b ∈ F W we have

$$\displaystyle\begin{array}{rcl} a \cdot (b \cdot (m \otimes n))& =& \sum _{(a)}\sum _{(b)}a_{(1)} \cdot (b_{(1)} \cdot m) \otimes a_{(2)} \cdot (b_{(2)} \cdot n) {}\\ & =& \sum _{(a)}\sum _{(b)}(a_{(1)}b_{(1)}) \cdot m \otimes (a_{(2)}b_{(2)}) \cdot n {}\\ & =& (ab) \cdot (m \otimes n) {}\\ \end{array}$$

where the last step holds because of (3.32). Hence we have an action of F W on MF N.

For the left F W -module M, the dual \({M}^{{\ast}} =\mathrm{ Hom}_{\mathrm{F}}(M,\mathrm{F})\) has a left F W -module structure defined by w ⋅ m  = m w −1 or more generally \(a \cdot {m}^{{\ast}} = {m}^{{\ast}}\circ {a}^{t}\) for w ∈ W and a ∈ F W . Consider the left F-linear isomorphism \({M}^{{\ast}}\otimes _{\mathrm{F}}N\mathop{\cong}\mathrm{Hom}_{\mathrm{F}}(M,N)\) given by \({m}^{{\ast}}\otimes n\mapsto (x\mapsto {m}^{{\ast}}(x)n)\). We define a left F W -module structure on HomF(M, N) by declaring that the above map is an isomorphism of left F W -modules. It is enough to consider the action of a = w: \(w \cdot {m}^{{\ast}}\otimes n = w \cdot {m}^{{\ast}}\otimes w \cdot n = {m}^{{\ast}}\circ {w}^{-1} \otimes w \cdot n\). This corresponds to the function \(x\mapsto {m}^{{\ast}}({w}^{-1}x)w \cdot n = w \cdot {m}^{{\ast}}({w}^{-1}x)n\). If f ∈ HomF(M, N) corresponds to m n then the above function corresponds to \(x\mapsto w \cdot f({w}^{-1}x)\), which is the required action.

Proof of Proposition 3.16.

We first compute

$$\displaystyle\begin{array}{rcl} \Delta (A_{i})& =& \Delta (\alpha _{i}^{-1}(1 - s_{ i})) {}\\ & =& \alpha _{i}^{-1}(1 \otimes 1 - s_{ i} \otimes s_{i}) {}\\ & =& \alpha _{i}^{-1}(1 \otimes 1 - s_{ i} \otimes 1 + s_{i} \otimes 1 - s_{i} \otimes s_{i}) {}\\ & =& A_{i} \otimes 1 + s_{i} \otimes A_{i} {}\\ & =& \alpha _{i}^{-1}(1 \otimes 1 - 1 \otimes s + 1 \otimes s_{ i} - s_{i} \otimes s_{i}) {}\\ & =& 1 \otimes A_{i} + A_{i} \otimes s_{i}. {}\\ \end{array}$$

This yields (3.35). It follows that the restriction of \(\Delta \) to \(\mathbb{A}\) has image in \(\mathbb{A} \otimes _{S}\mathbb{A}\) and inherits the required properties by Proposition 3.14. All other assertions follow directly.

B Appendix: Small Torus GKM Proofs

Proof.

We prove (1) as (2) follows from it. There is a commutative diagram of ring homomorphisms

(B.1)

The horizontal maps are restriction to torus-fixed points. \(\mbox{ For}_{T}^{T_{\mathrm{af}}}\) is the map that forgets from T af-equivariance to T-equivariance. The top map is an isomorphism by Theorem 3.45. It is not hard to show that For is surjective and that \({H}^{T}(\tilde{\mathrm{Fl}}_{\mathrm{af}})\) has an H T(pt)-basis given by the T-equivariant classes of Schubert varieties \({[{X}^{v}]}^{T} \in {H}^{T}(\tilde{\mathrm{Fl}}_{\mathrm{af}})\) for v ∈ W af. By commutativity of the diagram,

$$\displaystyle{ \mathrm{Im}(\mathrm{{res}}^{{\prime}}) =\mathrm{ Im}{(\pi }^{{\ast}}) =\bigoplus _{ v\in W_{\mathrm{af}}}{S\bar{\xi }}^{v}. }$$
(B.2)

The functions \(\bar{{\xi }}^{v}\) are independent since the matrix \((\pi (d_{v,w}))_{v,w\in W_{\mathrm{af}}}\) is triangular with nonvanishing diagonal entries.

It remains to show that

$$\displaystyle{ \mathrm{Im}{(\pi }^{{\ast}}) = \Lambda _{\mathrm{ af}}^{{\prime}}. }$$
(B.3)

Let v ∈ W af. Certainly \(\bar{{\xi }}^{v}\) satisfies (3.53) since ξ v does. We must check the conditions (4.1) and (4.2). Let w ∈ W af, \(\alpha \in \Phi \), and \(d \in \mathbb{Z}_{>0}\). Let W  ⊂ W af be the subgroup generated by \(t_{{\alpha }^{\vee }}\) and r α ; it is isomorphic to the affine Weyl group of SL 2. Define the function f: W  → S by \(f(x) {=\bar{\xi } }^{v}(xw)\). Since ξ v satisfies (3.53) for Flaf, f satisfies (3.53) for the affine flag variety Fl corresponding to α. It follows that f is an S-linear combination of Schubert classes in Fl. By Propositions B.1 and B.2 (proved below), πf satisfies (4.1) and (4.2), so that \(\bar{{\xi }}^{v} \in \Lambda _{\mathrm{af}}^{{\prime}}\), as required.

Conversely, suppose \(\xi \in \Lambda _{\mathrm{af}}^{{\prime}}\). We show that

$$\displaystyle{ \xi \in \bigoplus _{v\in W_{\mathrm{af}}}{S\,\bar{\xi }}^{v}. }$$
(B.4)

Let x = t λ u be of minimal length in the support of ξ, with u ∈ W and λ ∈ Q . It suffices to show that

$$\displaystyle{ \xi (x) {\in \bar{\xi }}^{x}(x)S. }$$
(B.5)

Suppose (B.5) holds. Define \({\xi }^{{\prime}}: W_{\mathrm{af}} \rightarrow S\) by \({\xi }^{{\prime}} =\xi -{(\xi (x){/\bar{\xi }}^{x}(x))\bar{\xi }}^{x}\). Since \(\Lambda _{\mathrm{af}}^{{\prime}}\) is an S-module, \({\xi }^{{\prime}} \in \Lambda _{\mathrm{af}}^{{\prime}}\). Moreover \(\Omega {(\xi }^{{\prime}}) \supsetneq \Omega (\xi )\) where \(\Omega (\xi )\) is defined by (3.64). By induction (B.4) holds for ξ and therefore it holds for ξ.

We now show (B.5). The elements \(\{\alpha \mid \alpha \in {\Phi }^{+}\}\) are relatively prime in S. Letting \(\alpha \in {\Phi }^{+}\), by (3.61) it suffices to show that \(\xi (x) \in J:{=\alpha }^{d}S\) where \(d = \vert \mathrm{Inv}_{\alpha }({x}^{-1})\vert \) and \(\mathrm{Inv}_{\alpha }({x}^{-1})\) is the set of roots in Inv(x −1) (see (2.21)) of the form ±α + k δ for some \(k \in \mathbb{Z}_{\geq 0}\).

Note that for \(\beta \in \Phi _{\mathrm{af}}^{+\mathrm{re}}\), β ∈ Inv(x −1) if and only if \({x}^{-1}\cdot \beta \in -\Phi _{\mathrm{af}}^{+\mathrm{re}}\). We have

$$\displaystyle{ {x}^{-1} \cdot (\pm \alpha + k\delta ) = {u}^{-1}t_{ -\lambda }\cdot (\pm \alpha + k\delta ) = \pm {u}^{-1}\alpha + (k\pm \langle \lambda \,,\,\alpha \rangle )\delta. }$$

Letting \(\chi =\chi (\alpha \in \mathrm{ Inv}({u}^{-1}))\) we have

$$\displaystyle{ \mathrm{Inv}_{\alpha }({x}^{-1}) = \left \{\begin{array}{@{}l@{\quad }l@{}} \{\alpha,\alpha +\delta,\mathop{\ldots },\alpha -(\langle \lambda \,,\,\alpha \rangle +1-\chi )\delta \} \quad &\mbox{ if $\langle \lambda \,,\,\alpha \rangle \leq 0$} \\ \{-\alpha +\delta,-\alpha + 2\delta,\mathop{\ldots },-\alpha + (\langle \lambda \,,\,\alpha \rangle -\chi )\delta \}\quad &\mbox{ if $\langle \lambda \,,\,\alpha \rangle > 0$.} \end{array} \right. }$$
(B.6)

Suppose first that \(\langle \lambda \,,\,\alpha \rangle > 0\). Then \(d =\langle \lambda \,,\,\alpha \rangle -\chi (\alpha \in \mathrm{ Inv}({u}^{-1}))\). Applying (4.2) to \(y = t_{{(1-d)\alpha }^{\vee }}x\), we have Z 1 ∈ J where

$$\displaystyle\begin{array}{rcl} Z_{1}& =& \xi ({(1 - t_{{\alpha }^{\vee }})}^{d-1}(1 - s_{\alpha })y) {}\\ & =& {(-1)}^{d-1}\xi ({(1 - t_{{ -\alpha }^{\vee }})}^{d-1}x) -\xi ({(1 - t_{{\alpha }^{\vee }})}^{d-1}s_{\alpha }y) {}\\ & =& {(-1)}^{d-1}\xi (x) -\xi ({(1 - t_{{\alpha }^{\vee }})}^{d-1}s_{\alpha }y) {}\\ \end{array}$$

where the last equality holds by the assumption on Supp(ξ) and a calculation of \(\mathrm{Inv}_{\alpha }({(s_{\alpha +d\delta }x)}^{-1})\), giving \(x > s_{\alpha +d\delta }x > t_{-{k\alpha }^{\vee }}x\) for all 1 ≤ k ≤ d − 1. By Lemma 4.1 we have Z 2 ∈ J where

$$\displaystyle{ Z_{2} =\xi ({(1 - t_{{\alpha }^{\vee }})}^{d-1}s_{\alpha }y) -\xi ({(1 - t_{{\alpha }^{\vee }})}^{d-1}t_{{ p\alpha }^{\vee }}s_{\alpha }y) }$$

for any \(p \in \mathbb{Z}\). Thus

$$\displaystyle{ Z_{1} + Z_{2} = {(-1)}^{d-1}\xi (x) -\xi ({(1 - t_{{\alpha }^{\vee }})}^{d-1}t_{{ p\alpha }^{\vee }}s_{\alpha }y) \in J. }$$

By the assumption on Supp(x) and (B.6),

$$\displaystyle{ \xi ({(1 - t_{{\alpha }^{\vee }})}^{d-1}t_{{ p\alpha }^{\vee }}s_{\alpha }y) = 0 }$$

for p = 2 − d. It follows that ξ(x) ∈ J.

Otherwise \(\langle \lambda \,,\,\alpha \rangle \leq 0\). By the previous case we may assume that \(t_{{d\alpha }^{\vee }}x\not\in \mathrm{Supp}(\xi )\). Thus

$$\displaystyle{ \xi (x) =\xi ({(1 - t_{{\alpha }^{\vee }})}^{d}x) \in J }$$

by induction on Supp(ξ) and (4.1). This proves (B.5) and (B.4) as required.

2.1 B.1 Small Torus GKM Condition for \(\widehat{\mathrm{sl}}_{2}\)

In this section we consider the root datum for \(\widehat{\mathrm{sl}}_{2}\). Let \({\Phi }^{+} =\{\alpha \}\) where α = α 1. Also δ = α 0 +α 1 so that π(α 0) = −π(α 1) and the level zero action of W af is given by s 0 ⋅ α = s 1 ⋅ α = −α. For \(i \in \mathbb{Z}_{\geq 0}\) let

$$\displaystyle{ \begin{array}{rcl} \sigma _{2i}& =&{(s_{1}s_{0})}^{i}\qquad \sigma _{-2i} = {(s_{0}s_{1})}^{i},\\ \sigma _{ 2i+1} & =&s_{0}\sigma _{2i}\qquad \sigma _{-(2i+1)} = s_{1}\sigma _{-2i}. \end{array} }$$
(B.7)

Then we have (σ j ) =  | j | for \(j \in \mathbb{Z}\), \(W_{\mathrm{af}}^{I} =\{\sigma _{j}\mid j \in \mathbb{Z}_{\geq 0}\}\), and

$$\displaystyle{ \sigma _{2i} = t_{-{i\alpha }^{\vee }}\qquad \mbox{ for $i \in \mathbb{Z}$.} }$$

Let \(\xi _{j}^{i}:=\pi {(\xi }^{\sigma _{i}}(\sigma _{j}))\) for \(i,j \in \mathbb{Z}\). For \(m,a \in \mathbb{Z}_{\geq 0}\),

$$\displaystyle{ \xi _{j}^{m} = {(-1)}^{mj}\binom{m + a}{{m}\alpha }^{m} }$$
(B.8)

where j ∈ { m + 2a, m + 2a + 1, −m − 2a − 1, −m − 2a − 2}. This is easily proved by induction using (3.60). The rest of the values for ξ m are zero by (3.58). For m < 0 we may use

$$\displaystyle{ \xi _{j}^{m} = {(-1)}^{m}\xi _{ -j}^{-m}\qquad \mbox{ for $m,j \in \mathbb{Z}$} }$$
(B.9)

which follows from the Dynkin symmetry \(0 \leftrightarrow 1\).

Proposition B.1.

For all d ≥ 1, \(m \in \mathbb{Z}\) , and w ∈ W af we have

$$\displaystyle{{ \xi }^{m}({(1 - t_{{\alpha }^{\vee }})}^{d}w) {\in \alpha }^{d}\mathbb{Z}[\alpha ]. }$$
(B.10)

Proof.

One may reduce to m ≥ 0 using (B.9). Equation (B.10) is proved for m = 2i, \(w = t_{{(-i-a)\alpha }^{\vee }}\) and d = (i + a) + (i + 1 + b) = 2i + a + b + 1 for \(a,b \in \mathbb{Z}_{\geq 0}\), as the other possibilities are similar or easier. Equation (B.10) can be rewritten as

$$\displaystyle{ Z:=\sum _{ k=0}^{d}{(-1)}^{k}\binom{d}{k}\xi _{ -2i-2-2b+2k}^{2i} {\in \alpha }^{d}\mathbb{Z}[\alpha ]. }$$
(B.11)

By (3.58) ξ 2p 2i = 0 for − 2i − 2 < 2p < 2i. Using (B.8),

$$\displaystyle\begin{array}{rcl} Z& =& \left (\sum _{k=0}^{b} +\sum _{ k=2i+1+b}^{d}\right ){(-1)}^{k}\binom{d}{k}\xi _{ -2i-2-2b+2k}^{2i} {}\\ & =& \sum _{k=0}^{b}{(-1)}^{k}\binom{d}{k}\xi _{ -2i-2-2b+2k}^{2i} +\sum _{ k=0}^{a}{(-1)}^{k+2i+1+b}\binom{d}{2i + 1 + b + k}\xi _{ 2i+2k}^{2i} {}\\ & =& {(-1)}^{b}\sum _{ k=0}^{b}{(-1)}^{k}\binom{d}{b - k}\xi _{ -2i-2-2k}^{2i} - {(-1)}^{b}\sum _{ k=0}^{a}{(-1)}^{k}\binom{d}{a - k}\xi _{ 2i+2k}^{2i} {}\\ & =& {(-1)}^{b}\sum _{ k=0}^{b}{(-1)}^{k}\binom{d}{b - k}\binom{2i + k}{{2i}\alpha }^{2i} - {(-1)}^{b}\sum _{ k=0}^{a}{(-1)}^{k}\binom{d}{a - k}\binom{2i + k}{{2i}\alpha }^{2i}. {}\\ \end{array}$$

Taking the coefficient of (−x)b in (1 − x)d∕(1 − x)2i+1 = (1 − x)a+b we have

$$\displaystyle{ \sum _{k=0}^{b}{(-1)}^{k}\binom{d}{b - k}\binom{2i + k}{2i} = \binom{a + b}{b}. }$$

Exchanging the roles of a and b we see that Z = 0 which implies (B.11).

Proposition B.2.

For all d ≥ 1, \(m \in \mathbb{Z}\) , and w ∈ W af we have

$$\displaystyle{{ \xi }^{m}({(1 - t_{{\alpha }^{\vee }})}^{d-1}(1 - s_{\alpha })w) {\in \alpha }^{d}\mathbb{Z}[\alpha ]. }$$

Proof.

Let \(y = {(1 - t_{{\alpha }^{\vee }})}^{d-1}\). Without loss of generality we may assume \(w = t_{{p\alpha }^{\vee }}\) for some \(p \in \mathbb{Z}\). By Proposition B.1, ξ m satisfies (4.1). We have

$$ \displaystyle\begin{array}{rcl} {\xi }^{m}(y(1 - s_{\alpha })t_{{ p\alpha }^{\vee }})& =& {\xi }^{m}(yt_{{ p\alpha }^{\vee }}) {-\xi }^{m}(yt_{ -{p\alpha }^{\vee }}s_{\alpha }) {}\\ & \equiv & {\xi }^{m}(y) {-\xi }^{m}{(ys_{\alpha })\mod \alpha }^{d}\mathbb{Z}[\alpha ], {}\\ \end{array}$$

using Lemma 4.1 twice. However

$$ \displaystyle\begin{array}{rcl} {\xi }^{m}(y) {-\xi }^{m}(ys_{\alpha })& =& {\xi }^{m}(y\alpha A_{\alpha }) {}\\ & =& (y\cdot \alpha )(A{_{\alpha }\bullet \xi }^{m})(y). {}\\ \end{array}$$

Now y ⋅ α = ±α and A α ξ m is 0 if m ≥ 0 and is equal to ξ m+1 if m < 0. Assuming the latter, by (4.1) for d − 1, ξ m+1(y) ∈ α d−1 S, so that \({\xi }^{m}(y) {-\xi }^{m}(ys_{\alpha }) {\in \alpha }^{d}S\) as required.

C Appendix: Homology of Gr

Let K be the maximal compact form of G and \(T_{\mathbb{R}} = K \cap T\). The based loop group \(\Omega K\) of continuous maps (S 1, 1) → (K, 1), has a \(T_{\mathbb{R}}\)-equivariant multiplication map \(\Omega K \times \Omega K \rightarrow \Omega K\) given by pointwise multiplication on K, and this induces the structure of a commutative and co-commutative Hopf-algebra on \({H}^{T_{\mathbb{R}}}(\Omega K)\). The co-commutativity of \({H}^{T_{\mathbb{R}}}(\Omega K)\) follows from the fact that \(\Omega K\) is a homotopy double-loop space.

Gr = Gr G and \(\Omega K\) are weakly homotopy equivalent [135]. By “fattening loops”, it follows that H T (Gr) and \({H}^{T}(\tilde{\mathrm{Gr}})\) are dual Hopf algebras over S = H T(pt) with duality given by an intersection pairing

$$\displaystyle{ \langle \cdot \,,\,\cdot \rangle: H_{T}(\mathrm{Gr}) \times {H}^{T}(\tilde{\mathrm{Gr}}) \rightarrow S. }$$
(C.1)

We may regard an element ξ ∈ H T (Gr) as an S-module homomorphism \({H}^{T}(\tilde{\mathrm{Gr}}) \rightarrow S\) by \(\xi (f) =\langle \xi \,,\,f\rangle\).

Lemma C.1.

Let λ,μ ∈ Q , and consider the maps \(i_{\lambda }^{{\ast}},i_{\mu }^{{\ast}}: {H}^{T}(\tilde{\mathrm{Gr}}) \rightarrow S\) as elements of H T (Gr ). Then in H T (Gr ), we have

$$\displaystyle{i_{\lambda }^{{\ast}}\;i_{\mu }^{{\ast}} = i_{\lambda +\mu }^{{\ast}}.}$$

Proof.

It suffices to work in \({H}^{T_{\mathbb{R}}}(\Omega K)\). The map \(i_{\lambda }^{{\ast}}\;i_{\mu }^{{\ast}}\) is induced by the map \(\mathrm{pt} \rightarrow \Omega K \times \Omega K \rightarrow \Omega K\) where the image of the first map is the pair \((t_{\lambda },t_{\mu }) \in \Omega K \times \Omega K\) of fixed points, and the latter map is multiplication. Treating \(t_{\lambda },t_{\mu }: {S}^{1} \rightarrow K\) as homomorphisms into K, the pointwise multiplication of t λ and t μ gives t λ+μ . Thus \(i_{\lambda }^{{\ast}}\;i_{\mu }^{{\ast}} = i_{\lambda +\mu }^{{\ast}}\).

The antipode of H T (Gr G ) is given by \(i_{\lambda }^{{\ast}}\mapsto i_{-\lambda }^{{\ast}}\), since the fixed points satisfy \(t_{\lambda }^{-1} = t_{-\lambda }\) in \(\Omega K\).

For \(w \in W_{\mathrm{af}}^{0}\) denote by \([X_{w}]_{T} \in H_{T}(\mathrm{Gr})\) the equivariant fundamental homology class of the Schubert variety \(X_{w}:= \overline{B_{\mathrm{af}}\dot{w}P_{\mathrm{af}}/P_{\mathrm{af}}} \subset G_{\mathrm{af}}/P_{\mathrm{af}} =\mathrm{ Gr}\).

Using the intersection pairing we have that \(\{[X_{w}]\mid w \in W_{\mathrm{af}}^{0}\}\) is the basis of H T (Gr G ) that is dual to the basis \(\{{[{X}^{w}]}^{T}\mid w \in W_{\mathrm{af}}^{0}\}\) of \({H}^{T}(\tilde{\mathrm{Gr}})\), which corresponds to the basis \(\{\bar{{\xi }}^{w}\mid w \in W_{\mathrm{af}}^{0}\}\) of \(\Lambda _{\mathrm{Fl}}^{{\prime}}\) under the isomorphism of Theorem 4.3.

Proof of Proposition 4.6.

Follows from the above discussion.

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer Science+Business Media New York

About this chapter

Cite this chapter

Lam, T., Lapointe, L., Morse, J., Schilling, A., Shimozono, M., Zabrocki, M. (2014). Affine Schubert Calculus. In: k-Schur Functions and Affine Schubert Calculus. Fields Institute Monographs, vol 33. Springer, New York, NY. https://doi.org/10.1007/978-1-4939-0682-6_4

Download citation

Publish with us

Policies and ethics