This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Double Grothendieck Polynomials for Symplectic and Odd Orthogonal Grassmannians

Thomas Hudson, Takeshi Ikeda, Tomoo Matsumura, Hiroshi Naruse
Abstract

We study the double Grothendieck polynomials of Kirillov–Naruse for the symplectic and odd orthogonal Grassmannians. These functions are explicitly written as sums of Pfaffian and are identified with the stable limits of the fundamental classes of Schubert varieties in the torus equivariant connective KK-theory of these isotropic Grassmannians. We also provide a combinatorial description of the ring formally spanned by double Grothendieck polynomials.

1 Introduction

In [9], Kirillov and Naruse introduced the double Grothendieck polynomials of classical types in order to represent the KK-theoretic Schubert classes for the corresponding flag varieties. In this paper, we study these functions for the odd orthogonal and symplectic Grassmannians. We first set up a combinatorial model for the ring formally spanned by the stable Schubert classes. This construction is based on previous work for the maximal isotropic Grassmannians ([8]). By utilizing our result in [5] and equivariant localization techniques ([10], [11]), we will explicitly specify the elements in the ring corresponding to the stable Schubert classes.

Fix integers kk and nn such that 0k<n0\leq k<n. Let 𝔽{\mathbb{F}} be an algebraically closed field of characteristic zero. We equip the vector space 𝔽2n+1{\mathbb{F}}^{2n+1} with a non-degenerate symmetric bilinear form and 𝔽2n{\mathbb{F}}^{2n} with a symplectic form. We denote respectively by OGk(n){OG}^{k}(n) and SGk(n)SG^{k}(n) the odd-orthogonal and symplectic Grassmannians of (nk)(n-k)-dimensional isotropic subspaces in 𝔽2n+1{\mathbb{F}}^{2n+1} and 𝔽2n{\mathbb{F}}^{2n}. We denote both these isotropic Grassmannians by IGk(n){{IG}}^{k}(n). Each of such spaces can be realized as a homogeneous space G/PkG/P_{k} where PkP_{k} is a parabolic subgroup either of the special orthogonal group G=SO2n+1(𝔽)G={{SO}}_{2n+1}({\mathbb{F}}) or of the symplectic group G=Sp2n(𝔽)G={Sp}_{2n}({\mathbb{F}}). Let TnT_{n} be a maximal torus of GG, BB a Borel subgroup of GG containing TnT_{n}, and BB_{-} a Borel subgroup opposite to BB, i.e., BB=Tn.B\cap B_{-}=T_{n}. For both SO2n+1(𝔽){{SO}}_{2n+1}({\mathbb{F}}) and Sp2n(𝔽){Sp}_{2n}({\mathbb{F}}), the Weyl group WnW_{n} is isomorphic to the hyperocthedral group given by {1}nSn\{-1\}^{n}\rtimes S_{n}. Let Wn(k)W_{n}^{(k)} be the set of minimal length coset representatives of Wn/Wn,kW_{n}/W_{n,k} where Wn,kW_{n,k} is the Weyl group of PkP_{k}. There is a natural bijection between Wn(k)W_{n}^{(k)} and the set 𝒮𝒫k(n){\mathcal{S}}{\mathcal{P}}^{k}(n) of kk-strict partitions contained in the (nk,n+k)(n-k,n+k) rectangle (cf. Buch–Kresch–Tamvakis [2]). Moveover, in both the odd-orthogonal and symplectic cases, Wn(k)W_{n}^{(k)} is also in bijection with the set of TnT_{n}-fixed points in IGk(n){{IG}}^{k}(n). For each λ𝒮𝒫k(n)\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n), let eλe_{\lambda} be the corresponding TnT_{n}-fixed point and define the Schubert variety Ωλ\Omega_{\lambda} associated to λ\lambda as the closure of its BB_{-}-orbit.

We will use the torus-equivariant connective KK-theory studied by Krishna [11] (cf. Dai–Levine [3]). For any non-singular variety XX endowed with an action of an algebraic torus Tn(𝔽×)nT_{n}\cong({\mathbb{F}}^{\times})^{n}, there is a {\mathbb{Z}}-graded algebra CKTn(X){C\!K}_{T_{n}}(X) over CKTn(pt){C\!K}_{T_{n}}(pt) which can be identified with the graded formal power series ring [β][[b1,,bn]]gr{\mathbb{Q}}[\beta][[b_{1},\ldots,b_{n}]]_{\operatorname{gr}} where deg(bi)=1\deg(b_{i})=1 and deg(β)=1\deg(\beta)=-1. It interpolates the TnT_{n}-equivariant chow ring CHTn(X){C\!H}_{T_{n}}^{*}(X) and the KK-theory KTn(X)K_{T_{n}}(X) of TnT_{n}-equivariant algebraic vector bundles on XX. By this we mean that it specializes to CHTn(X){C\!H}_{T_{n}}^{*}(X) for β=0\beta=0 and to KTn(X)K_{T_{n}}(X) for β=1\beta=-1. Since throughout the paper we will work with rational coefficients, here CHTn(X){C\!H}_{T_{n}}^{*}(X) stands for the equivariant chow ring with {\mathbb{Q}} coefficients and KTn(X)K_{T_{n}}(X) denotes the equivariant KK-theory tensored with {\mathbb{Q}}.

In CKTn(IGk(n)){C\!K}_{T_{n}}({{IG}}^{k}(n)), there exists a fundamental class [Ωλ]Tn[\Omega_{\lambda}]_{T_{n}} associated to the Schubert variety Ωλ\Omega_{\lambda}, which specializes to the corresponding TnT_{n}-equivariant Schubert class in CHTn(IGk(n)){C\!H}^{*}_{T_{n}}({{IG}}^{k}(n)) and to the class of the structure sheaf of Ωλ\Omega_{\lambda} in KTn(IGk(n))K_{T_{n}}({{IG}}^{k}(n)). As we are interested in [Ωλ]T[\Omega_{\lambda}]_{T}, the limit of the classes [Ωλ]Tn[\Omega_{\lambda}]_{T_{n}} in the graded projective limit of the rings CKTn(IGk(n)){C\!K}_{T_{n}}({{IG}}^{k}(n)), we introduce 𝕂Tfin(IGk){\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}) a subring which contains all of them (see Definition 4.5). More precisely, if we set

b=n0[β][[b1,,bn]]gr,{\mathcal{R}}_{b}=\bigcup_{n\geq 0}{\mathbb{Q}}[\beta][[b_{1},\ldots,b_{n}]]_{\operatorname{gr}},

then 𝕂Tfin(IGk){\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}) is the b{\mathcal{R}}_{b}-module on the formal basis given by [Ωλ]T[\Omega_{\lambda}]_{T} indexed by the all kk-strict partitions λ\lambda.

In order to deal with 𝕂Tfin(IGk){\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}) algebraically and combinatorially, we introduce the ring 𝒦{\mathcal{K}}_{\infty}, which governs the equivariant KK-theoretic Schubert calculus of the full-flag varieties of both types B and C. A cornerstone of the construction of 𝒦{\mathcal{K}}_{\infty} is the ring GΓ{{G\Gamma}} introduced in [8]. It is defined as a subring of [β][[x]]gr{\mathbb{Q}}[\beta][[x]]_{\operatorname{gr}} with variables x=(x1,x2,)x=(x_{1},x_{2},\dots) consisting of symmetric graded formal power series with a certain cancellation property (see §3.1). Then one sets

𝒦=GΓ[β]a[β]b.{\mathcal{K}}_{\infty}={G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{b}.

This construction is analogous to the one of the ring {\mathcal{R}}_{\infty} introduced as an algebraic model for the torus equivariant cohomology of full-flag varieties of type B, C, and D in [7]. There are two (called left and right) commuting actions of W=n1WnW_{\infty}=\bigcup_{n\geq 1}W_{n} on 𝒦{\mathcal{K}}_{\infty}. The left action commutes with a{\mathcal{R}}_{a} while the right one commutes with b{\mathcal{R}}_{b}. The ring 𝒦(k){\mathcal{K}}^{(k)}_{\infty} is then the invariant subring with respect to the right action of the group W(k)=n>kWn,kW_{(k)}=\bigcup_{n>k}W_{n,k}. In particular, the invariant ring 𝒦(0){\mathcal{K}}^{(0)}_{\infty} coincides with GΓ[β]b{G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{b}.

The first main result is the following algebraic realization of 𝕂Tfin(IGk){\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}), which we obtain by comparing the GKM descriptions of 𝒦(k){\mathcal{K}}^{(k)}_{\infty} and CKTn(IGk(n)){C\!K}_{T_{n}}({{IG}}^{k}(n)).

Theorem (Theorem 5.10 below) There are isomorphisms of graded b{\mathcal{R}}_{b}-algebras

𝒦(k)𝕂Tfin(OGk) and 𝒦(k)𝕂Tfin(SGk).{\mathcal{K}}^{(k)}_{\infty}\cong{\mathbb{K}}^{\operatorname{fin}}_{T}({OG}^{k})\ \ \mbox{ and }\ \ \ {\mathcal{K}}^{(k)}_{\infty}\cong{\mathbb{K}}^{\operatorname{fin}}_{T}(SG^{k}).

The second main result is the description of the those elements in 𝒦(k){\mathcal{K}}_{\infty}^{(k)} corresponding to [Ωλ]T[\Omega_{\lambda}]_{T} under the above isomorphisms. For each kk-strict partition λ\lambda, we define the functions GkΘλ{}_{k}{G\Theta^{\prime}}_{\!\lambda} and GkΘλ{}_{k}{G\Theta}_{\lambda} in 𝒦(k){\mathcal{K}}_{\infty}^{(k)} via a Pfaffian formula (Definition 3.13) analogous to the one obtained in [5] for the Schubert classes for CKTn(IGk(n)){C\!K}_{T_{n}}({{IG}}^{k}(n)) (cf.​ Theorem 4.7). This allows us to show that those functions correspond to the limits of the Schubert classes (Theorem 5.10). Note that the Pfaffian formula in [5] describes the Schubert classes in terms of the (equivariantly shifted) special classes which we also identify with certain functions in 𝒦(k){\mathcal{K}}_{\infty}^{(k)} (Lemma 5.9). It is worth mentioning that when β=0\beta=0, the functions GkΘλ{}_{k}{G\Theta}_{\lambda} specialize to Wilson’s double ϑ\vartheta-functions [16] which coincide with the type C double Schubert polynomials (see [6]).

For the maximal isotropic Grassmannians in type B and C, the isomorphisms in the above theorem were obtained in [8]. There have also been introduced the so-called GP{GP}- and GQ{GQ}-functions in a combinatorially explicit formula, which represent the KK-theory Schubert classes. Thus by our result we find that, in the case k=0k=0, they coincide with G0Θλ{}_{0}{G\Theta^{\prime}}_{\!\lambda} and G0Θλ{}_{0}{G\Theta}_{\lambda} respectively, hence establishing their Pfaffian formula simultaneaously.

Kirillov–Naruse [9] constructed the double Grothendieck polynomials representing the equivariant KK-theoretic Schubert classes for the full-flag varieties of type B and C. One can prove that our new functions consist of an appropriate subfamily of their polynomials, simultaneously establishing their explicit closed formula in the form of Pfaffians. We also remark that by their construction, those functions are polynomials in xx, aa, bb variables and β\beta with coefficients in 0{\mathbb{Z}}_{\geq 0}, once we choose a positive integer nn and set xi=0x_{i}=0 for all i>ni>n.

Given the previous remark related to Kirillov–Naruse’s constructions, let us mention some combinatorial problems that still remain with regard to our GΘ{G\Theta}- and GΘ{G\Theta^{\prime}}-functions. When k=0k=0, it was proved in [8] that GQ{GQ}- and GP{GP}-functions are given in terms of shifted set-valued tableaux, generalizing the classical tableaux formula of Schur QQ- and PP-functions (cf. [13, §III.8]). When β=0\beta=0 and bi=0b_{i}=0 for all ii but for general kk, Tamvakis [14] obtained an analogous combinatorial formula for the theta polynomials. Thus it would be an interesting problem to find similar combinatorial formulas for our GΘ{G\Theta}- and GΘ{G\Theta^{\prime}}-functions. It is also worth mentioning that, again from Kirillov–Naruse’s construction, we can deduce the fact that GΘ{G\Theta}-functions (resp. GΘ{G\Theta^{\prime}}-functions) can be expanded into a linear combination of GQ{GQ}-functions (resp. GP{GP}-functions) with coefficients in [β][a,b]{\mathbb{Z}}[\beta][a,b]. It is an open problem to find an explicit formula for the coefficients of the expansions. Note that when β=0\beta=0, such formula has been known in the work of Buch–Kresch–Tamvakis [2, Theorem 4] and Tamvakis–Wilson [15, Corollary 2].

Finally, we note that recently Anderson [1] proved a formula describing a larger family of KK-theoretic Schubert classes including the case of type D isotropic Grassmannians, generalizing our results for type B and C in [5]. Given this, we expect that it is possible to define the functions analogous to GΘ{G\Theta} and GΘ{G\Theta^{\prime}} explicitly, extending the main results of this paper to type D.

This paper is organized as follows. In Section 2, we recall the combinatorics related to the Weyl group of type B and C. In Section 3, we define the rings GΓ{G\Gamma}, 𝒦{\mathcal{K}}_{\infty} and 𝒦(k){\mathcal{K}}_{\infty}^{(k)} and introduce the functions GkΘλ{}_{k}{G\Theta^{\prime}}_{\!\lambda} and GkΘλ{}_{k}{G\Theta}_{\lambda}. In Section 4, we recall the definitions of the odd orthogonal and symplectic Grassmannians together with the torus equivariant KK-theory Schubert classes. In Section 5, we compare the GKM descriptions of 𝒦(k){\mathcal{K}}_{\infty}^{(k)} and CKTn(OGk(n)){C\!K}_{T_{n}}({OG}^{k}(n)) and define a map relating those two algebras. Finally we establish our main results, namely the isomorphisms in the above theorem, and the fact that GkΘλ{}_{k}{G\Theta^{\prime}}_{\!\lambda} and GkΘλ{}_{k}{G\Theta}_{\lambda} represent the Schubert classes.

2 Preliminary on combinatorics of type B and C

We fix notation on Weyl groups, root systems, and kk-strict partitions.

2.1 Weyl group and root systems

Let WW_{\infty} be the infinite hyperoctahedral group, which is defined by the generators si,i=0,1,s_{i},i=0,1,\ldots, and the relations

si2=e(i0),sisj=sjsi(|ij|2),s0s1s0s1=s1s0s1s0,sisi+1si=si+1sisi+1(i1).\begin{array}[]{c}s_{i}^{2}=e\ \ \ \ (i\geq 0),\ \ \ \ \ \ \ s_{i}s_{j}=s_{j}s_{i}\ \ \ \ (|i-j|\geq 2),\\ s_{0}s_{1}s_{0}s_{1}=s_{1}s_{0}s_{1}s_{0},\ \ s_{i}s_{i+1}s_{i}=s_{i+1}s_{i}s_{i+1}\ \ \ \ (i\geq 1).\\ \end{array} (2.1)

We identify WW_{\infty} with the group of signed permutations, i.e. all permutations ww of \{0}{\mathbb{Z}}\backslash\{0\} such that w(i)iw(i)\not=i for only finitely many i\{0}i\in{\mathbb{Z}}\backslash\{0\}, and w(i)¯=w(i¯)\overline{w(i)}=w(\bar{i}) for all ii. In this context i¯\bar{i} stands for i-i. The generators, often referred to as simple reflections, are identified with the transpositions s0=(1,1¯)s_{0}=(1,\bar{1}) and si=(i+1,i)(i¯,i+1¯)s_{i}=(i+1,i)(\overline{i},\overline{i+1}) for i1i\geq 1. The one-line notation of an element wWw\in W_{\infty} is the sequence w=(w(1),w(2),w(3),)w=(w(1),w(2),w(3),\cdots). The length of wWw\in W_{\infty} is denoted by (w)\ell(w).

For each nonnegative integer kk, let W(k)W_{(k)} be the subgroup of WW_{\infty} generated by all sis_{i} with iki\not=k. Let W(k)W_{\infty}^{(k)} be the set of minimum length coset representatives for W/W(k)W_{\infty}/W_{(k)}, which is described as

W(k)={wW|(wsi)>(w) for all ik}.W_{\infty}^{(k)}=\{w\in W_{\infty}\ |\ \ell(ws_{i})>\ell(w)\mbox{ for all }i\not=k\}.

An element of W(k)W_{\infty}^{(k)} is called kk-Grassmannian and it is given by the following one-line notation:

w=(v1,,vk,ζ1¯,,ζs¯,u1,u2,);0<v1<<vk,ζ1¯<<ζs¯<0<u1<u2<,\begin{array}[]{c}w=(v_{1},\cdots,v_{k},\overline{\zeta_{1}},\cdots,\overline{\zeta_{s}},u_{1},u_{2},\cdots);\\ 0<v_{1}<\cdots<v_{k},\ \ \ \overline{\zeta_{1}}<\cdots<\overline{\zeta_{s}}<0<u_{1}<u_{2}<\cdots,\end{array} (2.2)

where s0s\geq 0. For example, (1,3,4¯,2,5,6,7,)(1,3,\bar{4},2,5,6,7,\cdots) is a 22-Grassmannian element in WW_{\infty}.

Upon a choice of an integer n1n\geq 1, we let WnW_{n} be the subgroup of WW_{\infty} generated by s0,s1,,sn1s_{0},s_{1},\dots,s_{n-1}. Or, equivalently, it consists of the elements wWw\in W_{\infty} such that w(i)=iw(i)=i for all i>ni>n. We write the one-line notation of wWnw\in W_{n} as the finite sequence (w(1),w(2),,w(n))(w(1),w(2),\cdots,w(n)). We set Wn,(k):=WnW(k)W_{n,(k)}:=W_{n}\cap W_{(k)} and Wn(k):=WnW(k)W_{n}^{(k)}:=W_{n}\cap W_{\infty}^{(k)} so that Wn(k)Wn/Wn,(k)W_{n}^{(k)}\cong W_{n}/W_{n,(k)}.

Let L:=i=1εiL:=\bigoplus_{i=1}^{\infty}{\mathbb{Z}}\varepsilon_{i} be the free module generated by εi,i>0\varepsilon_{i},i\in{\mathbb{Z}}_{>0}. We define the set Δ+\Delta^{+} of positive roots in LL as

Δ+:={εi,1i}{εj±εi| 1i<j}for type B and\displaystyle\Delta^{+}:=\{\varepsilon_{i},1\leq i\}\cup\{\varepsilon_{j}\pm\varepsilon_{i}\ |\ 1\leq i<j\}\ \ \ \mbox{for type B and}
Δ+:={2εi,1i}{εj±εi| 1i<j}for type C.\displaystyle\Delta^{+}:=\{2\varepsilon_{i},1\leq i\}\cup\{\varepsilon_{j}\pm\varepsilon_{i}\ |\ 1\leq i<j\}\ \ \ \mbox{for type C.}

Let sαWs_{\alpha}\in W_{\infty} denote the simple reflection associated with the positive root αΔ+\alpha\in\Delta^{+}. Let Δn+:=Δ+Span{ε1,,εn}\Delta^{+}_{n}:=\Delta^{+}\cap\ \operatorname{Span}_{{\mathbb{Z}}}\{\varepsilon_{1},\dots,\varepsilon_{n}\}, then for αΔn+\alpha\in\Delta^{+}_{n}, the simple reflection sαs_{\alpha} is an element of WnW_{n}.

2.2 kk-strict partitions

Fix a nonnegative integer kk. A kk-strict partition is an infinite sequence (λ1,λ2,)(\lambda_{1},\lambda_{2},\cdots) of non-increasing nonnegative integers such that all but finitely many λi\lambda_{i}’s are zero, and such that λi>k\lambda_{i}>k implies λi>λi+1\lambda_{i}>\lambda_{i+1}. The length of a kk-strict partition is the largest index rr such that λr0\lambda_{r}\not=0. We often denote a kk-strict partition λ\lambda with length at most rr by λ=(λ1,,λr)\lambda=(\lambda_{1},\dots,\lambda_{r}). We use the following notation for the sets of kk-strict partitions:

𝒮𝒫k\displaystyle{\mathcal{S}}{\mathcal{P}}^{k} :\displaystyle: the set of all kk-strict partitions;
𝒮𝒫rk\displaystyle{\mathcal{S}}{\mathcal{P}}_{r}^{k} :\displaystyle: the set of all kk-strict partitions with length at most rr;
𝒮𝒫k(n)\displaystyle{\mathcal{S}}{\mathcal{P}}^{k}(n) :\displaystyle: the set of all k-strict partitions with length at most nk and λ1n+k.\displaystyle\mbox{the set of all $k$-strict partitions with length at most $n-k$ and $\lambda_{1}\leq n+k$}.

There is a bijection from W(k)W_{\infty}^{(k)} to 𝒮𝒫k{\mathcal{S}}{\mathcal{P}}^{k} given as follows. Let wW(k)w\in W_{\infty}^{(k)} be an element with the one-line notation (2.2). Let ν=(ν1,ν2,)\nu=(\nu_{1},\nu_{2},\dots) be a partition given by νi={p|vp>ui}\nu_{i}=\sharp\{p\ |\ v_{p}>u_{i}\}. We define a kk-strict partition λ\lambda by setting λi=ζi+k\lambda_{i}=\zeta_{i}+k if 1is1\leq i\leq s and λi=νis\lambda_{i}=\nu_{i-s} if s+1is+1\leq i. This bijection can be restricted to Wn(k)W_{n}^{(k)}, i.e. Wn(k)𝒮𝒫k(n)W_{n}^{(k)}\cong{\mathcal{S}}{\mathcal{P}}^{k}(n). See Buch–Kresch–Tamvakis [2] for details. Through this bijection W(k)𝒮𝒫kW_{\infty}^{(k)}\cong{\mathcal{S}}{\mathcal{P}}^{k}, the Weyl group WW_{\infty} acts naturally on 𝒮𝒫k{\mathcal{S}}{\mathcal{P}}^{k}. Similarly, WnW_{n} acts on 𝒮𝒫k(n){\mathcal{S}}{\mathcal{P}}^{k}(n) via the bijection Wn(k)𝒮𝒫k(n)W_{n}^{(k)}\cong{\mathcal{S}}{\mathcal{P}}^{k}(n).

3 Rings of double Grothendieck polynomials

We first define the ring GΓ{G\Gamma} and introduce its formal basis {GPλ(x)}λ𝒮𝒫0\{{GP}_{\lambda}(x)\}_{\lambda\in{\mathcal{S}}{\mathcal{P}}^{0}}. The ring 𝒦{\mathcal{K}}_{\infty} of double Grothendieck polynomials is then introduced. On this ring we have a natural (right) action of WW_{\infty}. For each nonnegative integer kk, we define the functions GkΘλ{}_{k}{G\Theta}_{\lambda} and GkΘλ{}_{k}{G\Theta^{\prime}}_{\!\!\lambda} associated to all λ𝒮𝒫k\lambda\in{\mathcal{S}}{\mathcal{P}}^{k} as elements in the invariant subring 𝒦(k){\mathcal{K}}^{(k)}_{\infty} with respect to the action of the subgroup W(k)W_{\infty}^{(k)} of WW_{\infty}.

3.1 The ring GΓ{G\Gamma} and its formal basis

Let x=(x1,x2,)x=(x_{1},x_{2},\ldots) be a sequence of indeterminates of degree 11. Let [β]{\mathbb{Q}}[\beta] be the polynomial ring in a variable β\beta with degβ=1\deg\beta=-1. Denote by [β][[x]]m{\mathbb{Q}}[\beta][[x]]_{m} the set of all power series in xx with coefficients in [β]{\mathbb{Q}}[\beta] of degree mm\in{\mathbb{Z}}. Then [β][[x]]gr:=m[β][[x]]m{\mathbb{Q}}[\beta][[x]]_{\operatorname{gr}}:=\bigoplus_{m\in{\mathbb{Z}}}{\mathbb{Q}}[\beta][[x]]_{m} is a graded [β]{\mathbb{Q}}[\beta]-algebra, called the ring of graded formal power series. We denote the ring of graded formal power series in a finite sequence (x1,,xn)(x_{1},\dots,x_{n}) of variables by [β][[x1,,xn]]gr{\mathbb{Q}}[\beta][[x_{1},\dots,x_{n}]]_{\operatorname{gr}}. Throughout the paper, we denote, for any variables uu and vv,

uv:=u+v+βuv,uv:=uv1+βv,u¯:=u1+βu.u\oplus v:=u+v+\beta uv,\ \ \ u\ominus v:=\frac{u-v}{1+\beta v},\ \ \ \bar{u}:=\frac{-u}{1+\beta u}.

where 1+βu=m0βmum1+\beta u=\sum_{m\geq 0}\beta^{m}u^{m}.

Definition 3.1 ([8]).

We denote by GΓn{G\Gamma}_{n} the graded subring of [β][[x1,,xn]]gr{\mathbb{Q}}[\beta][[x_{1},\ldots,x_{n}]]_{\operatorname{gr}} whose elements are the series f(x)f(x) such that:

  • (1)

    f(x)f(x) is symmetric in x1,,xnx_{1},\ldots,x_{n};

  • (2)

    f(t,t¯,x3,x4,,xn)=f(0,0,x3,x4,,xn)f(t,\bar{t},x_{3},x_{4},\dots,x_{n})=f(0,0,x_{3},x_{4},\dots,x_{n}).

These rings form a projective system with respect to the degree preserving homomorphism GΓn+1GΓn{G\Gamma}_{n+1}\to{G\Gamma}_{n} given by xn+1=0x_{n+1}=0. Let us denote by GΓ{G\Gamma} its graded projective limit. We can identify GΓ{G\Gamma} with the subring of [β][[x]]gr{\mathbb{Q}}[\beta][[x]]_{\operatorname{gr}} defined by the conditions (1)’ f(x)f(x) is symmetric in xi,i>0x_{i},i\in{\mathbb{Z}}_{>0} and (2)’ f(t,t¯,x3,x4,)=f(0,0,x3,x4,)f(t,\bar{t},x_{3},x_{4},\dots)=f(0,0,x_{3},x_{4},\dots), analogous to (1) and (2) above. Note that our GΓ{G\Gamma} is a completion of the one defined in [8].

Definition 3.2 ([8]).

For each strict partition λ=(λ1,,λr)\lambda=(\lambda_{1},\dots,\lambda_{r}) of length rr in 𝒮𝒫n0{\mathcal{S}}{\mathcal{P}}_{n}^{0}, the corresponding GPGP-functions are defined by

GPλ(x1,,xn)=1(nr)!wSnw[x1λ1xrλri=1rj=i+1nxixjxixj].{GP}_{\lambda}(x_{1},\dots,x_{n})=\frac{1}{(n-r)!}\sum_{w\in S_{n}}w\left[x_{1}^{\lambda_{1}}\cdots x_{r}^{\lambda_{r}}\prod_{i=1}^{r}\prod_{j=i+1}^{n}\frac{x_{i}\oplus x_{j}}{x_{i}\ominus x_{j}}\right].

The polynomial GPλ(x1,,xn){GP}_{\lambda}(x_{1},\dots,x_{n}) is an element of GΓn{G\Gamma}_{n} for all nn, and these define an element GPλ(x){GP}_{\lambda}(x) in GΓ{G\Gamma} as the limit.

The next proposition will be used in the proof of Lemma 5.5 at §5\S\ref{sec5}.

Proposition 3.3.

Any homogeneous element f(x)f(x) of GΓ{G\Gamma} with degree mm is uniquely expressed as a possibly infinite linear combination

f(x)=λ𝒮𝒫0cλGPλ(x),cλ[β]m|λ|.f(x)=\sum_{\lambda\in{\mathcal{S}}{\mathcal{P}}^{0}}c_{\lambda}{GP}_{\lambda}(x),\ \ \ c_{\lambda}\in{\mathbb{Q}}[\beta]_{m-|\lambda|}.
Proof.

The claim follows formally from the fact that GPλ(x1,,xn)(λ𝒮𝒫n0){GP}_{\lambda}(x_{1},\dots,x_{n})\ (\lambda\in{\mathcal{S}}{\mathcal{P}}_{n}^{0}) form a formal basis of GΓn{G\Gamma}_{n}. That is, a homogeneous element f(x1,,xn)f(x_{1},\dots,x_{n}) in GΓn{G\Gamma}_{n} of degree rr is uniquely expressed as a possibly infinite linear combination

f(x1,,xn)=λ𝒮𝒫n0cλGPλ(x1,,xn),cλ[β]r|λ|.f(x_{1},\dots,x_{n})=\sum_{\lambda\in{\mathcal{S}}{\mathcal{P}}_{n}^{0}}c_{\lambda}{GP}_{\lambda}(x_{1},\dots,x_{n}),\ \ \ c_{\lambda}\in{\mathbb{Q}}[\beta]_{r-|\lambda|}.

This fact is a slight modification of Theorem 3.1 [8]: here we must work in the ring of graded formal power series. We leave the details to the reader since it is parallel to the original one. ∎

3.2 Rings of double Grothendieck polynomials

For infinite sequences of variables a=(a1,a2,)a=(a_{1},a_{2},\dots) and b=(b1,b2,)b=(b_{1},b_{2},\dots), consider the rings

a\displaystyle{\mathcal{R}}_{a} :=\displaystyle:= m=0[β][[a1,,am]]gr,b:=m=0[β][[b1,,bm]]gr.\displaystyle\bigcup_{m=0}^{\infty}{\mathbb{Q}}[\beta][[a_{1},\dots,a_{m}]]_{\operatorname{gr}},\ \ \ {\mathcal{R}}_{b}:=\bigcup_{m=0}^{\infty}{\mathbb{Q}}[\beta][[b_{1},\dots,b_{m}]]_{\operatorname{gr}}.

and define the b{\mathcal{R}}_{b}-algebra 𝒦{\mathcal{K}}_{\infty} by

𝒦:=GΓ[β]a[β]b.{\mathcal{K}}_{\infty}:={G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{b}.

We define an action of WW_{\infty} on 𝒦{\mathcal{K}}_{\infty} as follows. For f(x;a)GΓ[β]af(x;a)\in{G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a} we set

(s0f)(x;a)\displaystyle(s_{0}f)(x;a) =\displaystyle= f(a1,x1,x2,;a1¯,a2,)\displaystyle f(a_{1},x_{1},x_{2},\ldots;\overline{a_{1}},a_{2},\ldots)
(sif)(x;a)\displaystyle(s_{i}f)(x;a) =\displaystyle= f(x1,x2,;a1,a2,,ai+1,ai,)(i1)\displaystyle f(x_{1},x_{2},\ldots;a_{1},a_{2},\ldots,a_{i+1},a_{i},\ldots)\quad(i\geq 1)

and extend these as automorphisms of b{\mathcal{R}}_{b}-algebras. One can check that this gives an action of WW_{\infty} on 𝒦{\mathcal{K}}_{\infty}. For example, s02=1s_{0}^{2}=1 follows from the definition of GΓ{G\Gamma}. We call this the right action, while the left action is similarly defined by replacing the roles of aa and bb. In this paper, we only use the right one.

Now for each k0k\geq 0, we define 𝒦(k){\mathcal{K}}_{\infty}^{(k)} to be the subalgebra of 𝒦{\mathcal{K}}_{\infty} invariant under the W(k)W_{(k)}-action:

𝒦(k):=𝒦W(k)=(GΓ[β]a)W(k)[β]b.{\mathcal{K}}_{\infty}^{(k)}:={\mathcal{K}}_{\infty}^{W_{(k)}}=({G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a})^{W_{(k)}}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{b}.

Note that the subring (GΓ[β]a)W(k)({G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a})^{W_{(k)}} of GΓ[β]a{G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a} invariant under the action of W(k)W_{(k)}, is contained in GΓ[[a1,,ak]]gr{G\Gamma}[[a_{1},\dots,a_{k}]]_{\operatorname{gr}} since each element of a{\mathcal{R}}_{a} involves only finitely many aia_{i}’s.

Remark 3.4.

The ring 𝒦{\mathcal{K}}_{\infty} (resp. 𝒦(k){\mathcal{K}}_{\infty}^{(k)}) is the KK-theoretic version of {\mathcal{R}}_{\infty} introduced in [7] (resp. (k){\mathcal{R}}_{\infty}^{(k)} in [6]). The corresponding double Grothendieck polynomials constructed by Kirillov-Naruse [9] represent the KK-theoretic equivariant Schubert classes.

3.3 Definitions of GΘλ{G\Theta}_{\lambda} and GΘλ{G\Theta^{\prime}}_{\lambda}

Definition 3.5.

We introduce the functions GkΘm(x,a)GΓ[[a1,,ak]]gr{}_{k}{G\Theta}_{m}(x,a)\in{G\Gamma}[[a_{1},\dots,a_{k}]]_{\operatorname{gr}} (m)(m\in{\mathbb{Z}}) by the following generating function

GkΘ(x,a;u):=mGkΘm(x,a)um=11+βu1i=11+(u+β)xi1+(u+β)x¯ii=1k(1+(u+β)ai).\displaystyle{}_{k}{G\Theta}(x,a;u):=\sum_{m\in{\mathbb{Z}}}{}_{k}{G\Theta}_{m}(x,a)u^{m}=\frac{1}{1+\beta u^{-1}}\prod_{i=1}^{\infty}\frac{1+(u+\beta)x_{i}}{1+(u+\beta)\bar{x}_{i}}\prod_{i=1}^{k}(1+(u+\beta)a_{i}).
Lemma 3.6.

The functions GkΘm(x,a){}_{k}{G\Theta}_{m}(x,a) are elements of (GΓ[β]a)W(k)({G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a})^{W_{(k)}} for all mm\in{\mathbb{Z}}.

Proof.

First, we observe that GkΘ(x,a;u){}_{k}{G\Theta}(x,a;u) satisfies the conditions (1)’ and (2)’. Thus GkΘm(x,a){}_{k}{G\Theta}_{m}(x,a) are in GΓ[β]a{G\Gamma}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{a}. Next we see that GkΘ(x,a;u){}_{k}{G\Theta}(x,a;u) is symmetric in a1,,aka_{1},\dots,a_{k} and hence GkΘm(x,a){}_{k}{G\Theta}_{m}(x,a) are invariant under the actions of s1a,,sk1as_{1}^{a},\dots,s_{k-1}^{a}. Furthermore, it is obvious from definition that s0a(GkΘ(x,a;u))=GkΘ(x,a;u)s_{0}^{a}({}_{k}{G\Theta}(x,a;u))={}_{k}{G\Theta}(x,a;u). Thus GkΘm(x,a){}_{k}{G\Theta}_{m}(x,a) are also invariant under the action of s0as_{0}^{a} (cf. [6, Proposition 5.1]). ∎

Lemma 3.6 allows us to define functions GkΘm()(x,a|b){}_{k}{G\Theta}_{m}^{(\ell)}(x,a|b) and GkΘm()(x,a|b){}_{k}{G\Theta^{\prime}}_{m}^{(\ell)}(x,a|b) in 𝒦(k){\mathcal{K}}_{\infty}^{(k)} as follows.

Definition 3.7.

We define the elements GkΘm()(x,a|b){}_{k}{G\Theta}_{m}^{(\ell)}(x,a|b) and GkΘm()(x,a|b){}_{k}{G\Theta^{\prime}}_{m}^{(\ell)}(x,a|b) in 𝒦(k){\mathcal{K}}_{\infty}^{(k)} (m,)(m,\ell\in{\mathbb{Z}}) by the following generating functions

mGkΘm()(x,a|b)um\displaystyle\sum_{m\in{\mathbb{Z}}}{}_{k}{G\Theta}_{m}^{(\ell)}(x,a|b)u^{m} =\displaystyle= {GkΘ(x,a;u)i=1||11+(u+β)b¯i(<0),GkΘ(x,a;u)i=1(1+(u+β)bi)(0),\displaystyle\begin{cases}{}_{k}{G\Theta}(x,a;u)\displaystyle\prod_{i=1}^{|\ell|}\dfrac{1}{1+(u+\beta)\bar{b}_{i}}&(\ell<0),\\ {}_{k}{G\Theta}(x,a;u)\displaystyle\prod_{i=1}^{\ell}(1+(u+\beta)b_{i})&(\ell\geq 0),\end{cases}
mGkΘm()(x,a|b)um\displaystyle\sum_{m\in{\mathbb{Z}}}{}_{k}{G\Theta^{\prime}}_{m}^{(\ell)}(x,a|b)u^{m} =\displaystyle= {GkΘ(x,a;u)i=1||11+(u+β)b¯i(<0),12+βu1GkΘ(x,a;u)i=1(1+(u+β)bi)(0).\displaystyle\begin{cases}{}_{k}{G\Theta}(x,a;u)\displaystyle\prod_{i=1}^{|\ell|}\dfrac{1}{1+(u+\beta)\bar{b}_{i}}&(\ell<0),\\ \displaystyle\frac{1}{2+\beta u^{-1}}{}_{k}{G\Theta}(x,a;u)\displaystyle\prod_{i=1}^{\ell}(1+(u+\beta)b_{i})&(\ell\geq 0).\end{cases}

We denote GkXm()=GkΘm()(x,a|b){}_{k}{{GX}}_{m}^{(\ell)}={}_{k}{G\Theta}_{m}^{(\ell)}(x,a|b) for type C and GkXm()=GkΘm()(x,a|b){}_{k}{{GX}}_{m}^{(\ell)}={}_{k}{G\Theta^{\prime}}_{m}^{(\ell)}(x,a|b) for type BB.

Remark 3.8.

A direct computation shows that GkΘm()(x,a|b)=(β)m{}_{k}{G\Theta}_{m}^{(\ell)}(x,a|b)=(-\beta)^{-m} for each m0m\leq 0. Moreover, if 0\ell\geq 0, we have

GkΘm()(x,a|b)=12s0(β2)sGkΘm+s()(x,a|b).{}_{k}{G\Theta^{\prime}}_{m}^{(\ell)}(x,a|b)=\frac{1}{2}\sum_{s\geq 0}\left(\displaystyle\frac{-\beta}{2}\right)^{s}{}_{k}{G\Theta}_{m+s}^{(\ell)}(x,a|b).

In order to define GΘλ{G\Theta}_{\lambda} and GΘλ{G\Theta^{\prime}}_{\lambda} in terms of Pfaffians, we prepare some notations related to kk-strict partitions.

Definition 3.9.

For λ𝒮𝒫k\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}, we set

C(λ)\displaystyle C(\lambda) :=\displaystyle:= {(i,j)2| 1i<j,λi+λj>2k+ji},\displaystyle\Big{\{}(i,j)\in{\mathbb{N}}^{2}\ |\ 1\leq i<j,\ \ \lambda_{i}+\lambda_{j}>2k+j-i\Big{\}},
γj\displaystyle\gamma_{j} :=\displaystyle:= {i|(i,j)C(λ)}for each j>0.\displaystyle\sharp\Big{\{}i\in{\mathbb{N}}\ |\ (i,j)\in C(\lambda)\Big{\}}\ \ \ \ \mbox{for each $j>0$}.

and define the associated characteristic index χ=(χ1,χ2,)\chi=(\chi_{1},\chi_{2},\dots) by

χj:=λjj+γjk.\chi_{j}:=\lambda_{j}-j+\gamma_{j}-k.
Remark 3.10.

For k=0k=0, a kk-strict partition λ\lambda is called a strict partition. In this case, one has χi=λi1\chi_{i}=\lambda_{i}-1 for all i=1,,ri=1,\dots,r, where rr is the length of λ\lambda.

Remark 3.11.

We have χi+χj0\chi_{i}+\chi_{j}\geq 0 if and only if λi+λj>2k+ji\lambda_{i}+\lambda_{j}>2k+j-i for i<ji<j (see [6, Lemma 3.3]). As a consequence one has C(λ)={(i,j)| 1i<j,χi+χj0}C(\lambda)=\{(i,j)\ |\ 1\leq i<j,\ \ \chi_{i}+\chi_{j}\geq 0\}.

Definition 3.12.

Let λ\lambda be a kk-strict partition of length rr. We set

Δr\displaystyle\Delta_{r} :=\displaystyle:= {(i,j)| 1i<jr}\displaystyle\{(i,j)\ |\ 1\leq i<j\leq r\}
D(λ)\displaystyle D(\lambda) :=\displaystyle:= {(i,j)Δr|χi+χj<0}=Δr\C(λ).\displaystyle\Big{\{}(i,j)\in\Delta_{r}\ |\ \chi_{i}+\chi_{j}<0\Big{\}}=\Delta_{r}\backslash C(\lambda).

For each ID(λ)I\subset D(\lambda), we set

aiI:={j|(i,j)I},cjI:={i|(i,j)I},anddiI:=aiIciI.a_{i}^{I}:=\sharp\{j\ |\ (i,j)\in I\},\ \ \ c_{j}^{I}:=\sharp\{i\ |\ (i,j)\in I\},\ \ \ \mbox{and}\ \ \ d_{i}^{I}:=a_{i}^{I}-c_{i}^{I}.

Denote m:=rm:=r if rr is even and m:=r+1m:=r+1 if rr is odd. Consider the following rational function of variables tit_{i} and tjt_{j} (1i,jm1\leq i,j\leq m): recall t¯=t1+βt\bar{t}=\frac{-t}{1+\beta t} and set

Fi,jI(t):=1(1+βti)miciI11(1+βtj)mjcjI1t¯i/t¯j1ti/t¯j.F_{i,j}^{I}(t):=\frac{1}{(1+\beta t_{i})^{m-i-c_{i}^{I}-1}}\frac{1}{(1+\beta t_{j})^{m-j-c_{j}^{I}}}\frac{1-\bar{t}_{i}/\bar{t}_{j}}{1-t_{i}/\bar{t}_{j}}.

Let fpqij,I{f}_{pq}^{ij,I} be the coefficient of the expansion of Fi,jI(t)F_{i,j}^{I}(t) as the following Laurent series

Fi,jI(t)=p0,p+q0fpqij,Itiptjq.F_{i,j}^{I}(t)=\sum_{p\geq 0,\atop{p+q\geq 0}}{f}_{pq}^{ij,I}t_{i}^{p}t_{j}^{q}. (3.1)
Definition 3.13.

Let λ𝒮𝒫k(n)\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n) of length rr with its characteristic index χ\chi. Let m:=rm:=r if rr is even and m:=r+1m:=r+1 if rr is odd. We define the element GkXλ𝒦(k){}_{k}{{GX}}_{\lambda}\in{\mathcal{K}}_{\infty}^{(k)} for type C and B by

GkXλ=ID(λ)Pf(p,qp0,p+q0fpqij,IGkXλi+diI+p(χi)GkXλj+djI+q(χj))1i<jm.{}_{k}{{GX}}_{\lambda}=\sum_{I\subset D(\lambda)}{\operatorname{Pf}}\left(\sum_{p,q\in{\mathbb{Z}}\atop{p\geq 0,p+q\geq 0}}{f}_{pq}^{ij,I}{}_{k}{{GX}}_{\lambda_{i}+d_{i}^{I}+p}^{(\chi_{i})}\cdot{}_{k}{{GX}}_{\lambda_{j}+d_{j}^{I}+q}^{(\chi_{j})}\right)_{1\leq i<j\leq m}.

We denote GkΘλ(x,a|b):=GkXλ(x,a|b){}_{k}{G\Theta}_{\lambda}(x,a|b):={}_{k}{{GX}}_{\lambda}(x,a|b) for type C and GkΘλ(x,a|b):=GkXλ(x,a|b){}_{k}{G\Theta^{\prime}}_{\lambda}(x,a|b):={}_{k}{{GX}}_{\lambda}(x,a|b) for type B. If there is no fear of confusion, we denote GkXλ{}_{k}{{GX}}_{\lambda} simply by GXλ{{GX}}_{\lambda}.

Example 3.14.

For a kk-strict partition λ=(λ1)\lambda=(\lambda_{1}) of length 11, the corresponding characteristic index is χ=(λ1k1)\chi=(\lambda_{1}-k-1). In this case, we have GkΘ(λ1)(x,a|b)=GkΘλ1(λ1k1)(x,a|b){}_{k}{G\Theta}_{(\lambda_{1})}(x,a|b)={}_{k}{G\Theta}_{\lambda_{1}}^{(\lambda_{1}-k-1)}(x,a|b) and GkΘ(λ1)(x,a|b)=GkΘλ1(λ1k1)(x,a|b){}_{k}{G\Theta^{\prime}}_{(\lambda_{1})}(x,a|b)={}_{k}{G\Theta^{\prime}}_{\lambda_{1}}^{\,(\lambda_{1}-k-1)}(x,a|b).

4 KK-theoretic Schubert classes of IGk(n){{IG}}^{k}(n)

4.1 Preliminary on equivariant connective KK-theory

First of all, let us briefly recall the some facts about torus equivariant connective KK-theory and the KK-theoretic Segre classes which will be needed later. For more details readers are referred to [11], [5, Section 2] and, for an alternative presentation which has no restriction on the characteristic of the base field, to [1, Appendix].

Let TnT_{n} be a standard algebraic torus (𝔽×)n({\mathbb{F}}^{\times})^{n}. For a smooth variety XX with an action of TnT_{n}, its equivariant connective KK-theory is denoted by CKTn(X){C\!K}^{*}_{T_{n}}(X). In this paper, we work with the rational coefficients {\mathbb{Q}} instead of the integral coefficients {\mathbb{Z}}. Let EE be a TnT_{n}-equivariant vector bundles, then its ii-th TnT_{n}-equivariant Chern class is denoted by ci(E)c_{i}(E), while its total Chern class of EE is denoted by c(E;u)=i=0rankEci(E)uic(E;u)=\sum_{i=0}^{{\operatorname{rank}}E}c_{i}(E)u^{i}. If FF is another TnT_{n}-equivariant vector bundle, then the total Chern class of the virtual bundle EFE-F is defined by c(EF;u)=i0ci(EF)ui=c(E;u)/c(F;u)c(E-F;u)=\sum_{i\geq 0}c_{i}(E-F)u^{i}=c(E;u)/c(F;u).

The equivariant connective KK-theory CKTn(X){C\!K}^{*}_{T_{n}}(X) is in fact a graded algebra over CKTn(pt){C\!K}^{*}_{T_{n}}(pt) which we identify with a ring of graded formal power series as follows. We regard ε1,,εn\varepsilon_{1},\dots,\varepsilon_{n} as the standard basis of the character group of TnT_{n}. Let LiL_{i} be the one dimensional representation of TnT_{n} with character εi\varepsilon_{i}. We use the following isomorphism

CKTn(pt)[β][[b1,,bn]]gr;c1(Li)bi{C\!K}^{*}_{T_{n}}(pt)\to{\mathbb{Q}}[\beta][[b_{1},\dots,b_{n}]]_{\operatorname{gr}};\ \ c_{1}(L_{i})\mapsto b_{i} (4.1)

of graded [β]{\mathbb{Q}}[\beta]-algebras ([11, §2.6]). For simplicity, we denote CKTn:=CKTn(pt){C\!K}^{*}_{T_{n}}:={C\!K}^{*}_{T_{n}}(pt). We regard a CKTn{C\!K}^{*}_{T_{n}}-algebra as a [β][[b]]gr{\mathbb{Q}}[\beta][[b]]_{\operatorname{gr}}-algebra or b{\mathcal{R}}_{b}-algebra via the projection [β][[b]]grCKTn{\mathbb{Q}}[\beta][[b]]_{\operatorname{gr}}\to{C\!K}^{*}_{T_{n}} or bCKTn{\mathcal{R}}_{b}\to{C\!K}^{*}_{T_{n}} defined by bi=0b_{i}=0 for all i>ni>n respectively.

Remark 4.1.

The ring CKTn=[β][[b1,,bn]]gr{C\!K}^{*}_{T_{n}}={\mathbb{Q}}[\beta][[b_{1},\dots,b_{n}]]_{\operatorname{gr}} specializes to KTn(pt)=[[b1,,bn]]K_{T_{n}}({pt})={\mathbb{Q}}[[b_{1},\dots,b_{n}]] at β=1\beta=-1. It can be also identified with a completion of the representation ring R(Tn)R(T_{n}) of TnT_{n} with rational coefficients as follows. First recall that we have

R(Tn)=[e±ε1,,e±εn],R(T_{n})\otimes_{{\mathbb{Z}}}{\mathbb{Q}}={\mathbb{Q}}[e^{\pm\varepsilon_{1}},\dots,e^{\pm\varepsilon_{n}}],

where the class of LiL_{i} corresponds to eεie^{-\varepsilon_{i}}. Now 1eεi1-e^{\varepsilon_{i}} is identified with the first Chern class bi=c1(Li)b_{i}=c_{1}(L_{i}) of LiL_{i} (cf. Krishna [12, Theorem 7.3]).

Although the Segre classes can be defined geometrically, we take the following definition in terms of a generating function due to [5].

Definition 4.2.

For a virtual bundle EFE-F, we define the TnT_{n}-equivariant relative Segre classes 𝒮m(EF){\mathscr{S}}_{m}(E-F) (m)(m\in{\mathbb{Z}}) in CKTn(X){C\!K}^{*}_{T_{n}}(X) by

𝒮(EF;u):=m𝒮m(EF)um=11+βu1c(EF;β)c(EF;u).{\mathscr{S}}(E-F;u):=\sum_{m\in{\mathbb{Z}}}{\mathscr{S}}_{m}(E-F)u^{m}=\frac{1}{1+\beta u^{-1}}\frac{c(E-F;\beta)}{c(E-F;-u)}. (4.2)
Remark 4.3.

The Chern classes of the connective KK-theory are governed essentially by the multiplicative formal group law xy=x+y+βxyx\oplus y=x+y+\beta xy. Namely if LL and MM are line bundles, then c1(LM)=c1(L)c1(M)c_{1}(L\otimes M)=c_{1}(L)\oplus c_{1}(M) and c1(L)=c1(L)1+βc1(L)c_{1}(L^{\vee})=\frac{-c_{1}(L)}{1+\beta c_{1}(L)}. This being said, the identity 1+βx1xu=11+(u+β)x\frac{1+\beta x}{1-xu}=\frac{1}{1+(u+\beta)x} implies

𝒮(EF;u)=11+βu1c(F;u+β)c(E;u+β).{\mathscr{S}}(E-F;u)=\frac{1}{1+\beta u^{-1}}\frac{c(F^{\vee};u+\beta)}{c(E^{\vee};u+\beta)}.

4.2 Schubert varieties and the stability of their classes

For an integer nkn\geq k, let E(n)E^{(n)} be a vector space 𝔽2n{\mathbb{F}}^{2n} or 𝔽2n+1{\mathbb{F}}^{2n+1} of dimension 2n2n or 2n+12n+1 respectively. We fix bases by

𝔽2n=Span{𝒆i¯,𝒆i| 1in} and 𝔽2n+1=Span{𝒆i¯,𝒆i| 1in}Span{𝒆0},{\mathbb{F}}^{2n}=\operatorname{Span}\{\boldsymbol{e}_{\bar{i}},\boldsymbol{e}_{i}\>|\;1\leq i\leq n\}\ \ \ \ \mbox{ and }\ \ \ {\mathbb{F}}^{2n+1}=\operatorname{Span}\{\boldsymbol{e}_{\bar{i}},\boldsymbol{e}_{i}\>|\;1\leq i\leq n\}\oplus\operatorname{Span}\{\boldsymbol{e}_{0}\},

together with the symplectic form i=1n𝒆i𝒆i¯\sum_{i=1}^{n}\boldsymbol{e}_{i}^{*}\wedge\boldsymbol{e}_{\bar{i}}^{*} and the non-degenerate symmetric form 𝒆0𝒆0+i=1n𝒆i𝒆i¯\boldsymbol{e}_{0}^{*}\otimes\boldsymbol{e}_{0}^{*}+\sum_{i=1}^{n}\boldsymbol{e}_{i}^{*}\otimes\boldsymbol{e}_{\bar{i}}^{*} respectively where {𝒆i}\{\boldsymbol{e}_{i}^{*}\} denotes the dual basis of {𝒆i}\{\boldsymbol{e}_{i}\}. We define the action of TnT_{n} on E(n)E^{(n)} by setting the weights of 𝒆i\boldsymbol{e}_{i} and 𝒆i¯\boldsymbol{e}_{\bar{i}} to be εi\varepsilon_{i} and εi-\varepsilon_{i} respectively for 1in1\leq i\leq n, while the weight of 𝒆0\boldsymbol{e}_{0} is 0. This identifies TnT_{n} with maximal tori of Sp2n(𝔽){Sp}_{2n}({\mathbb{F}}) and SO2n+1(𝔽){{SO}}_{2n+1}({\mathbb{F}}) respectively.

Let IGk(n){{IG}}^{k}(n) be the Grassmannians of nkn-k dimensional isotropic subspaces in E(n)E^{(n)}, i.e. IGk(n){{IG}}^{k}(n) is the symplectic Grassmannian SGk(n)SG^{k}(n) if E(n)=𝔽2nE^{(n)}={\mathbb{F}}^{2n} (type C) and the odd orthogonal Grassmannian OGk(n){OG}^{k}(n) if E(n)=𝔽2n+1E^{(n)}={\mathbb{F}}^{2n+1} (type B). Consider the subspaces of E(n)E^{(n)}

F=Span{𝒆n,,𝒆+1}(0n),\displaystyle F^{\ell}=\operatorname{Span}\{\boldsymbol{e}_{n},\dots,\boldsymbol{e}_{\ell+1}\}\ \ \ (0\leq\ell\leq n),
F=(F0)Span{𝒆1¯,,𝒆¯}(1n).\displaystyle F^{-\ell}=(F^{0})^{\perp}\oplus\operatorname{Span}\{\boldsymbol{e}_{\bar{1}},\cdots,\boldsymbol{e}_{\bar{\ell}}\}\ \ \ (1\leq\ell\leq n).

It is known that the Schubert varieties in IGk(n){{IG}}^{k}(n) are described as follows. For a kk-strict partition λ𝒮𝒫k(n)\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n) with length rr and characteristic index χ\chi, the associated Schubert variety ΩλX\Omega_{\lambda}^{X} in IGk(n){{IG}}^{k}(n) is given by

ΩλX={UIGk(n)|dim(UFχi)i,i=1,,r},\Omega_{\lambda}^{X}=\{U\in{{IG}}^{k}(n)\ |\ \dim(U\cap F^{\chi_{i}})\geq i,\ \ \ i=1,\dots,r\},

where we write X=CX=C for the symplectic case, and X=BX=B for the odd orthogonal case.

Since the Schubert variety ΩλX\Omega_{\lambda}^{X} is TnT_{n}-stable, it defines the TnT_{n}-equivariant class [ΩλX]Tn[\Omega_{\lambda}^{X}]_{T_{n}} in CKTn(IGk(n)){C\!K}^{*}_{T_{n}}({{IG}}^{k}(n)). As a CKTn{C\!K}^{*}_{T_{n}}-module, CKTn(IGk(n)){C\!K}^{*}_{T_{n}}({{IG}}^{k}(n)) is freely generated by [ΩλX]Tn,λ𝒮𝒫k(n)[\Omega_{\lambda}^{X}]_{T_{n}},\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n). See [11].

Let E(n)E(n+1)E^{(n)}\to E^{(n+1)} be the injective linear map defined by the inclusion of the basis elements. It induces an embedding jn:IGk(n)IGk(n+1)j_{n}:{{IG}}^{k}(n)\to{{IG}}^{k}(n+1), which is equivariant with respect to the corresponding inclusion TnTn+1T_{n}\to T_{n+1}. Consider its pullback

jn:CKTn+1(IGk(n+1))CKTn(IGk(n)),j_{n}^{*}:{C\!K}^{*}_{T_{n+1}}({{IG}}^{k}(n+1))\to{C\!K}^{*}_{T_{n}}({{IG}}^{k}(n)),

and then we have

jn[ΩλX]Tn+1={[ΩλX]Tn if λ𝒮𝒫k(n),0 if λ𝒮𝒫k(n).j_{n}^{*}[\Omega_{\lambda}^{X}]_{T_{n+1}}=\begin{cases}[\Omega_{\lambda}^{X}]_{T_{n}}&\mbox{ if }\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n),\\ 0&\mbox{ if }\lambda\not\in{\mathcal{S}}{\mathcal{P}}^{k}(n).\end{cases} (4.3)

Consider the graded projective limit with respect to jnj_{n}^{*}

𝕂T(IGk):=mlimnCKTnm(IGk(n)),{\mathbb{K}}^{\infty}_{T}({{IG}}^{k}):=\bigoplus_{m\in{\mathbb{Z}}}\lim_{\longleftarrow\atop{n}}{C\!K}^{m}_{T_{n}}({{IG}}^{k}(n)),

and then by (4.3) one obtains a unique element [ΩλX]T[\Omega_{\lambda}^{X}]_{T} as a limit of the classes [ΩλX]Tn[\Omega_{\lambda}^{X}]_{T_{n}}. Since the Schubert classes form a CKTn{C\!K}^{*}_{T_{n}}-module basis, we can conclude the following.

Lemma 4.4.

Any element ff of 𝕂T(IGk){\mathbb{K}}^{\infty}_{T}({{IG}}^{k}) can be expressed uniquely as a possibly infinite [β][[b]]gr{\mathbb{Q}}[\beta][[b]]_{\operatorname{gr}}-linear combination of the classes [ΩλX]T[\Omega_{\lambda}^{X}]_{T}:

f=λ𝒮𝒫kcλ[ΩλX]T,cλ[β][[b]]gr.f=\sum_{\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}}c_{\lambda}[\Omega_{\lambda}^{X}]_{T},\ \ \ \ \ c_{\lambda}\in{\mathbb{Q}}[\beta][[b]]_{\operatorname{gr}}.
Definition 4.5.

Let 𝕂Tfin(IGk){\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}) be the b{\mathcal{R}}_{b}-submodule of 𝕂T(IGk){\mathbb{K}}^{\infty}_{T}({{IG}}^{k}) consisting of the following possibly infinite b{\mathcal{R}}_{b}-linear combinations of the classes [ΩλX]T[\Omega_{\lambda}^{X}]_{T}

f=λ𝒮𝒫kcλ[ΩλX]T,cλb.f=\sum_{\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}}c_{\lambda}[\Omega_{\lambda}^{X}]_{T},\ \ \ \ \ c_{\lambda}\in{\mathcal{R}}_{b}.

It will be shown below that there is an isomorphism of b{\mathcal{R}}_{b}-algebras 𝒦(k)𝕂Tfin(IGk){\mathcal{K}}_{\infty}^{(k)}\cong{\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}).

4.3 Pfaffian formula of the equivariant Schubert classes

Let UU be the tautological isotropic bundle of IGk(n){{IG}}^{k}(n) of rank nkn-k. We abuse notation and denote by FiF^{i} the vector bundle over IGk(n){{IG}}^{k}(n) with the fiber FiF^{i}. We have

c(E/F;u)\displaystyle c(E/F^{\ell};u) =\displaystyle= i=1n(1+b¯iu)i=1(1+biu)(0n),\displaystyle\prod_{i=1}^{n}(1+\bar{b}_{i}u)\cdot\prod_{i=1}^{\ell}(1+b_{i}u)\ \ \ \ \ \ (0\leq\ell\leq n),
c(E/F;u)\displaystyle c(E/F^{-\ell};u) =\displaystyle= i=+1n(1+b¯iu)(1n),\displaystyle\prod_{i=\ell+1}^{n}(1+\bar{b}_{i}u)\ \ \ \ \ \ \ \ (1\leq\ell\leq n),

where EE denotes the vector bundle with fiber E(n)E^{(n)}. Let us point out that for the odd orthogonal case, the action of TnT_{n} on Span{𝒆0}\operatorname{Span}\{\boldsymbol{e}_{0}\} is trivial so that c((F0)/F0;u)=1c((F^{0})^{\perp}/F^{0};u)=1.

Definition 4.6.

We define the classes 𝒳m()CKTn(IGk(n)){\mathscr{X}}_{m}^{(\ell)}\in{C\!K}^{*}_{T_{n}}({{IG}}^{k}(n)) for mm\in{\mathbb{Z}} and =n,,n\ell=-n,\cdots,n as follows. For the symplectic case IGk(n)=SGk(n){{IG}}^{k}(n)=SG^{k}(n), we let

𝒳m():=𝒞m()=𝒮m(U(E/F)){\mathscr{X}}_{m}^{(\ell)}:={\mathscr{C}}_{m}^{(\ell)}={\mathscr{S}}_{m}(U^{\vee}-(E/F^{\ell})^{\vee})

and for the odd orthogonal case IGk(n)=OGk(n){{IG}}^{k}(n)={OG}^{k}(n), we let

𝒳m():=m()={𝒮m((UE/F))(n<0),12s0(β2)s𝒮m+s((UE/F))(0n).{\mathscr{X}}_{m}^{(\ell)}:={\mathscr{B}}_{m}^{(\ell)}=\begin{cases}{\mathscr{S}}_{m}((U-E/F^{\ell})^{\vee})&(-n\leq\ell<0),\\ \displaystyle\frac{1}{2}\sum_{s\geq 0}\left(\displaystyle\frac{-\beta}{2}\right)^{s}{\mathscr{S}}_{m+s}((U-E/F^{\ell})^{\vee})&(0\leq\ell\leq n).\end{cases}

Let us point out that by Definition 4.2 and Remark 4.3, we have

m𝒞m()um=11+βu1c(E/F;u+β)c(U;u+β).\sum_{m\in{\mathbb{Z}}}{\mathscr{C}}_{m}^{(\ell)}u^{m}=\frac{1}{1+\beta u^{-1}}\frac{c(E/F^{\ell};u+\beta)}{c(U;u+\beta)}. (4.4)
Theorem 4.7 ([5]).

Let λ\lambda be a kk-strict partition in 𝒮𝒫k(n){\mathcal{S}}{\mathcal{P}}^{k}(n) of length rr and χ\chi its characteristic index. In CKTn(IGk(n)){C\!K}^{*}_{T_{n}}({{IG}}^{k}(n)), the Schubert class [ΩλX]Tn[\Omega_{\lambda}^{X}]_{T_{n}} is given by

[ΩλX]Tn=ID(λ)Pf(p,qp0,p+q0fpqij,I𝒳λi+diI+p(χi)𝒳λj+djI+q(χj))1i<jm,[\Omega_{\lambda}^{X}]_{T_{n}}=\sum_{I\subset D(\lambda)}{\operatorname{Pf}}\left(\sum_{p,q\in{\mathbb{Z}}\atop{p\geq 0,p+q\geq 0}}{f}_{pq}^{ij,I}{\mathscr{X}}_{\lambda_{i}+d_{i}^{I}+p}^{(\chi_{i})}{\mathscr{X}}_{\lambda_{j}+d_{j}^{I}+q}^{(\chi_{j})}\right)_{1\leq i<j\leq m},

where m=rm=r if rr is even and m=r+1m=r+1 if rr is odd, and 𝒳i(n1):=(β)i{\mathscr{X}}_{-i}^{(-n-1)}:=(-\beta)^{i} for i0i\geq 0.

This theorem follows from [5, Theorem 5.20, 6.16]. Indeed, let BTnBT_{n} be the classifying space of TnT_{n} and ETnBTnET_{n}\rightarrow BT_{n} the universal bundle. Consider the bundle ETn×TnEET_{n}\times_{T_{n}}E over ETn×TnIGk(n)ET_{n}\times_{T_{n}}{{IG}}^{k}(n). We can apply Theorem 5.20 or 6.16 in [5] to every finite approximation of this bundle. Then the functoriality of Chern classes implies the claim.

5 GKM descriptions in algebra and geometry

We use techniques of the equivariant localization maps to study both rings 𝒦(k){\mathcal{K}}_{\infty}^{(k)} and 𝕂T(IGk){\mathbb{K}}^{\infty}_{T}({{IG}}^{k}). By using this we have so-called GKM (after Goresky–MacPherson–Kottwitz [4]) description for these rings. These maps also enable us to establish the connection between GΘ/GΘ{G\Theta}/{G\Theta^{\prime}}-functions and the stable limits of the torus equivariant Schubert classes.

5.1 GKM description for 𝒦(k){\mathcal{K}}_{\infty}^{(k)}

First we study the GKM description of the ring 𝒦{\mathcal{K}}_{\infty}. Let Fun(W,b)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b}) be the algebra of maps from WW_{\infty} to b{\mathcal{R}}_{b} whose algebra structure is naturally given by the one of b{\mathcal{R}}_{b}. An element ψ\psi in Fun(W,b)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b}) is denoted by (vψ(v))vW(v\mapsto\psi(v))_{v\in W_{\infty}}. We define an action of WW_{\infty} on Fun(W,b)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b}) by

w(ψ)(v):=ψ(vw),ψFun(W,b),w,vW.w(\psi)(v):=\psi(vw),\ \ \psi\in\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b}),w,v\in W_{\infty}.
Definition 5.1.

We introduce the following homomorphism of b{\mathcal{R}}_{b}-algebras

Φ:𝒦Fun(W,b);Φ(f):=(vΦv(f))vW,\Phi_{\infty}:{\mathcal{K}}_{\infty}\to\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b});\ \ \ \ \Phi_{\infty}(f):=(v\mapsto\Phi_{v}(f))_{v\in W_{\infty}},

where Φv:𝒦b\Phi_{v}:{\mathcal{K}}_{\infty}\to{\mathcal{R}}_{b} is the b{\mathcal{R}}_{b}-algebra homomorphism given by the substitution

xi{bv(i) if v(i)<00 if v(i)>0 and aibv(i).x_{i}\mapsto\begin{cases}b_{v(i)}&\mbox{ if $v(i)<0$}\\ 0&\mbox{ if $v(i)>0$}\end{cases}\ \ \ \ \mbox{ and }\ \ a_{i}\mapsto b_{-v(i)}.

where we set bm:=b¯mb_{m}:=\bar{b}_{-m} if m<0m<0.

Definition 5.2.

Recall that Δ+\Delta^{+} denotes the root system of type B and C in the lattice L=i=1εiL=\bigoplus_{i=1}^{\infty}{\mathbb{Z}}\varepsilon_{i} (see §2.1). We define a map e:Lbe:L\to{\mathcal{R}}_{b} by

e(εi)=bi,e(εi)=b¯i,e(α+γ)=e(α)e(γ) and e(αγ)=e(α)e(γ)(α,γL).e(\varepsilon_{i})=b_{i},\ \ e(-\varepsilon_{i})=\bar{b}_{i},\ \ e(\alpha+\gamma)=e(\alpha)\oplus e(\gamma)\ \mbox{ and }\ e(\alpha-\gamma)=e(\alpha)\ominus e(\gamma)\ \ \ \ (\alpha,\gamma\in L).

For each αΔn+\alpha\in\Delta^{+}_{n} one has e(α)CKTn[β][[b1,,bn]]gre(\alpha)\in{C\!K}^{*}_{T_{n}}\cong{\mathbb{Q}}[\beta][[b_{1},\dots,b_{n}]]_{\operatorname{gr}}.

Definition 5.3.

Let 𝔎{\mathfrak{K}}_{\infty} be the subalgebra of Fun(W,b)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b}) consisting of maps ψ\psi such that

ψ(sαv)ψ(v)e(α)b, for all vW and αΔ+.\psi(s_{\alpha}v)-\psi(v)\in e(\alpha)\cdot{\mathcal{R}}_{b},\ \ \mbox{ for all }v\in W_{\infty}\ \ \mbox{ and }\alpha\in\Delta^{+}.
Remark 5.4.

By the fact that e(2εi)=bibi=bi(2+βbi)e(2\varepsilon_{i})=b_{i}\oplus b_{i}=b_{i}(2+\beta b_{i}) and since 2+βbi2+\beta b_{i} is invertible in b{\mathcal{R}}_{b}, we can see that 𝔎n(k){\mathfrak{K}}_{n}^{(k)} is independent of the type B and C.

Lemma 5.5.

The map Φ\Phi_{\infty} is injective and its image lies in 𝔎Fun(W,b){\mathfrak{K}}_{\infty}\subset\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b}).

Proof.

The proof of the latter claim is similar to the one in the proof of Lemma 7.1. in [8]. We leave the details to the reader. Below we prove the injectivity.

By the definition of 𝒦{\mathcal{K}}_{\infty} and Proposition 3.3, a homogeneous element ff of 𝒦{\mathcal{K}}_{\infty} of degree dd can be uniquely written as

f=λ𝒮𝒫0cλ(a;b)GPλ(x),cλ(a;b)(a[β]b)d|λ|.f=\sum_{\lambda\in{\mathcal{S}}{\mathcal{P}}^{0}}c_{\lambda}(a;b){GP}_{\lambda}(x),\ \ \ c_{\lambda}(a;b)\in({\mathcal{R}}_{a}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{b})_{d-|\lambda|}. (5.1)

By the definition of a[β]b{\mathcal{R}}_{a}\otimes_{{\mathbb{Q}}[\beta]}{\mathcal{R}}_{b}, there exist m,nm,n such that for all λ𝒮𝒫0\lambda\in{\mathcal{S}}{\mathcal{P}}^{0}

cλ(a;b)=cλ(a1,,an;b1,,bm)[β][[a1,,an,b1,,bm]]d|λ|.c_{\lambda}(a;b)=c_{\lambda}(a_{1},\dots,a_{n};b_{1},\dots,b_{m})\in{\mathbb{Q}}[\beta][[a_{1},\dots,a_{n},b_{1},\dots,b_{m}]]_{d-|\lambda|}.

Now suppose that Φ(f)=0\Phi_{\infty}(f)=0. We choose an integer N>m+nN>m+n and consider an element vv of WW_{\infty} given by

v=(m+1,,m+n,1,,m,m+n+1¯,,m+n+N¯).v=(m+1,\dots,m+n,1,\dots,m,\overline{m+n+1},\dots,\overline{m+n+N}).

Applying Φv\Phi_{v} to (5.1) we obtain

Φv(f)=λ𝒮𝒫0cλ(b¯m+1,,b¯m+n;b1,,bm)GPλ(b¯m+n+1,,b¯m+n+N,0,0,)=0.\Phi_{v}(f)=\sum_{\lambda\in{\mathcal{S}}{\mathcal{P}}^{0}}c_{\lambda}(\bar{b}_{m+1},\dots,\bar{b}_{m+n};b_{1},\dots,b_{m}){GP}_{\lambda}(\bar{b}_{m+n+1},\dots,\bar{b}_{m+n+N},0,0,\dots)=0.

Finally we can conclude that cλ(b¯m+1,,b¯m+n;b1,,bn)=0c_{\lambda}(\bar{b}_{m+1},\dots,\bar{b}_{m+n};b_{1},\dots,b_{n})=0 for all λ𝒮𝒫0\lambda\in{\mathcal{S}}{\mathcal{P}}^{0} by the facts that GPλ(x1,,xN),λ𝒮𝒫N0{GP}_{\lambda}(x_{1},\dots,x_{N}),\lambda\in{\mathcal{S}}{\mathcal{P}}^{0}_{N} form a formal basis of GΓN{G\Gamma}_{N} (see the proof of Proposition 3.3) and that NN can be chosen arbitrary as long as it is greater than m+nm+n. This completes the proof. ∎

Next we will obtain the GKM description of 𝒦(k){\mathcal{K}}_{\infty}^{(k)}.

Definition 5.6.

Let 𝔎(k){\mathfrak{K}}^{(k)}_{\infty} be the subalgebra of Fun(𝒮𝒫k,b)\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k},{\mathcal{R}}_{b}) consisting of functions ψ\psi such that

ψ(sαμ)ψ(μ)e(α)b, for all μ𝒮𝒫k and αΔ+.\psi(s_{\alpha}\mu)-\psi(\mu)\in e(\alpha)\cdot{\mathcal{R}}_{b},\ \ \mbox{ for all }\mu\in{\mathcal{S}}{\mathcal{P}}^{k}\ \ \mbox{ and }\alpha\in\Delta^{+}.

The following proposition is essentially a consequence of Lemma 5.5.

Proposition 5.7.

The map Φ\Phi_{\infty} naturally induces an injective b{\mathcal{R}}_{b}-algebra homomorphism

Φ(k):𝒦(k)Fun(𝒮𝒫k,b)\Phi_{\infty}^{(k)}:{\mathcal{K}}_{\infty}^{(k)}\to\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k},{\mathcal{R}}_{b})

and its image lies in 𝔎(k){\mathfrak{K}}_{\infty}^{(k)}.

Proof.

Since the map Φ\Phi_{\infty} is WW_{\infty}-equivariant (cf. [7, Proposition 7.3]), we find that Φ\Phi_{\infty} restricts to the injective map 𝒦(k)Fun(W,b)W(k){\mathcal{K}}_{\infty}^{(k)}\to\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b})^{W_{(k)}} by taking the W(k)W_{\infty}^{(k)}-invariant parts. Now we can identify Fun(𝒮𝒫k,b)\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k},{\mathcal{R}}_{b}) with Fun(W,b)W(k)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b})^{W_{(k)}} as b{\mathcal{R}}_{b}-algebras as follows. For each λ𝒮𝒫k\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}, let wλw_{\lambda} be the corresponding kk-Grassmannian element in W(k)W_{\infty}^{(k)}. For each ψFun(𝒮𝒫k,b)\psi\in\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k},{\mathcal{R}}_{b}), the corresponding element of Fun(W,b)W(k)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b})^{W_{(k)}} is a map sending each vWv\in W_{\infty} to ψ(λ)\psi(\lambda) if vwλW(k)v\in w_{\lambda}W_{(k)}. Thus we obtain an injective b{\mathcal{R}}_{b}-algebra homomorphism Φ(k):𝒦(k)Fun(𝒮𝒫k,b)\Phi_{\infty}^{(k)}:{\mathcal{K}}_{\infty}^{(k)}\to\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k},{\mathcal{R}}_{b}). Furthermore, we can observe that the WW_{\infty}-action on Fun(W,b)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b}) preserves 𝔎{\mathfrak{K}}_{\infty} and notice that, under the identification Fun(W,b)W(k)Fun(𝒮𝒫k,b)\operatorname{Fun}(W_{\infty},{\mathcal{R}}_{b})^{W_{(k)}}\cong\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k},{\mathcal{R}}_{b}), the W(k)W_{(k)}-invariant part of 𝔎{\mathfrak{K}}_{\infty} coincides with 𝔎(k){\mathfrak{K}}_{\infty}^{(k)}. This completes the proof. ∎

5.2 Map Ψ(k)\Psi_{\infty}^{(k)} through GKM descriptions

First we recall the GKM description of CKTn(IGk(n)){C\!K}^{*}_{T_{n}}({{IG}}^{k}(n)) following [10, Corollary (3.20)] (cf. [11, Theorem 7.8 ]).

The set (IGk(n))Tn({{IG}}^{k}(n))^{T_{n}} of TnT_{n}-fixed points in IGk(n){{IG}}^{k}(n) is bijective to 𝒮𝒫k(n){\mathcal{S}}{\mathcal{P}}^{k}(n). For each λ𝒮𝒫k(n)\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n), let eλe_{\lambda} denote the corresponding fixed point. Let Fun(𝒮𝒫k(n),CKTn)\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k}(n),{C\!K}^{*}_{T_{n}}) be the algebra of maps from 𝒮𝒫k(n){\mathcal{S}}{\mathcal{P}}^{k}(n) to CKTn{C\!K}^{*}_{T_{n}} whose algebra structure is naturally given by the one of CKTn{C\!K}^{*}_{T_{n}}. We can identify CKTn((IGk(n))Tn){C\!K}^{*}_{T_{n}}(({{IG}}^{k}(n))^{T_{n}}) with Fun(𝒮𝒫k(n),CKTn)\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k}(n),{C\!K}^{*}_{T_{n}}) as graded CKTn{C\!K}^{*}_{T_{n}}-algebras. The inclusion ιn:(IGk(n))TnIGk(n)\iota_{n}:({{IG}}^{k}(n))^{T_{n}}\hookrightarrow{{IG}}^{k}(n) defines, by pull-back, an injective homomorphism of CKTn{C\!K}^{*}_{T_{n}}-algebras

ιn:CKTn(IGk(n))Fun(𝒮𝒫k(n),CKTn).\iota_{n}^{*}:{C\!K}^{*}_{T_{n}}({{IG}}^{k}(n))\to\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k}(n),{C\!K}^{*}_{T_{n}}).

Let 𝔎n(k){\mathfrak{K}}_{n}^{(k)} be the graded CKTn{C\!K}^{*}_{T_{n}}-subalgebra of Fun(𝒮𝒫k(n),CKTn)\operatorname{Fun}({\mathcal{S}}{\mathcal{P}}^{k}(n),{C\!K}^{*}_{T_{n}}) defined as follows. A map ψ:𝒮𝒫k(n)CKTn\psi:{\mathcal{S}}{\mathcal{P}}^{k}(n)\to{C\!K}^{*}_{T_{n}} is in 𝔎n(k){\mathfrak{K}}_{n}^{(k)} if and only if

ψ(sαμ)ψ(μ)e(α)CKTn for all μ𝒮𝒫k(n) and αΔn+\psi(s_{\alpha}\mu)-\psi(\mu)\in e(\alpha)\cdot{C\!K}^{*}_{T_{n}}\ \ \mbox{ for all }\mu\in{\mathcal{S}}{\mathcal{P}}^{k}(n)\ \mbox{ and }\ \alpha\in\Delta^{+}_{n}

where e(α)e(\alpha) was introduced at Definition 5.2. Then the image of the map ιn\iota_{n}^{*} coincides with 𝔎n(k){\mathfrak{K}}_{n}^{(k)}. In other words, we have the isomorphism of CKTn{C\!K}^{*}_{T_{n}}-algebras

ιn:CKTn(IGk(n))𝔎n(k).\iota_{n}^{*}:{C\!K}^{*}_{T_{n}}({{IG}}^{k}(n))\cong{\mathfrak{K}}_{n}^{(k)}. (5.2)

Now we construct an injective homomorphism from 𝒦(k){\mathcal{K}}_{\infty}^{(k)} to 𝕂T(IGk){\mathbb{K}}^{\infty}_{T}({{IG}}^{k}) using the injective homomorphism Φ(k)\Phi^{(k)}_{\infty} given in Proposition 5.7 together with the isomorphism ιn\iota_{n}^{*} at (5.2).

Proposition 5.8.

There is an injective homomorphism of graded b{\mathcal{R}}_{b}-algebras

Ψ(k):𝒦(k)𝕂T(IGk).\Psi_{\infty}^{(k)}:{\mathcal{K}}_{\infty}^{(k)}\to{\mathbb{K}}^{\infty}_{T}({{IG}}^{k}).
Proof.

Below all the maps are considered as homomorphisms of graded b{\mathcal{R}}_{b}-algebras. There is a natural map 𝔎(k)𝔎n(k){\mathfrak{K}}^{(k)}_{\infty}\to{\mathfrak{K}}^{(k)}_{n} defined by restricting the domain of each function of 𝔎(k){\mathfrak{K}}^{(k)}_{\infty} from 𝒮𝒫k{\mathcal{S}}{\mathcal{P}}^{k} to 𝒮𝒫k(n){\mathcal{S}}{\mathcal{P}}^{k}(n) and projecting its values from b{\mathcal{R}}_{b} to CKTn{C\!K}^{*}_{T_{n}}. Similarly, one has maps 𝔎n+1(k)𝔎n(k){\mathfrak{K}}^{(k)}_{n+1}\to{\mathfrak{K}}^{(k)}_{n} for all nn. By the commutativity of these maps, we obtain an injection

𝔎(k)limn𝔎n(k),{\mathfrak{K}}^{(k)}_{\infty}\to\lim_{\longleftarrow\atop{n}}{\mathfrak{K}}^{(k)}_{n},

where the limit on the right hand side is the direct sum of the projective limits of each graded piece. On the other hand, the isomorphism ιn\iota_{n}^{*} naturally induces an isomorphism

𝕂T(IGk)limn𝔎n(k).{\mathbb{K}}^{\infty}_{T}({{IG}}^{k})\cong\lim_{\longleftarrow\atop{n}}{\mathfrak{K}}^{(k)}_{n}.

Composing the above maps with Φ(k)\Phi_{\infty}^{(k)}, we obtain the desired injective homomorphism:

Ψ(k):𝒦(k)Φ(k)𝔎(k)limn𝔎n(k)𝕂T(IGk).\Psi_{\infty}^{(k)}:{\mathcal{K}}_{\infty}^{(k)}\stackrel{{\scriptstyle\Phi_{\infty}^{(k)}}}{{\longrightarrow}}{\mathfrak{K}}^{(k)}_{\infty}\longrightarrow\lim_{\longleftarrow\atop{n}}{\mathfrak{K}}^{(k)}_{n}\cong{\mathbb{K}}^{\infty}_{T}({{IG}}^{k}).

5.3 GΘλ{G\Theta}_{\lambda} and GΘλ{G\Theta^{\prime}}_{\lambda} represent Schubert classes

Recall the following notations.

IGk(n)𝒳m()ΩλXGkXm()GkXλX=CSGk(n)𝒞m()ΩλCGkΘm()GkΘλX=BOGk(n)m()ΩλBGkΘm()GkΘλ\begin{array}[]{c|c|c|c|c|c}&{{IG}}^{k}(n)&{\mathscr{X}}_{m}^{(\ell)}&\Omega_{\lambda}^{X}&{}_{k}{{GX}}_{m}^{(\ell)}&{}_{k}{{GX}}_{\lambda}\\ \hline\cr X=C&SG^{k}(n)&{\mathscr{C}}_{m}^{(\ell)}&\Omega_{\lambda}^{C}&{}_{k}{G\Theta}_{m}^{(\ell)}&{}_{k}{G\Theta}_{\lambda}\\ \hline\cr X=B&{OG}^{k}(n)&{\mathscr{B}}_{m}^{(\ell)}&\Omega_{\lambda}^{B}&{}_{k}{G\Theta^{\prime}}_{m}^{\,(\ell)}&{}_{k}{G\Theta^{\prime}}_{\!\!\lambda}\end{array}

Define the b{\mathcal{R}}_{b}-algebra homomorphisms Ψn(k)\Psi_{n}^{(k)} for n1n\geq 1 by compositions:

Ψn(k):𝒦(k)Φ(k)𝔎(k)𝔎n(k)CKTn(IGk(n)).\Psi_{n}^{(k)}:{\mathcal{K}}_{\infty}^{(k)}\stackrel{{\scriptstyle\Phi_{\infty}^{(k)}}}{{\longrightarrow}}{\mathfrak{K}}^{(k)}_{\infty}\longrightarrow{\mathfrak{K}}^{(k)}_{n}\cong{C\!K}^{*}_{T_{n}}({{IG}}^{k}(n)).

Since Ψn(k)=jnΨn+1(k)\Psi_{n}^{(k)}=j_{n}^{*}\circ\Psi_{n+1}^{(k)}, they induce a map 𝒦(k)𝕂T(IGk){\mathcal{K}}_{\infty}^{(k)}\to{\mathbb{K}}^{\infty}_{T}({{IG}}^{k}) which coincides with Ψ(k)\Psi_{\infty}^{(k)}.

Lemma 5.9.

For nn-n\leq\ell\leq n and mm\in{\mathbb{Z}}, we have Ψn(k)(GkXm()(x,a|b))=𝒳m()\Psi_{n}^{(k)}({}_{k}{{GX}}_{m}^{(\ell)}(x,a|b))={\mathscr{X}}_{m}^{(\ell)}. Furthermore, we have

Ψn(k)(GXλ(x,a|b))={[ΩλX]Tnif λ𝒮𝒫k(n)0if λ𝒮𝒫k(n).\Psi_{n}^{(k)}({{GX}}_{\lambda}(x,a|b))=\begin{cases}[\Omega_{\lambda}^{X}]_{T_{n}}&\mbox{if $\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n)$}\\ 0&\mbox{if $\lambda\not\in{\mathcal{S}}{\mathcal{P}}^{k}(n)$.}\end{cases} (5.3)
Proof.

The first claim is a generalization of [7, Lemma 10.3]. The proof is analogous and it follows from the comparison of the localizations at μ𝒮𝒫k(n)\mu\in{\mathcal{S}}{\mathcal{P}}^{k}(n). Let wμw_{\mu} be the element of Wn(k)W_{n}^{(k)} corresponding to μ\mu and suppose that its one line notation is given as (2.2). Let ιμ:CKTn(IGk(n))CKTn\iota_{\mu}^{*}:{C\!K}^{*}_{T_{n}}({{IG}}^{k}(n))\to{C\!K}^{*}_{T_{n}} be the pullback of the inclusion ιμ:{eμ}IGk(n)\iota_{\mu}:\{e_{\mu}\}\to{{IG}}^{k}(n) and Φwμ\Phi_{w_{\mu}} the map introduced at Definition 5.1. We prove for the symplectic case. It suffices to show that

Φwμ(mGkΘm()um)=ιμ(m𝒞m()um).\Phi_{w_{\mu}}\left(\sum_{m\in{\mathbb{Z}}}{}_{k}{G\Theta}_{m}^{(\ell)}u^{m}\right)=\iota_{\mu}^{*}\left(\sum_{m\in{\mathbb{Z}}}{\mathscr{C}}_{m}^{(\ell)}u^{m}\right). (5.4)

We have

Φwμ(mGkΘm()um)\displaystyle\Phi_{w_{\mu}}\left(\sum_{m\in{\mathbb{Z}}}{}_{k}{G\Theta}_{m}^{(\ell)}u^{m}\right)
=\displaystyle= {11+βu1i=1s1+(u+β)b¯ζi1+(u+β)bζii=1k(1+(u+β)b¯vi)i=1(1+(u+β)bi)(0),11+βu1i=1s1+(u+β)b¯ζi1+(u+β)bζii=1k(1+(u+β)b¯vi)i=1||11+(u+β)b¯i(0).\displaystyle\begin{cases}\dfrac{1}{1+\beta u^{-1}}\prod_{i=1}^{s}\dfrac{1+(u+\beta)\bar{b}_{\zeta_{i}}}{1+(u+\beta)b_{\zeta_{i}}}\prod_{i=1}^{k}(1+(u+\beta)\bar{b}_{v_{i}})\prod_{i=1}^{\ell}(1+(u+\beta)b_{i})&(\ell\geq 0),\\ \dfrac{1}{1+\beta u^{-1}}\prod_{i=1}^{s}\dfrac{1+(u+\beta)\bar{b}_{\zeta_{i}}}{1+(u+\beta)b_{\zeta_{i}}}\prod_{i=1}^{k}(1+(u+\beta)\bar{b}_{v_{i}})\prod_{i=1}^{|\ell|}\dfrac{1}{1+(u+\beta)\bar{b}_{i}}&(\ell\leq 0).\end{cases}

On the other hand, if 𝐳k+1,,𝐳n{\mathbf{z}}_{k+1},\dots,{\mathbf{z}}_{n} are the Chern roots of UU, by (4.4) we have

m𝒞m()um\displaystyle\sum_{m\in{\mathbb{Z}}}{\mathscr{C}}_{m}^{(\ell)}u^{m} =\displaystyle= {11+βu1i=1n(1+(u+β)b¯i)i=k+1n(1+(u+β)𝐳i)i=1(1+(u+β)bi)(0),11+βu1i=1n(1+(u+β)b¯i)i=k+1n(1+(u+β)𝐳i)i=1||11+(u+β)b¯i(0).\displaystyle\begin{cases}\dfrac{1}{1+\beta u^{-1}}\dfrac{\prod_{i=1}^{n}(1+(u+\beta)\bar{b}_{i})}{\prod_{i=k+1}^{n}(1+(u+\beta)\mathbf{z}_{i})}\prod_{i=1}^{\ell}(1+(u+\beta)b_{i})&(\ell\geq 0),\\ \dfrac{1}{1+\beta u^{-1}}\dfrac{\prod_{i=1}^{n}(1+(u+\beta)\bar{b}_{i})}{\prod_{i=k+1}^{n}(1+(u+\beta)\mathbf{z}_{i})}\prod_{i=1}^{|\ell|}\dfrac{1}{1+(u+\beta)\bar{b}_{i}}&(\ell\leq 0).\end{cases}

Now we use the identity

ιμ(c(U;u))=i=1s(1+bζiu)i=1nks(1+b¯uiu)\iota_{\mu}^{*}(c(U;u))=\prod_{i=1}^{s}(1+b_{\zeta_{i}}u)\cdot\prod_{i=1}^{n-k-s}(1+\bar{b}_{u_{i}}u)

which follows from (the proof of) Proposition 10.1 [7] and obtain (5.4). The proof for the case IGk(n)=OGk(n){{IG}}^{k}(n)={OG}^{k}(n) is similar.

Finally we show the latter claim. If λ𝒮𝒫k(n)\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n), then [ΩλX]Tn[\Omega_{\lambda}^{X}]_{T_{n}} and GXλ{{GX}}_{\lambda} are given by the same Pfaffian formulas except that the entries are given in terms of 𝒳mi(χi){\mathscr{X}}_{m_{i}}^{(\chi_{i})} or GkXmi(χi){}_{k}{{GX}}_{m_{i}}^{(\chi_{i})} respectively. Therefore (5.3) for the case when λ𝒮𝒫k(n)\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n) follows from the former claim. For the vanishing, it suffices to show the case when λ𝒮𝒫k(n+1)\𝒮𝒫k(n)\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}(n+1)\backslash{\mathcal{S}}{\mathcal{P}}^{k}(n). In that case, by Ψn(k)=jnΨn+1(k)\Psi_{n}^{(k)}=j_{n}^{*}\circ\Psi_{n+1}^{(k)}, we have Ψn(k)(GXλ(x,a|b))=jn[ΩλX]Tn+1\Psi_{n}^{(k)}({{GX}}_{\lambda}(x,a|b))=j_{n}^{*}[\Omega_{\lambda}^{X}]_{T_{n+1}} which is 0 by (4.3). ∎

Theorem 5.10.

The map Ψ(k)\Psi_{\infty}^{(k)} restricts to an b{\mathcal{R}}_{b}-algebra isomorphism

𝒦(k)𝕂Tfin(IGk){\mathcal{K}}_{\infty}^{(k)}\cong{\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k})

which sends GkXλ{}_{k}{{GX}}_{\lambda} to [ΩλX]T[\Omega_{\lambda}^{X}]_{T}. In particular, 𝕂Tfin(IGk){\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}) is an b{\mathcal{R}}_{b}-algebra.

Proof.

The injectivity of Ψ(k)\Psi_{\infty}^{(k)} and Lemma 4.4 imply that GXλ(x,a|b)(λ𝒮𝒫k){{GX}}_{\lambda}(x,a|b)(\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}) form a formal basis 𝒦(k){\mathcal{K}}_{\infty}^{(k)} over b{\mathcal{R}}_{b}. That is, any element ff of 𝒦(k){\mathcal{K}}_{\infty}^{(k)} can be expressed uniquely as a possibly infinite b{\mathcal{R}}_{b}-linear combination

f=λ𝒮𝒫kcλX(b)GXλ(x,a|b),cλX(b)b.f=\sum_{\lambda\in{\mathcal{S}}{\mathcal{P}}^{k}}c_{\lambda}^{X}(b){{GX}}_{\lambda}(x,a|b),\ \ \ \ c_{\lambda}^{X}(b)\in{\mathcal{R}}_{b}.

On the other hand, Lemma 5.9 shows that Ψ(k)\Psi_{\infty}^{(k)} sends GXλ(x,a|b){{GX}}_{\lambda}(x,a|b) to [ΩλX]T[\Omega_{\lambda}^{X}]_{T}. Therefore we can conclude that the image of Ψ(k)\Psi_{\infty}^{(k)} coincides with 𝕂Tfin(IGk){\mathbb{K}}^{\operatorname{fin}}_{T}({{IG}}^{k}). ∎

Remark 5.11.

If k=0k=0 and λ\lambda is a strict partition, then one result shows that GQλ(x|b){GQ}_{\lambda}(x|b) and GPλ(x|0,b){GP}_{\lambda}(x|0,b) defined by Ikeda–Naruse in [8] coincide with the functions G0Θλ(x|b){}_{0}{G\Theta}_{\lambda}(x|b) and G0Θλ(x|b){}_{0}{G\Theta^{\prime}}_{\lambda}(x|b) respectively, simultaneously establishing their Pfaffian formula.


Acknowledgements. A considerable part of this work developed while the first and third authors were affiliated to KAIST, which they would like to thank for the excellent working conditions. A part of this work was developed while the first author was affiliated to POSTECH, which he would like to thank for the excellent working conditions. He would also like to gratefully acknowledge the support of the National Research Foundation of Korea (NRF) through the grants funded by the Korea government (MSIP) (2014-001824 and 2011-0030044). The second author is supported by Grant-in-Aid for Scientific Research (C) 18K03261, 15K04832. The third author is supported by Grant-in-Aid for Young Scientists (B) 16K17584. The fourth author is supported by Grant-in-Aid for Scientific Research (C) 25400041, (B) 16H03921.

References

  • [1] Anderson, D. K-theoretic Chern class formulas for vexillary degeneracy loci. ArXiv e-prints (Jan. 2017).
  • [2] Buch, A. S., Kresch, A., and Tamvakis, H. A Giambelli formula for isotropic Grassmannians. Selecta Math. (N.S.) 23, 2 (2017), 869–914.
  • [3] Dai, S., and Levine, M. Connective algebraic KK-theory. J. K-Theory 13, 1 (2014), 9–56.
  • [4] Goresky, M., Kottwitz, R., and MacPherson, R. Equivariant cohomology, Koszul duality, and the localization theorem. Invent. Math. 131, 1 (1998), 25–83.
  • [5] Hudson, T., Ikeda, T., Matsumura, T., and Naruse, H. Degeneracy loci classes in KK-theory — determinantal and Pfaffian formula. Adv. Math. 320 (2017), 115–156.
  • [6] Ikeda, T., and Matsumura, T. Pfaffian sum formula for the symplectic Grassmannians. Math. Z. 280, 1-2 (2015), 269–306.
  • [7] Ikeda, T., Mihalcea, L. C., and Naruse, H. Double Schubert polynomials for the classical groups. Adv. Math. 226, 1 (2011), 840–886.
  • [8] Ikeda, T., and Naruse, H. K-theoretic analogues of factorial Schur P- and Q-functions. Adv. Math. 226, 1 (2011), 840–886.
  • [9] Kirillov, A. N., and Naruse, H. Construction of double Grothendieck polynomials of classical types using idCoxeter algebras. Tokyo J. Math. 39, 3 (2017), 695–728.
  • [10] Kostant, B., and Kumar, S. TT-equivariant KK-theory of generalized flag varieties. J. Differential Geom. 32, 2 (1990), 549–603.
  • [11] Krishna, A. Equivariant cobordism for torus actions. Adv. Math. 231, 5 (2012), 2858–2891.
  • [12] Krishna, A. Equivariant cobordism of schemes. Doc. Math. 17 (2012), 95–134.
  • [13] Macdonald, I. G. Symmetric functions and Hall polynomials, second ed. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York, 1995. With contributions by A. Zelevinsky, Oxford Science Publications.
  • [14] Tamvakis, H. Giambelli, Pieri, and tableau formulas via raising operators. J. Reine Angew. Math. 652 (2011), 207–244.
  • [15] Tamvakis, H., and Wilson, E. Double theta polynomials and equivariant Giambelli formulas. Math. Proc. Cambridge Philos. Soc. 160, 2 (2016), 353–377.
  • [16] Wilson, V. Equivariant Giambelli Formulae for Grassmannians. Ph.D. thesis. University of Maryland (2010).

Thomas Hudson, Fachgruppe Mathematik und Informatik, Bergische Universität Wuppertal, 42119 Wuppertal, Germany

email address: hudson@math.uni-wuppertal.de

Takeshi Ikeda, Department of Applied Mathematics, Okayama University of Science, Okayama 700-0005, Japan

email address: ike@xmath.ous.ac.jp

Tomoo Matsumura, Department of Applied Mathematics, Okayama University of Science, Okayama 700-0005, Japan

email address: matsumur@xmath.ous.ac.jp

Hiroshi Naruse, Graduate School of Education, University of Yamanashi, Yamanashi 400-8510, Japan

email address: hnaruse@yamanashi.ac.jp