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

Integral generalized equivariant cohomologies of weighted Grassmann orbifolds

Koushik Brahma Department of Mathematics, Indian Institute of Technology Madras, India koushikbrahma95@gmail.com  and  Soumen Sarkar Department of Mathematics, Indian Institute of Technology Madras, India soumen@iitm.ac.in
Abstract.

We introduce a new definition of weighted Grassmann orbifolds. We study their several invariant qq-cell structures and the orbifold singularities on these qq-cells. We discuss when the integral cohomology of a weighted Grassmann orbifold has no pp-torsion. We compute the equivariant KK-theory ring of weighted Grassmann orbifolds with rational coefficients. We introduce divisive weighted Grassmann orbifolds and show that they have invariant cell structures. We calculate the equivariant cohomology ring, equivariant KK-theory ring and equivariant cobordism ring of a divisive weighted Grassmann orbifold with integer coefficients. We discuss how to compute the weighted structure constants for the integral equivariant cohomology ring of a divisive weighted Grassmann orbifold.

Key words and phrases:
Weighted Grassmann orbifold, qq-cell structure, divisive weighted Grassmann orbifold, equivariant cohomology ring, equivariant KK-theory ring, equivariant cobordism ring, weighted structure constant.
2020 Mathematics Subject Classification:
14M15, 57R18, 55N91, 19L47, 57R85

1. Introduction

We consider the nn-dimensional complex vector space n\mathbb{C}^{n} and a positive integer dd satisfying 1d<n1\leq d<n. Then the set of all dd-dimensional vector subspaces of n\mathbb{C}^{n} is called a (complex) Grassmann manifold and denoted by Gr(d,n)\mbox{Gr}(d,n). In particular, the space Gr(1,n)\mbox{Gr}(1,n) is called the (n1)(n-1)-dimensional complex projective space. The space Gr(d,n)\mbox{Gr}(d,n) has a manifold structure of dimension d(nd){d(n-d)}, see [Muk15, Chapter 1]. This is a projective variety via the Plücker embedding. The natural ()n(\mathbb{C}^{*})^{n}-action on n\mathbb{C}^{n} induces a ()n(\mathbb{C}^{*})^{n}-action on Gr(d,n)\mbox{Gr}(d,n). Grassmann manifolds are central objects of study in algebraic geometry, algebraic topology and differential geometry. Several interesting topological and geometrical properties of Grassmann manifolds can be found in [Lak72, KT03, JP03].

The orbifold version of a complex projective space was introduced in [Kaw73] and was called a twisted projective space. Orbifolds, a generalization of manifolds, were introduced by Satake [Sat56, Sat57] with the name VV-manifolds. Later, Thurston [Thu80] used the terminology orbifolds instead. In the past two decades, several development have been appeared to study orbifolds arising in algebraic geometry, differential geometry and string topology. Some cohomology theories such as the de Rham cohomology [ALR07, Chapter 2], the singular cohomology [Hat02], the Dolbeault cohomology [Bai56], Chen-Ruan cohomology ring [CR04] and orbifold KK-theory [ALR07, Chapter 3] for a class of orbifolds were studied either with rational, real or, complex coefficients. One can construct a CW-complex structure on an effective orbifold following [Gor78]. However, in general, the computation of the singular integral cohomology of an orbifold is considerably difficult.

Let GG be a topological group and XX a GG-space. Then the equivariant map X{pt}X\longrightarrow\{pt\} induces a graded G(pt)\mathcal{E}_{G}^{*}(pt)-algebra structure on G(X)\mathcal{E}_{G}^{*}(X). The readers are referred to [May96] for the definitions and several results on the GG-equivariant generalized cohomology theory G\mathcal{E}_{G}^{*}. If G=HG\mathcal{E}_{G}^{*}=H_{G}^{*}, then it is known as the equivariant cohomology theory defined by

HG(X):=H(EG×GX).H_{G}^{*}(X):=H^{*}(EG\times_{G}X).

The ring HG(X)H_{G}^{*}(X) is called the Borel equivariant cohomology of XX. If G=KG\mathcal{E}_{G}^{*}=K_{G}^{*}, then it is known as the equivariant KK-theory. If XX is compact then KG0(X)K_{G}^{0}(X) is the equivalence classes of GG-equivariant complex vector bundles on XX [Seg68]. If XX is a point with trivial action, then KG(pt)K_{G}^{*}(pt) is isomorphic to R(G)[z,z1]R(G)[z,z^{-1}] where R(G)R(G) is complex representation ring of GG and zz is the Bott element of cohomological dimension 2-2. The GG-equivariant ring MUG(X)MU_{G}^{*}(X) is known as equivariant complex cobordism ring see [tD70]. Sinha [Sin01] and Hanke [Han05] have shown several development on MUGMU_{G}^{*}. However, many interesting questions on MUG(X)MU_{G}^{*}(X) remain undetermined. For example, MUG(pt)MU_{G}^{*}(pt) is not completely known for non-trivial groups GG.

Corti and Reid [CR02] introduced the weighted projective analogs of Grassmann manifolds and called them weighted Grassmannians. Then Abe and Matsumura [AM15] defined them explicitly and studied the equivariant cohomology ring of weighted Grassmannians with rational coefficients. The weighted Grassmannians are projective varieties with orbifold singularities. The simplest weighted Grassmannians are the weighted projective spaces. Kawasaki [Kaw73] proved that the integral cohomology of weighted projective spaces have no torsion and is concentrated in even degrees. The equivariant cohomology ring of a weighted projective space has been studied in [BFR09] in terms of piecewise polynomials. The equivariant KK-theory and equivariant cobordism rings of divisive weighted projective spaces have been discussed in [HHRW16] in terms of piecewise Laurent polynomials and piecewise cobordism forms respectively.

Inspired by the above works, we introduce a different definition of weighted Grassmann orbifolds and study their several topological properties such as torsion in the integral cohomology, equivariant cohomology ring, equivariant KK-theory ring and equivariant cobordism ring with integer coefficients. We note that [CR02, AM15] used the name ‘weighted Grassmannians’. However, keeping other naming in mind like Milnor manifolds and Seifert manifolds, we prefer to use Grassmann manifolds and weighted Grassmann orbifolds.

The paper is organized as follows. In Section 2, analogously to the definition of Grassmann manifold discussed in [Muk15], we introduce another definition of a weighted Grassmann orbifold WGr(d,n)\mbox{WGr}(d,n) for d<nd<n, a ‘weight vector’ W:=(w1,,wn)(0)nW:=(w_{1},\ldots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n} and a1a\in\mathbb{Z}_{\geq 1}. Interestingly, this definition is equivalent to the previous one appeared in [AM15]. We recall the definition of Schubert symbols for d<nd<n and discuss how to get a total ordering on the Schubert symbols. Using this total order we show that there is a ‘weighted Plücker embedding’ from a weighted Grassmann orbifold to a weighted projective space see Lemma 2.5. We describe a qq-cell structure of WGr(d,n)\mbox{WGr}(d,n) in Proposition 2.6. Then we discuss a ()n(\mathbb{C}^{*})^{n}-invariant filtration

{pt}=X0X1X2Xm=WGr(d,n)\{pt\}=X_{0}\subset X_{1}\subset X_{2}\subset\dots\subset X_{m}=\mbox{WGr}(d,n)

of WGr(d,n)\mbox{WGr}(d,n) using the qq-cell decomposition, where m:=(nd)1m:={n\choose d}-1. Here, we consider qq-cell structure in the sense of [PS10, Section 4]. We note that one may get different qq-cell structures depending on the choice of the total orderings on the set of all Schubert symbols for d<nd<n. Accordingly, one may obtain different ()n(\mathbb{C}^{*})^{n}-invariant filtration of WGr(d,n)\mbox{WGr}(d,n).

In Section 3, first we recall that there is an equivariant homeomorphism from 𝕎P(rc0,rc1,,rcm)\mathbb{W}P(rc_{0},rc_{1},\ldots,rc_{m}) to 𝕎P(c0,c1,,cm)\mathbb{W}P(c_{0},c_{1},\ldots,c_{m}) for any 1r1\leq r\in\mathbb{N}. Using this technique, we show how the orbifold singularity on a qq-cell of some subcomplexes of WGr(d,n)\mbox{WGr}(d,n) can be reduced, see Lemma 3.3. Consequently, we get a new qq-cell structure of these subcomplexes including WGr(d,n)\mbox{WGr}(d,n) possibly with less singularity on each qq-cell, see Theorem 3.4. We show in Theorem 3.5 that two weighted Grassmann orbifolds are weakly equivariantly homeomorphic if their weight vectors differ by a permutation σSn\sigma\in S_{n}. We define ‘admissible permutation’ σSn\sigma\in S_{n} for a prime pp and WGr(d,n)\mbox{WGr}(d,n), see Definition 3.8. The following result says when H(WGr(d,n);)H^{*}({\rm WGr}(d,n);\mathbb{Z}) has no pp-torsion.

Theorem A (Theorem 3.10).

If there exists an admissible permutation σSn{\sigma\in S_{n}} for a prime pp and WGr(d,n){\rm WGr}(d,n) then Hodd(WGr(d,n);p)H^{odd}({\rm WGr}(d,n);\mathbb{Z}_{p}) is trivial and H(WGr(d,n);)H^{*}({\rm WGr}(d,n);\mathbb{Z}) has no pp-torsion.

We introduce ‘divisive’ weighted Grassmann orbifolds. We note that this definition coincides with the concept of divisive weighted projective space of [HHRW16] when 1=d<n1=d<n. We prove the following.

Theorem B (Theorem 3.19).

If WGr(d,n){\rm WGr}(d,n) is a divisive weighted Grassmann orbifold then it has a ()n(\mathbb{C}^{*})^{n}-invariant cell structure with only even dimensional cells. Moreover, the ()n(\mathbb{C}^{*})^{n}-action on these cells can be described explicitly.

This result implies that the integral cohomology of a divisive weighted Grassmann orbifold has no torsion and is concentrated in even degrees. We discuss a class of non-trivial examples of divisive weighted Grassmann orbifolds. We remark that the weighted Grassmann orbifold in Example 3.12 is not divisive. However, its integral cohomology has no torsion.

In Section 4, we show that the ()n(\mathbb{C}^{*})^{n}-invariant stratification

{pt}=X0X1Xm=WGr(d,n)\{pt\}=X_{0}\subset X_{1}\subset\cdots\subset X_{m}=\mbox{WGr}(d,n)

has the following property. The quotient Xi/Xi1X_{i}/X_{i-1} is homeomorphic to the Thom space of an orbifold ()n(\mathbb{C}^{*})^{n}-bundle

ξi:(λi)/Gi{pt}\xi^{i}\colon\mathbb{C}^{\ell(\lambda^{i})}/G_{i}\to\{pt\}

for some (λi)(1)\ell(\lambda^{i})\in(\mathbb{Z}_{\geq 1}) and finite groups GiG_{i} for i=1,,mi=1,\ldots,m, see Proposition 4.1. We compute the equivariant KK-theory ring of any weighted Grassmann orbifolds with rational coefficients, see Theorem 4.4. If WGr(d,n)\mbox{WGr}(d,n) is divisive then GiG_{i} is trivial for i=1,,mi=1,\ldots,m. Then considering Tn:=(S1)n()nT^{n}:=(S^{1})^{n}\subset(\mathbb{C}^{*})^{n}, the following result describes the integral equivariant cohomology of certain weighted Grassmann orbifolds.

Theorem C (Theorem 4.7).

Let WGr(d,n){\rm WGr}(d,n) be a divisive weighted Grassmann orbifold for d<nd<n. Then the generalized TnT^{n}-equivariant cohomology with integer coefficient Tn(WGr(d,n);)\mathcal{E}^{\ast}_{T^{n}}({\rm WGr}(d,n);\mathbb{Z}) can be given by

{(fi)i=0mTn(pt;)|eTn(ξij)dividesfifjfor j<iand|λjλi|=d1}\Big{\{}(f_{i})\in\bigoplus_{i=0}^{m}\mathcal{E}^{*}_{T^{n}}(pt;\mathbb{Z})~{}\big{|}~{}e_{T^{n}}(\xi^{ij})~{}\mbox{divides}~{}f_{i}-f_{j}~{}\mbox{for }~{}j<i~{}\mbox{and}~{}|\lambda^{j}\cap\lambda^{i}|=d-1\Big{\}}

for Tn=HTn,KTn\mathcal{E}^{\ast}_{T^{n}}=H^{\ast}_{T^{n}},K^{*}_{T^{n}} and MUTnMU^{*}_{T^{n}}.

The computation of eTn(ξij)e_{T^{n}}(\xi^{ij}) is discussed in (4.4). We compute the equivariant cohomology ring of some weighted Grassmann orbifold with integer coefficients which are not divisive, see Theorem 4.10. For m2m\geq 2, corresponding to each pair of positive integers (n,d)(n,d) such that d<nd<n and m+1=(nd)m+1={n\choose d} we have a TnT^{n}-action on 𝕎P(c0,c1,,cm)\mathbb{W}P(c_{0},c_{1},\dots,c_{m}). For each pair (n,d)(n,d), we discuss the generalized TnT^{n}-equivariant cohomology of a divisive 𝕎P(c0,c1,,cm)\mathbb{W}P(c_{0},c_{1},\dots,c_{m}) with integer coefficients, see Theorem 4.11.

In Section 5, we show that there exist equivariant Schubert classes {wS~λi}i=0m\{w\widetilde{S}_{\lambda^{i}}\}_{i=0}^{m} which form a basis for the integral TnT^{n}-equivariant cohomology of a divisive weighted Grassmann orbifold, see Proposition 5.3. We study some properties of weighted structure constant, see Lemma 5.5. Then we show the following multiplication rule.

Proposition D (Proposition 5.7).

[Weighted Pieri rule]

wS~λ1wS~λj=(wS~λ1|λj)wS~λj+λiλjc0cjwS~λi.w\widetilde{S}_{\lambda^{1}}~{}w\widetilde{S}_{\lambda^{j}}=({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{j}})~{}w\widetilde{S}_{\lambda^{j}}+\sum_{\lambda^{i}\to\lambda^{j}}\dfrac{c_{0}}{c_{j}}~{}w\widetilde{S}_{\lambda^{i}}.

Moreover, we deduce a recurrence relation which helps to compute the weighted structure constants {wcijk}\{wc_{ij}^{k}\} corresponding to this Schubert basis {wS~λi}i=0m\{w\widetilde{S}_{\lambda^{i}}\}_{i=0}^{m} with integral coefficients.

Proposition E ( Proposition 5.8).

For any three Schubert symbols λi,λj\lambda^{i},\lambda^{j} and λk\lambda^{k}, we have the following recurrence relation.

(wS~λ1|λkwS~λ1|λi)wcijk=(λsλic0ciwcsjkλkλtc0ctwcijt).({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{k}}-{w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}})~{}wc_{ij}^{k}=(\sum_{\lambda^{s}\to\lambda^{i}}\dfrac{c_{0}}{c_{i}}~{}wc_{sj}^{k}-\sum_{\lambda^{k}\to\lambda^{t}}\dfrac{c_{0}}{c_{t}}~{}wc_{ij}^{t}).

2. Invariant qq-cell structure on weighted Grassmann orbifolds

In this section, we introduce another definition of weighted Grassmann orbifold WGr(d,n)\mbox{WGr}(d,n) where d<nd<n. We recall the definition of a Schubert symbol for d<n{d<n} and discuss some (total) ordering on the set of Schubert symbols. We show that there is an equivariant embedding from a weighted Grassmann orbifold to a weighted projective space. We show that our definition of weighted Grassmann orbifold is equivalent to the previous one appeared in [AM15]. We study the orbifold and qq-cell structures of weighted Grassmann orbifolds generalizing the manifolds counter part discussed in [MS74].

A Schubert symbol λ\lambda for d<nd<n is a sequence of dd integers (λ1,λ2,,λd)(\lambda_{1},\lambda_{2},\dots,\lambda_{d}) such that 1λ1<λ2<<λdn1\leq\lambda_{1}<\lambda_{2}<\cdots<\lambda_{d}\leq n. The length (λ)\ell(\lambda) of a Schubert symbol λ:=(λ1,λ2,,λd)\lambda:=(\lambda_{1},\lambda_{2},\dots,\lambda_{d}) is defined by (λ):=(λ11)+(λ22)++(λdd)\ell(\lambda):=(\lambda_{1}-1)+(\lambda_{2}-2)+\cdots+(\lambda_{d}-d). There are (nd){n\choose d} many Schubert symbols for d<nd<n. One can define a partial order ‘\preceq’ on the Schubert symbols for d<nd<n by

(2.1) λμ if λiμi for all i=1,2,,d.\lambda\preceq\mu\text{ if }\lambda_{i}\leq\mu_{i}\text{ for all }i=1,2,\dots,d.

Then the set of all Schubert symbols for d<nd<n form a poset with respect to this partial order ‘\preceq’.

Let Md(n,d)M_{d}(n,d) be the set of all complex n×dn\times d matrix of rank dd and GL(d,)\mbox{GL}(d,\mathbb{C}) the set of all non-singular complex matrix of order dd. We denote a matrix AMd(n,d)A\in M_{d}(n,d) as follows

A=(a11a12a1da21a22a2dan1an2and)=(𝐚1𝐚2𝐚n)A=\begin{pmatrix}a_{11}&a_{12}&\cdots&a_{1d}\\ a_{21}&a_{22}&\cdots&a_{2d}\\ \vdots&\vdots&\vdots&\vdots\\ a_{n1}&a_{n2}&\cdots&a_{nd}\end{pmatrix}=\begin{pmatrix}{\bf a}_{1}\\ {\bf a}_{2}\\ \vdots\\ {\bf a}_{n}\end{pmatrix}

where 𝐚id{\bf a}_{i}\in\mathbb{C}^{d} for i=1,,ni=1,\ldots,n.

Definition 2.1.

Let W:=(w1,w2,,wn)(0)nW:=(w_{1},w_{2},\dots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n} and a1a\in\mathbb{Z}_{\geq 1}. We define an equivalence relation w\sim_{w} on Md(n,d)M_{d}(n,d) by

(𝐚𝟏𝐚𝟐𝐚𝐧)w(tw1𝐚𝟏tw2𝐚𝟐twn𝐚𝐧)T\begin{pmatrix}\bf{a}_{1}\\ \bf{a}_{2}\\ \vdots\\ \bf{a}_{n}\end{pmatrix}\sim_{w}\begin{pmatrix}t^{w_{1}}\bf{a}_{1}\\ t^{w_{2}}\bf{a}_{2}\\ \vdots\\ t^{w_{n}}\bf{a}_{n}\end{pmatrix}T

for TGL(d,)T\in{\rm GL}(d,\mathbb{C}) and tt\in\mathbb{C}^{*} such that ta=det(T)t^{a}=\det(T)\in\mathbb{C}^{*}. We denote the identification space by

WGr(d,n):=Md(n,d)w.{\rm WGr}(d,n):=\dfrac{M_{d}(n,d)}{\sim_{w}}.

The quotient map

(2.2) πw:Md(n,d)WGr(d,n)\pi_{w}\colon M_{d}(n,d)\to{\rm WGr}(d,n)

is defined by πw(A)=[A]w\pi_{w}(A)=[A]_{\sim_{w}}. The topology on WGr(d,n){\rm WGr}(d,n) is given by the quotient topology via the map πw\pi_{w}.

Remark 2.2.

If W=(0,0,,0)W=(0,0,\dots,0) and a=1a=1 then WGr(d,n){\rm WGr}(d,n) is the Grassmann manifold Gr(d,n){\rm Gr}(d,n). We denote the corresponding quotient map by

(2.3) π:Md(n,d)Gr(d,n).\pi\colon M_{d}(n,d)\to{\rm Gr}(d,n).

The space Gr(d,n){\rm Gr}(d,n) is a d(nd)d(n-d)-dimensional smooth manifold and represents the set of all dd-dimensional vector subspaces in n\mathbb{C}^{n}. Several basic properties such as manifold and a cell structure of Gr(d,n){\rm Gr}(d,n) can be found in [MS74].

Remark 2.3.

If d=1d=1 then Md(n,d)=M1(n,1)=n{0}M_{d}(n,d)=M_{1}(n,1)=\mathbb{C}^{n}-\{0\} and GL(1,)={\rm GL}(1,\mathbb{C})=\mathbb{C}^{*}. The corresponding w\sim_{w} is given by

(z1,z2,,zn)w(ta+w1z1,ta+w2z2,,ta+wnzn).(z_{1},z_{2},\dots,z_{n})\sim_{w}(t^{a+w_{1}}z_{1},t^{a+w_{2}}z_{2},\dots,t^{a+w_{n}}z_{n}).

The quotient space M1(n,1)w\frac{M_{1}(n,1)}{\sim_{w}} is called the weighted projective space with weights (a+w1,a+w2,,a+wn)(a+w_{1},a+w_{2},\dots,a+w_{n}) and denoted by 𝕎P(c0,c1,,cn1)\mathbb{W}P(c_{0},c_{1},\dots,c_{n-1}) where ci=a+wi+1c_{i}=a+w_{i+1} for i{0,1,,n1}i\in\{0,1,\dots,n-1\}. For the weighted projective space, we denote w\sim_{w} by c\sim_{c} when c=(c0,c1,,cn1)c=(c_{0},c_{1},\dots,c_{n-1}). This identification c\sim_{c} is called a weighted \mathbb{C}^{*}-action on n{0}\mathbb{C}^{n}-\{0\} with weights (c0,c1,,cn1)(c_{0},c_{1},\dots,c_{n-1}). In addition, if W=(0,0,,0)W=(0,0,\dots,0) and a=1a=1 then c0=1=c1==cn1c_{0}=1=c_{1}=\dots=c_{n-1} and 𝕎P(c0,c1,,cn1)\mathbb{W}P(c_{0},c_{1},\dots,c_{n-1}) is Pn1=Gr(1,n)\mathbb{C}P^{n-1}={\rm Gr}(1,n).

Definition 2.4.

Let λ=(λ1,λ2,,λd)\lambda=(\lambda_{1},\lambda_{2},\dots,\lambda_{d}) and μ=(μ1,μ2,,μd)\mu=(\mu_{1},\mu_{2},\dots,\mu_{d}) be two Schubert symbols for d<nd<n. We say that λ<μ\lambda<\mu if (λ)<(μ)\ell(\lambda)<\ell(\mu), otherwise we use the dictionary order if (λ)=(μ)\ell(\lambda)=\ell(\mu).

This gives a total order on the set of all Schubert symbols. Note that the total order ‘<<’ in Definition 2.4 preserves the partial order ‘\preceq’ in (2.1). That is, for two Schubert symbols λ\lambda and μ\mu, λμλμ\lambda\preceq\mu\implies\lambda\leq\mu, but the converse may not be true in general. Observe that there may exist several other total orders on the set of all Schubert symbols which preserve the partial order ‘\preceq’. For example, the dictionary order also gives a total order on the Schubert symbols. By a total order on the set of all Schubert symbols for d<nd<n, we mean one of these total orders on it. For m:=(nd)1m:={n\choose d}-1, let

(2.4) λ0<λ1<λ2<<λm\lambda^{0}<\lambda^{1}<\lambda^{2}<\dots<\lambda^{m}

be a total order on the Schubert symbols for d<nd<n.

For W=(w1,w2,,wn)(0)nW=(w_{1},w_{2},\dots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n}, a1a\in\mathbb{Z}_{\geq 1} and i{0,1,,m}i\in\{0,1,\dots,m\}, let

(2.5) ci:=a+j=1dwλjic_{i}:=a+\sum_{j=1}^{d}w_{\lambda_{j}^{i}}

where λi=(λ1i,λ2i,,λdi)\lambda^{i}=(\lambda_{1}^{i},\lambda_{2}^{i},\dots,\lambda_{d}^{i}) is the ii-th Schubert symbol given in (2.4). Then ci1c_{i}\geq 1 for any i{0,,m}i\in\{0,\ldots,m\}. Therefore, one can define the weighted projective space 𝕎P(c0,c1,,cm)\mathbb{W}P(c_{0},c_{1},\dots,c_{m}) from Remark 2.3. We denote the associated orbit map m+1{0}𝕎P(c0,c1,,cm)\mathbb{C}^{m+1}-\{0\}\to\mathbb{W}P(c_{0},c_{1},\dots,c_{m}) by πc\pi^{\prime}_{c} which can be written as

(2.6) πc(z0,z1,,zm)=[z0:z1::zm]c.\pi^{\prime}_{c}(z_{0},z_{1},\ldots,z_{m})=[z_{0}:z_{1}:\cdots:z_{m}]_{\sim_{c}}.

Note that when c0=c1==cm=1c_{0}=c_{1}=\cdots=c_{m}=1, then the corresponding orbit map is denoted by

π:m+1{0}Pm.\pi^{\prime}\colon\mathbb{C}^{m+1}-\{0\}\to\mathbb{C}P^{m}.

Let (t1,t2,,tn)()n(t_{1},t_{2},\dots,t_{n})\in(\mathbb{C}^{*})^{n} and A=(𝐚1,𝐚2,,𝐚n)trMd(n,d)A=({\bf a}_{1},{\bf a}_{2},\dots,{\bf a}_{n})^{tr}\in M_{d}(n,d). Then ()n(\mathbb{C}^{*})^{n} acts on Md(n,d)M_{d}(n,d) defined by

(2.7) (t1,,tn)(𝐚1,𝐚2,,𝐚n)tr:=(t1𝐚1,t2𝐚2,,tn𝐚n)tr.(t_{1},\ldots,t_{n})({\bf a}_{1},{\bf a}_{2},\dots,{\bf a}_{n})^{tr}:=(t_{1}{\bf a}_{1},t_{2}{\bf a}_{2},\dots,t_{n}{\bf a}_{n})^{tr}.

This induces a natural ()n(\mathbb{C}^{*})^{n}-action on WGr(d,n)\mbox{WGr}(d,n) such that the orbit map πw\pi_{w} of (2.2) is ()n(\mathbb{C}^{*})^{n}-equivariant.

We remark that the standard ordered basis {e1,e2,,en}\{e_{1},e_{2},\dots,e_{n}\} of n\mathbb{C}^{n} induces an ordered basis {eλ0,eλ1,,eλm}\{e_{\lambda^{0}},e_{\lambda^{1}},\dots,e_{\lambda^{m}}\} of Λd(n)\Lambda^{d}(\mathbb{C}^{n}). Therefore, we can identify Λd(n)\Lambda^{d}(\mathbb{C}^{n}) with m+1(={eλ0,eλ1,,eλm})\mathbb{C}^{m+1}(=\mathbb{C}\{e_{\lambda^{0}},e_{\lambda^{1}},\dots,e_{\lambda^{m}}\}). The standard action of ()n(\mathbb{C}^{*})^{n} on n\mathbb{C}^{n} induces an action of ()n(\mathbb{C}^{*})^{n} on m+1{0}\mathbb{C}^{m+1}-\{0\} which is defined by

(2.8) (t1,t2,,tn)(i=0maieλi)=i=0maitλieλi(t_{1},t_{2},\dots,t_{n})(\sum_{i=0}^{m}a_{i}e_{\lambda^{i}})=\sum_{i=0}^{m}a_{i}t_{\lambda^{i}}e_{\lambda^{i}}

where tλ=tλ1tλdt_{\lambda}=t_{\lambda_{1}}\cdots t_{\lambda_{d}} and eλ=eλ1eλde_{\lambda}=e_{\lambda_{1}}\wedge\cdots\wedge e_{\lambda_{d}} for the Schubert symbol λ=(λ1,λ2,,λd)\lambda=(\lambda_{1},\lambda_{2},\dots,\lambda_{d}). This induces a ()n(\mathbb{C}^{*})^{n}-action on the weighted projective space 𝕎P(c0,c1,,cm)\mathbb{W}P(c_{0},c_{1},\ldots,c_{m}) such that the orbit map πc\pi^{{}^{\prime}}_{c} in (2.6) is ()n(\mathbb{C}^{*})^{n}-equivariant.

For each Schubert symbol λ=(λ1,λ2,,λd)\lambda=(\lambda_{1},\lambda_{2},\dots,\lambda_{d}), let AλA_{\lambda} be the matrix with row vectors 𝐚λ1,𝐚λ2,,𝐚λd{\bf a}_{\lambda_{1}},{\bf a}_{\lambda_{2}},\dots,{\bf a}_{\lambda_{d}}. Consider the map P:Md(n,d)m+1P\colon M_{d}(n,d)\to\mathbb{C}^{m+1} defined by

(2.9) P(A)=𝐯1𝐯2𝐯d=i=0mdet(Aλi)eλi,P(A)={\bf v}_{1}\wedge{\bf v}_{2}\wedge\dots\wedge{\bf v}_{d}=\sum_{i=0}^{m}\det(A_{\lambda^{i}})e_{\lambda^{i}},

where 𝐯1,𝐯2,,𝐯dn{\bf v}_{1},{\bf v}_{2},\ldots,{\bf v}_{d}\in\mathbb{C}^{n} are the columns of AA. Observe that P(A)0P(A)\neq 0 because AMd(n,d)A\in M_{d}(n,d) has rank dd.

Lemma 2.5.

The map in (2.9) induces a weighted Plücker embedding

Plw:WGr(d,n)𝕎P(c0,c1,c2,,cm).Pl_{w}\colon{\rm WGr}(d,n)\rightarrow\mathbb{W}P(c_{0},c_{1},c_{2},\dots,c_{m}).
Proof.

From (2.9) we have

P(DAT)=i=0mdet((DAT)λi)eλi=i=0mtcidet(Aλi)eλi,P(DAT)=\sum_{i=0}^{m}\det((DAT)_{\lambda^{i}})e_{\lambda^{i}}=\sum_{i=0}^{m}t^{c_{i}}\det(A_{\lambda^{i}})e_{\lambda^{i}},

where TGL(d,)T\in\mbox{GL}(d,\mathbb{C}), D=diag(tw1,tw2,,twn)D=\mbox{diag}(t^{w_{1}},t^{w_{2}},\ldots,t^{w_{n}}) is the diagonal matrix for tt\in\mathbb{C}^{*} such that ta=det(T)t^{a}=\det(T) and cic_{i} is defined in (2.5) for i=0,1,2,,mi=0,1,2,\dots,m. Therefore, this induces a map

Plw:WGr(d,n)𝕎P(c0,c1,,cm)Pl_{w}\colon\mbox{WGr}(d,n)\rightarrow\mathbb{W}P(c_{0},c_{1},\ldots,c_{m})

defined by

(2.10) Plw([A]w)=[det(Aλ0):det(Aλ1)::det(Aλm)]c.Pl_{w}([A]_{\sim_{w}})=[\det(A_{\lambda^{0}}):\det(A_{\lambda^{1}}):\dots:\det(A_{\lambda^{m}})]_{\sim_{c}}.

This map satisfies the following commutative diagram

Md(n,d){M_{d}(n,d)}m+1{0}{\mathbb{C}^{m+1}-\{0\}}WGr(d,n){\mbox{WGr}(d,n)}𝕎P(c0,c1,,cm).{\mathbb{W}P(c_{0},c_{1},\dots,c_{m}).}P\scriptstyle{P}πw\scriptstyle{\pi_{w}}πc\scriptstyle{\pi_{c}^{{}^{\prime}}}Plw\scriptstyle{Pl_{w}}

Thus the map PlwPl_{w} is continuous, since πw\pi_{w} and πc\pi^{{}^{\prime}}_{c} are quotient maps.

Let [A]wWGr(d,n)[A]_{\sim_{w}}\in\mbox{WGr}(d,n) for some AMd(n,d)A\in M_{d}(n,d). So, there exists a Schubert symbol λi\lambda^{i} such that det(Aλi)0\det(A_{\lambda^{i}})\neq 0. Without loss of generality, we can assume that Aλi=IdA_{\lambda^{i}}=\mbox{I}_{d}, where Id\mbox{I}_{d} is the identity matrix of order dd. If AλiIdA_{\lambda^{i}}\neq\text{I}_{d} then one can calculate ss\in\mathbb{C}^{*} such that sci=1/det(Aλi)s^{c_{i}}=1/\det(A_{\lambda^{i}}). Now we consider the matrices D=diag(sw1,sw2,,swn)D=\mbox{diag}(s^{w_{1}},s^{w_{2}},\dots,s^{w_{n}}) and T=(DλiAλi)1T=(D_{\lambda^{i}}A_{\lambda^{i}})^{-1}. Then det(T)=sa\det(T)=s^{a} and (DAT)λi=Id(DAT)_{\lambda^{i}}=\text{I}_{d}. Note [DAT]w=[A]wWGr(d,n)[DAT]_{\sim_{w}}=[A]_{\sim_{w}}\in\mbox{WGr}(d,n).

We prove that PlwPl_{w} is injective. Let [A]w[A]_{\sim_{w}} and [B]wWGr(d,n)[B]_{\sim_{w}}\in\mbox{WGr}(d,n) such that Plw([A]w)=Plw([B]w)Pl_{w}([A]_{\sim_{w}})=Pl_{w}([B]_{\sim_{w}}) for some A,BMd(n,d)A,B\in M_{d}(n,d). Now

(2.11) Plw([A]w)=Plw([B]w)det(Aλj)=tcjdet(Bλj)Pl_{w}([A]_{\sim_{w}})=Pl_{w}([B]_{\sim_{w}})\implies\det(A_{\lambda^{j}})=t^{c_{j}}\det(B_{\lambda^{j}})

for some tt\in\mathbb{C}^{*} and for all j{0,1,,m}j\in\{0,1,\dots,m\}. Since AMd(n,d)A\in M_{d}(n,d) there exists a Schubert symbol λi=(λ1i,,λdi)\lambda^{i}=(\lambda_{1}^{i},\dots,\lambda_{d}^{i}) such that det(Aλi)0\det(A_{\lambda^{i}})\neq 0. Then using (2.11), det(Bλi)0\det(B_{\lambda^{i}})\neq 0. So we can assume Aλi=Bλi=IdA_{\lambda^{i}}=B_{\lambda^{i}}=\text{I}_{d}. Then tci=1t^{c_{i}}=1. Consider the matrices D=diag(tw1,tw2,,twn)D=\mbox{diag}(t^{w_{1}},t^{w_{2}},\dots,t^{w_{n}}) and T=diag(twλ1i,,twλdi)T=\mbox{diag}(t^{-w_{\lambda^{i}_{1}}},\dots,t^{-w_{\lambda^{i}_{d}}}). Thus, we have Bλi=(DAT)λi.B_{\lambda^{i}}=(DAT)_{\lambda^{i}}.

Let akla_{kl} and bklb_{kl} be the (k,l)(k,l)-th entries of the matrices AA and BB respectively for k(λ1i,,λdi)k\notin(\lambda_{1}^{i},\dots,\lambda_{d}^{i}) and 1ld1\leq l\leq d. For a fixed ll, let λj\lambda^{j} be the Schubert symbol obtained by replacing λli\lambda_{l}^{i} by kk in λi\lambda^{i} and then ordering the later set. Then det(Aλj)=akl\det(A_{\lambda^{j}})=a_{kl} and det(Bλj)=bkl\det(B_{\lambda^{j}})=b_{kl}. Thus using (2.11), we get

bkl=tcjaklbkl=tcjciaklbkl=twkwλliakl.b_{kl}=t^{c_{j}}a_{kl}\implies b_{kl}=t^{c_{j}-c_{i}}a_{kl}\implies b_{kl}={t^{w_{k}-w_{\lambda_{l}^{i}}}}a_{kl}.

The above condition holds for all 1kn1\leq k\leq n and 1ld1\leq l\leq d. This gives B=DATB=DAT. Then we have [A]w=[B]w[A]_{\sim_{w}}=[B]_{\sim_{w}}. Hence, PlwPl_{w} is an injective map.

Observe that, if W=(0,0,,0)W=(0,0,\dots,0) and a=1a=1 then the map PlwPl_{w} is the usual Plücker map

Pl:Gr(d,n)Pm.Pl\colon\mbox{Gr}(d,n)\rightarrow\mathbb{C}P^{m}.

It is well known that PlPl is an embedding. Moreover, we have the following commutative diagrams.

(2.12) WGr(d,n){\mbox{WGr}(d,n)}𝕎P(c0,c1,,cm){\mathbb{W}P(c_{0},c_{1},\dots,c_{m})}Md(n,d){M_{d}(n,d)}m+1{0}{\mathbb{C}^{m+1}-\{0\}}Gr(d,n){\mbox{Gr}(d,n)}Pm.{\mathbb{C}P^{m}.}Plw\scriptstyle{Pl_{w}}P\scriptstyle{P}πw\scriptstyle{\pi_{w}}π\scriptstyle{\pi}πc\scriptstyle{\pi_{c}^{{}^{\prime}}}π\scriptstyle{\pi^{{}^{\prime}}}Pl\scriptstyle{Pl}

Let UU be an open subset of WGr(d,n)\mbox{WGr}(d,n). Then πw1(U)\pi_{w}^{-1}(U) is an open subset of Md(n,d)M_{d}(n,d). Since the map π\pi in (2.3) is an orbit map so π(πw1(U))\pi(\pi_{w}^{-1}(U)) is an open subset of Gr(d,n)\mbox{Gr}(d,n). Thus Pl(π(πw1(U)))=π(P(πw1(U)))Pl(\pi(\pi_{w}^{-1}(U)))=\pi^{\prime}(P(\pi_{w}^{-1}(U))) is an open subset of Pl(Gr(d,n))Pl(\mbox{Gr}(d,n)). Then P(πw1(U))P(\pi_{w}^{-1}(U)) is an open subset of P(Md(n,d))P(M_{d}(n,d)). Therefore, Plw(U)=πc(P(πw1(U)))Pl_{w}(U)=\pi_{c}^{\prime}(P(\pi_{w}^{-1}(U))) is an open subset of Plw(WGr(d,n))Pl_{w}(\mbox{WGr}(d,n)). Thus PlwPl_{w} is an embedding. ∎

Note that the actions of ()n(\mathbb{C}^{*})^{n} on WGr(d,n)\mbox{WGr}(d,n) and 𝕎P(c0,c1,,cm)\mathbb{W}P(c_{0},c_{1},\ldots,c_{m}) implies that the weighted Plücker embedding PlwPl_{w} in (2.10) is ()n(\mathbb{C}^{*})^{n}-equivariant, and Plw(WGr(d,n))Pl_{w}(\mbox{WGr}(d,n)) is a ()n(\mathbb{C}^{*})^{n}-invariant subset of 𝕎P(c0,c1,,cm)\mathbb{W}P(c_{0},c_{1},\ldots,c_{m}). Thus all the maps in the diagram (2.12) are ()n(\mathbb{C}^{*})^{n}-equivariant.

Now we show that Definition 2.1 is equivalent to the definition of a weighted Grassmannian studied in [AM15]. The algebraic torus ()n+1(\mathbb{C}^{*})^{n+1} acts on Λd(n)\Lambda^{d}(\mathbb{C}^{n}) by

(t1,t2,,tn,t)i=0maλieλi=i=0mttλiaλieλi(t_{1},t_{2},\dots,t_{n},t)\sum_{i=0}^{m}a_{\lambda^{i}}e_{\lambda^{i}}=\sum_{i=0}^{m}t\cdot t_{\lambda^{i}}a_{\lambda^{i}}e_{\lambda^{i}}

where tλ=tλ1tλdt_{\lambda}=t_{\lambda_{1}}\cdots t_{\lambda_{d}} for λ=(λ1,,λd)\lambda=(\lambda_{1},\ldots,\lambda_{d}). Consider the subgroup WDWD of ()n+1(\mathbb{C}^{*})^{n+1} defined by

WD:={(tw1,tw2,,twn,ta)|t}.WD:=\{(t^{w_{1}},t^{w_{2}},\dots,t^{w_{n}},t^{a})~{}|~{}t\in\mathbb{C}^{*}\}.

Then the restricted action of WDWD on Λd(n){0}\Lambda^{d}(\mathbb{C}^{n})-\{0\} is given by

(tw1,tw2,,twn,ta)i=0maλieλi=i=0mtciaλieλi.(t^{w_{1}},t^{w_{2}},\dots,t^{w_{n}},t^{a})\sum_{i=0}^{m}a_{\lambda^{i}}e_{\lambda^{i}}=\sum_{i=0}^{m}t^{c_{i}}a_{\lambda^{i}}e_{\lambda^{i}}.

Observe that this action of WDWD is same as the weighted \mathbb{C}^{*}-action in Remark 2.3. Then we have Λd(n){0}WD=𝕎P(c0,,cm)\dfrac{\Lambda^{d}(\mathbb{C}^{n})-\{0\}}{WD}=\mathbb{W}P(c_{0},\dots,c_{m}) and by the commutativity of the diagram (2.12) we have

Plw(WGr(d,n))=P(Md(n,d))WD.Pl_{w}(\mbox{WGr}(d,n))=\dfrac{P(M_{d}(n,d))}{WD}.

Therefore the topologies on WGr(d,n)\mbox{WGr}(d,n) and P(Md(n,d))WD\dfrac{P(M_{d}(n,d))}{WD} are equivalent. Abe and Matsumura [AM15] called the quotient P(Md(n,d))WD\dfrac{P(M_{d}(n,d))}{WD} a weighted Grassmannian and showed that it has an orbifold structure. We call WGr(d,n)\mbox{WGr}(d,n) a weighted Grassmann orbifold associated to the pair (W,a)(W,a).

Next, we recall the Schubert cell decomposition of Gr(d,n)\mbox{Gr}(d,n) following [MS74]. For knk\leq n, we identify

k={(z1,z2,,zk,0,,0)n}.\mathbb{C}^{k}=\{(z_{1},z_{2},...,z_{k},0,...,0)\in\mathbb{C}^{n}\}.

For the Schubert symbol λ=(λ1,λ2,,λd)\lambda=(\lambda_{1},\lambda_{2},\dots,\lambda_{d}), the Schubert cell E(λ)E(\lambda) is defined by

E(λ):={XGr(d,n)|dim(Xλj)=j,dim(Xλj1)=j1;j[d]},\displaystyle E(\lambda):=\{X\in\mbox{Gr}(d,n)~{}|~{}\text{dim}(X\cap\mathbb{C}^{\lambda_{j}})=j,~{}\text{dim}(X\cap\mathbb{C}^{\lambda_{j}-1})=j-1;\forall~{}j\in[d]\},

where [d]:={1,2,,d}[d]:=\{1,2,\dots,d\}. We have the following homeomorphism from [MS74, Chapter-6].

(2.13) E(λ){[100000100000001000000]| and ej is the λj-th row for j[d]}.E(\lambda)\cong\Big{\{}\begin{bmatrix}*&*&\dots&*\\ \vdots&\vdots&&\vdots\\ *&*&\dots&*\\ 1&0&\dots&0\\ 0&*&\dots&*\\ \vdots&\vdots&&\vdots\\ 0&*&\dots&*\\ 0&1&\dots&0\\ 0&0&\dots&*\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&*\\ 0&0&\dots&1\\ 0&0&\dots&0\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&0\\ \end{bmatrix}\big{|}~{}*\in\mathbb{C}\text{ and }e_{j}\text{ is the }\lambda_{j}\text{-th row for }j\in[d]\Big{\}}.

Note that jj-th column in the matrices in (2.13) has λj\lambda_{j}-th entry 11 and all subsequent entries of this column are zero for j[d]j\in[d]. Then E(λ)E(\lambda) is an open cell of dimension (λ)=(λ11)+(λ22)++(λdd)\ell(\lambda)=(\lambda_{1}-1)+(\lambda_{2}-2)+\dots+(\lambda_{d}-d).

Proposition 2.6.

There is a qq-cell structure on WGr(d,n){\rm WGr}(d,n) for 0<d<n0<d<n.

Proof.

For each i{0,1,,m}i\in\{0,1,\dots,m\}, we define E~(λi):=π1(E(λi))\widetilde{E}(\lambda^{i}):=\pi^{-1}(E(\lambda^{i})) where the map π\pi is defined in (2.3). The Schubert cell decomposition of Gr(d,n)\mbox{Gr}(d,n) gives that Gr(d,n)=i=0mE(λi)\mbox{Gr}(d,n)=\sqcup_{i=0}^{m}E(\lambda^{i}). This implies

(2.14) Md(n,d)=i=0mE~(λi),M_{d}(n,d)=\sqcup_{i=0}^{m}\widetilde{E}(\lambda^{i}),

since the map π\pi is surjective. Note that

E~(λi)={AMd(n,d)|det(Aλi)0,det(Aλj)=0,for j>i}.\widetilde{E}(\lambda^{i})=\{A\in M_{d}(n,d)~{}|~{}\det(A_{\lambda^{i}})\neq 0,\det(A_{\lambda^{j}})=0,\text{for }j>i\}.

Let AE~(λi)A\in\widetilde{E}(\lambda^{i}) and AwBA\sim_{w}B for a matrix BMd(n,d)B\in M_{d}(n,d). Then BE~(λi)B\in\widetilde{E}(\lambda^{i}). Therefore we have the following decomposition of WGr(d,n)\mbox{WGr}(d,n).

WGr(d,n)=πw(E~(λ0))πw(E~(λ1))πw(E~(λi)).\mbox{WGr}(d,n)=\pi_{w}(\widetilde{E}(\lambda^{0}))\sqcup\pi_{w}(\widetilde{E}(\lambda^{1}))\sqcup\dots\sqcup\pi_{w}(\widetilde{E}(\lambda^{i})).

By the commutativity of the diagram (2.12), we get

Plw(πw(E~(λi)))=πc(P(E~(λi))) and P(E~(λi))=(π)1(Pl(E(λi))).Pl_{w}(\pi_{w}(\widetilde{E}(\lambda^{i})))=\pi^{\prime}_{c}(P(\widetilde{E}({\lambda^{i}})))\text{ and }P(\widetilde{E}(\lambda^{i}))=(\pi^{{}^{\prime}})^{-1}(Pl(E(\lambda^{i}))).

The map π\pi^{\prime} is a principal \mathbb{C}^{*}-bundle, and E(λi)E(\lambda^{i}) is contractible. So, there is a bundle isomorphism

ϕi:P(E~(λi))E(λi)×.\phi_{i}\colon P(\widetilde{E}(\lambda^{i}))\to E(\lambda^{i})\times\mathbb{C}^{*}.

This map can be defined by ϕi(P(A))=(π(A),det(Aλi))\phi_{i}(P(A))=(\pi(A),\det(A_{\lambda^{i}})). The inverse map is defined by (π(A),s)(s(det(Aλi))1P(A))(\pi(A),s)\mapsto(s(\det(A_{\lambda^{i}}))^{-1}P(A)).

Let π(A)Gr(d,n)\pi(A)\in\mbox{Gr}(d,n) for some A=(𝐚1,𝐚2,,𝐚n)trMd(n,d)A=({\bf a}_{1},{\bf a}_{2},\dots,{\bf a}_{n})^{tr}\in M_{d}(n,d) and tt\in\mathbb{C}^{*}. There is an action of \mathbb{C}^{*} on Gr(d,n)\mbox{Gr}(d,n) defined by

(2.15) t.π(A)=t.π((𝐚1,𝐚2,,𝐚n)tr):=π((tw1𝐚1,tw2𝐚2,,twn𝐚n)tr).t.\pi(A)=t.\pi(({\bf a}_{1},{\bf a}_{2},\dots,{\bf a}_{n})^{tr}):=\pi((t^{w_{1}}{\bf a}_{1},t^{w_{2}}{\bf a}_{2},\dots,t^{w_{n}}{\bf a}_{n})^{tr}).

If π(A)=π(B)\pi(A)=\pi(B), then A=BTDA=DBTA=BT\iff DA=DBT for a diagonal matrix DD and TGL(d,)T\in\text{GL}(d,\mathbb{C}). Thus t.π(A)=t.π(B)t.\pi(A)=t.\pi(B). Then ϕi\phi_{i} becomes \mathbb{C}^{*}-equivariant with the following weighted \mathbb{C}^{*}-action on E(λi)×E(\lambda^{i})\times\mathbb{C}^{*} given by

t.(π(A),s)=(t.π(A),tcis),t.(\pi(A),s)=(t.\pi(A),t^{c_{i}}s),

where t.π(A)t.\pi(A) is defined in (2.15) and cic_{i} is defined in (2.5). Let G(ci)G(c_{i}) be the group of cic_{i}-th roots of unity defined by

G(ci):={t|tci=1},G(c_{i}):=\{t\in\mathbb{C}^{*}~{}|~{}t^{c_{i}}=1\},

for i=0,1,,mi=0,1,\dots,m. Then the finite group G(ci)G(c_{i}) acts on the second factor of E(λi)×E(\lambda^{i})\times\mathbb{C}^{*} trivially. Thus

πc(P(E~(λi)))=P(E~(λi)) weighted -action E(λi)× weighted -action E(λi)G(ci).\pi^{\prime}_{c}(P(\widetilde{E}({\lambda^{i}})))=\frac{P(\widetilde{E}(\lambda^{i}))}{\mbox{ weighted }\mathbb{C}^{*}\mbox{-action }}\cong\frac{E(\lambda^{i})\times\mathbb{C}^{*}}{\mbox{ weighted }\mathbb{C}^{*}\mbox{-action }}\cong\frac{E(\lambda^{i})}{G(c_{i})}.

Therefore we get a qq-cell decomposition of WGr(d,n)\mbox{WGr}(d,n) given by

Plw(WGr(d,n))=E(λ0)G(c0)E(λ1)G(c1)E(λ2)G(c2)E(λm)G(cm).Pl_{w}(\mbox{WGr}(d,n))=\frac{E(\lambda^{0})}{G(c_{0})}\sqcup\frac{E(\lambda^{1})}{G(c_{1})}\sqcup\frac{E(\lambda^{2})}{G(c_{2})}\sqcup\dots\sqcup\frac{E(\lambda^{m})}{G(c_{m})}.

For each k{0,1,2,,m}k\in\{0,1,2,\dots,m\}, let Xk:=i=0kE(λi)G(ci)WGr(d,n)X_{k}:=\sqcup_{i=0}^{k}\frac{E(\lambda^{i})}{G(c_{i})}\subset\mbox{WGr}(d,n). Here XkX_{k} is built inductively by attaching the qq-cells E(λ0)G(c0),,E(λk)G(ck)\frac{E(\lambda^{0})}{G(c_{0})},\ldots,\frac{E(\lambda^{k})}{G(c_{k})} so that XkX_{k} remains a subset of WGr(d,n)\mbox{WGr}(d,n). Then we have the following filtration of qq-CW complexes which are invariant under ()n(\mathbb{C}^{*})^{n}-action on WGr(d,n)\mbox{WGr}(d,n),

(2.16) {pt}=X0X1X2Xm=WGr(d,n).\{pt\}=X_{0}\subset X_{1}\subset X_{2}\subset\dots\subset X_{m}=\mbox{WGr}(d,n).

We note that the paper [AM15] discussed a qq-cell structure of WGr(d,n)\mbox{WGr}(d,n). However, our approach is different and helps to study torsions in the integral cohomology of WGr(d,n)\mbox{WGr}(d,n).

Remark 2.7.

For each k{0,1,2,,m}k\in\{0,1,2,\dots,m\}, consider X~kMd(n,d)\widetilde{X}_{k}\subset M_{d}(n,d) defined by

X~k:={AMd(n,d)|det(Aλj)=0, for j>k}.\widetilde{X}_{k}:=\{A\in M_{d}(n,d)~{}|~{}\text{det}(A_{\lambda^{j}})=0,\text{ for }j>k\}.

Then X~k=i=0kE~(λi)Md(n,d)\widetilde{X}_{k}=\sqcup_{i=0}^{k}\widetilde{E}({\lambda^{i}})\subset M_{d}(n,d) and we have Xk=X~kw=πw(X~k)X_{k}=\frac{\widetilde{X}_{k}}{\sim_{w}}=\pi_{w}(\widetilde{X}_{k}).

3. Integral cohomology of certain weighted Grassmann orbifolds

In this section, we study several qq-cell structure on a weighted Grassmann orbifold. We show how a permutation on the weight vector affects the weighted Grassmann orbifold. We define admissible permutation σSn\sigma\in S_{n} for a prime pp and WGr(d,n)\mbox{WGr}(d,n). Then we discuss when H(WGr(d,n);)H^{*}({\rm WGr}(d,n);\mathbb{Z}) has no pp-torsion. We introduce the concept of divisive weighted Grassmann orbifolds which incorporates the divisive weighted projective spaces of [HHRW16]. We show that a divisive weighted Grassmann orbifold has a ()n(\mathbb{C}^{*})^{n}-invariant cell structure. We describe this action on each cell explicitly. As a consequence, we get that the integral cohomology of a divisive weighted Grassmann orbifold has no torsion and is concentrated in even degrees.

The following lemma is well-known, but for our purpose we may need its proof.

Lemma 3.1.

The map πc:m+1{0}𝕎P(c0,c1,,cm)\pi_{c}^{\prime}\colon\mathbb{C}^{m+1}-\{0\}\to\mathbb{W}P(c_{0},c_{1},\dots,c_{m}) induces an equivariant homeomorphism 𝕎P(rc0,rc1,,rcm)𝕎P(c0,c1,,cm)\mathbb{W}P(rc_{0},rc_{1},\dots,rc_{m})\to\mathbb{W}P(c_{0},c_{1},\dots,c_{m}) for any positive integer rr.

Proof.

The weighted \mathbb{C}^{*}-action on m+1{0}\mathbb{C}^{m+1}-\{0\} for 𝕎P(rc0,rc1,,rcm)\mathbb{W}P(rc_{0},rc_{1},\dots,rc_{m}) is given by

t(z0,z1,,zm)=(trc0z0,trc1z1,,trcmzm).t(z_{0},z_{1},\dots,z_{m})=(t^{rc_{0}}z_{0},t^{rc_{1}}z_{1},\dots,t^{rc_{m}}z_{m}).

We denote the equivalence class by [z0:z1::zm]rc[z_{0}:z_{1}:\dots:z_{m}]_{\sim_{rc}}.

One can define a map f:𝕎P(rc0,rc1,,rcm)𝕎P(c0,,cm)f\colon\mathbb{W}P(rc_{0},rc_{1},\dots,rc_{m})\to\mathbb{W}P(c_{0},\dots,c_{m}) by

f([z0:z1::zm]rc)=[z0:z1::zm]cf([z_{0}:z_{1}:\dots:z_{m}]_{\sim_{rc}})=[z_{0}:z_{1}:\dots:z_{m}]_{\sim_{c}}

and a map g:𝕎P(c0,c1,,cm)𝕎P(rc0,rc1,,rcm)g\colon\mathbb{W}P(c_{0},c_{1},\dots,c_{m})\to\mathbb{W}P(rc_{0},rc_{1},\dots,rc_{m}) by

g([z0:z1::zm]c)=[z0:z1::zm]rc.g([z_{0}:z_{1}:\dots:z_{m}]_{\sim_{c}})=[z_{0}:z_{1}:\dots:z_{m}]_{\sim_{rc}}.

Thus the following diagram commutes

m+1{0}\textstyle{\mathbb{C}^{m+1}-\{0\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Id\scriptstyle{\rm{Id}}πrc\scriptstyle{\pi^{\prime}_{rc}}m+1{0}\textstyle{\mathbb{C}^{m+1}-\{0\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πc\scriptstyle{\pi^{\prime}_{c}}𝕎P(rc0,,rcm)\textstyle{\mathbb{W}P(rc_{0},\dots,rc_{m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}𝕎P(c0,,cm).\textstyle{\mathbb{W}P(c_{0},\dots,c_{m}).\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}

Observe that, we have fg=Id𝕎P(c0,,cm)f\circ g=\text{Id}_{\mathbb{W}P(c_{0},\dots,c_{m})} and gf=Id𝕎P(rc0,,rcm)g\circ f=\text{Id}_{\mathbb{W}P(rc_{0},\dots,rc_{m})}. Thus ff is a bijective map with the inverse map gg.

Let UU be an open subset of 𝕎P(c0,,cm)\mathbb{W}P(c_{0},\dots,c_{m}) Then (πc)1(U)=(πrc)1f1(U)(\pi_{c}^{{}^{\prime}})^{-1}(U)=(\pi_{rc}^{{}^{\prime}})^{-1}\circ f^{-1}(U). Since πc\pi_{c}^{{}^{\prime}} is a quotient map then (πc)1(U)(\pi_{c}^{{}^{\prime}})^{-1}(U) is an open subset of m+1{0}\mathbb{C}^{m+1}-\{0\}. Thus f1(U)f^{-1}(U) is an open subset of 𝕎P(rc0,,rcm)\mathbb{W}P(rc_{0},\dots,rc_{m}) as πrc\pi_{rc}^{{}^{\prime}} is a quotient map. Thus ff is continuous. By the similar arguments, we can show that gg is continuous. Hence ff is a homeomorphism and also it is equivariant with respect to the ()n(\mathbb{C}^{*})^{n}-action on 𝕎P(c0,,cm)\mathbb{W}P(c_{0},\dots,c_{m}) and 𝕎P(rc0,,rcm)\mathbb{W}P(rc_{0},\dots,rc_{m}) defined after (2.8). ∎

Lemma 3.2.

Let BB be a subset of m+1{0}\mathbb{C}^{m+1}-\{0\}. Let Bc:=πc(B)B^{{}^{\prime}}_{c}:=\pi_{c}^{{}^{\prime}}(B) and Brc:=πrc(B)B^{{}^{\prime}}_{rc}:=\pi_{rc}^{{}^{\prime}}(B). Then f|Brc:BrcBcf|_{B^{{}^{\prime}}_{rc}}\colon B^{{}^{\prime}}_{rc}\to B_{c}^{{}^{\prime}} is a homeomorphism.

Proof.

Consider the following commutative diagram

B{B}B{B}Brc{B^{{}^{\prime}}_{rc}}Bc.{B^{{}^{\prime}}_{c}.}Id\scriptstyle{\rm{Id}}πrc\scriptstyle{\pi_{rc}^{{}^{\prime}}}πc\scriptstyle{\pi_{c}^{{}^{\prime}}}f|Brc\scriptstyle{f|_{B^{{}^{\prime}}_{rc}}}

The map ff is well defined and one-one. It follows that f|Brcf|_{B^{{}^{\prime}}_{rc}} is also well defined and one-one. Note that f|Brcf|_{B^{{}^{\prime}}_{rc}} is defined by f|Brc(πrc(b))=πc(b)f|_{B^{{}^{\prime}}_{rc}}(\pi_{rc}^{{}^{\prime}}(b))=\pi_{c}^{{}^{\prime}}(b). Therefore, πrc(b)Brc\pi_{rc}^{{}^{\prime}}(b)\in B^{{}^{\prime}}_{rc} is the inverse image of an element πc(b)Bc\pi_{c}^{{}^{\prime}}(b)\in B_{c}^{{}^{\prime}}. So f|Brcf|_{B^{{}^{\prime}}_{rc}} is bijective. Also (f|Brc)1=g|Bc.(f|_{B^{{}^{\prime}}_{rc}})^{-1}=g|_{B_{c}^{{}^{\prime}}}. To conclude, f|Bcf|_{B^{{}^{\prime}}_{c}} is a homeomorphism, recall that the restriction of a continuous map is also continuous. ∎

We apply the previous result onto some subsets of P(Md(n,d))m+1{0}P(M_{d}(n,d))\subseteq\mathbb{C}^{m+1}-\{0\} for m+1=(nd)m+1={n\choose d}. For all k{0,1,,m}k\in\{0,1,\dots,m\}, recall the space X~k\widetilde{X}_{k} from Remark 2.7. Then

P(X~k)=i=0kP(E~(λi))P(Md(n,d))P(\widetilde{X}_{k})=\sqcup_{i=0}^{k}P(\widetilde{E}(\lambda^{i}))\subseteq P(M_{d}(n,d))

using (2.14). Also P(X~k)k+1{0}m+1{0}P(\widetilde{X}_{k})\subseteq\mathbb{C}^{k+1}-\{0\}\subseteq\mathbb{C}^{m+1}-\{0\}, for k{0,1,,m}k\in\{0,1,\ldots,m\}.

One can calculate cic_{i} for all i{0,1,,m}i\in\{0,1,\dots,m\} from (2.5) for an weighted Grassmann orbifold WGr(d,n)\mbox{WGr}(d,n). Let rk:=gcd{c0,c1,,ck}r_{k}:=\gcd\{c_{0},c_{1},\dots,c_{k}\} for all k{1,2,m}k\in\{1,2,\dots m\} and G(rk)G(r_{k}) be the group of rkr_{k}-th roots of unity. Then G(rk)G(r_{k}) is a subgroup of G(ci)G(c_{i}) and G(ci)/G(rk)G(c_{i})/G(r_{k}) is isomorphic to G(ci/rk)G(c_{i}/r_{k}) for i{0,1,2,,k}i\in\{0,1,2,\dots,k\}.

Lemma 3.3.

The space πc(P(X~k))\pi_{c}^{{}^{\prime}}(P(\widetilde{X}_{k})) is homeomorphic to πcrk(P(X~k))\pi_{\frac{c}{r_{k}}}^{{}^{\prime}}(P(\widetilde{X}_{k})). Moreover, E(λk)/G(ck)E(\lambda^{k})/G(c_{k}) is homeomorphic to E(λk)/G(ck/rk)E(\lambda^{k})/G(c_{k}/r_{k}).

Proof.

The following diagram is commutative.

P(X~k){P(\widetilde{X}_{k})}P(X~k){P(\widetilde{X}_{k})}πc(P(X~k)){\pi_{c}^{{}^{\prime}}(P(\widetilde{X}_{k}))}πcrk(P(X~k)).{\pi_{\frac{c}{r_{k}}}^{{}^{\prime}}(P(\widetilde{X}_{k})).}Id\scriptstyle{\rm{Id}}πc\scriptstyle{\pi_{c}^{{}^{\prime}}}πcrk\scriptstyle{\pi_{\frac{c}{r_{k}}}^{{}^{\prime}}}f|πc(P(X~k))\scriptstyle{f|_{\pi_{c}^{{}^{\prime}}(P(\widetilde{X}_{k}))}}

By Lemma 3.2, the lower horizontal map is a homeomorphism. The second statement of the Lemma follows by the similar arguments with P(X~k)P(\widetilde{X}_{k}) is replaced by P(E~(λk))P(\widetilde{E}(\lambda^{k})). ∎

Theorem 3.4.

The collection {E(λi)G(ci/rk)}i=0k\{\frac{E(\lambda^{i})}{G(c_{i}/r_{k})}\}_{i=0}^{k} gives a qq-cell structure of πcrk(P(X~k))\pi^{{}^{\prime}}_{\frac{c}{r_{k}}}(P(\widetilde{X}_{k})) for k=1,2,,mk=1,2,\dots,m. Moreover, {E(λi)/G(ci/ri)}i=0m\{E(\lambda^{i})/G(c_{i}/r_{i})\}_{i=0}^{m} gives a qq-cell structure of WGr(d,n){\rm WGr}(d,n) where r0=c0r_{0}=c_{0}.

Proof.

Note that the sets P(E~(λi))P(\widetilde{E}(\lambda^{i})) and P(Md(n,d))=i=0mP(E~(λi))P(M_{d}(n,d))=\sqcup_{i=0}^{m}P(\widetilde{E}(\lambda^{i})) are invariant under the weighted \mathbb{C}^{*}-action defined in Remark 2.3. Then we have the following commutative diagram.

P(X~k){P(\widetilde{X}_{k})}{\subset}k+1{0}{\mathbb{C}^{k+1}-\{0\}}πcrk(P(X~k)){\pi_{\frac{c}{r_{k}}}^{{}^{\prime}}(P(\widetilde{X}_{k}))}{\subset}𝕎P(c0rk,c1rk,,ckrk).{\mathbb{W}P(\frac{c_{0}}{r_{k}},\frac{c_{1}}{r_{k}},\dots,\frac{c_{k}}{r_{k}}).}πcrk\scriptstyle{\pi_{\frac{c}{r_{k}}}^{{}^{\prime}}}πcrk\scriptstyle{\pi_{\frac{c}{r_{k}}}^{{}^{\prime}}}

Thus the first part follows from

πcrk(P(X~k))\displaystyle\pi^{{}^{\prime}}_{\frac{c}{r_{k}}}(P(\widetilde{X}_{k})) =πcrk(i=0kP(E~(λi)))\displaystyle=\pi^{{}^{\prime}}_{\frac{c}{r_{k}}}(\sqcup_{i=0}^{k}P(\widetilde{E}(\lambda^{i})))
=i=0kπcrk(P(E~(λi)))\displaystyle=\sqcup_{i=0}^{k}\pi^{{}^{\prime}}_{\frac{c}{r_{k}}}(P(\widetilde{E}(\lambda^{i})))
=i=0kP(E~(λi))c/rki=0kE(λi)G(ci/rk).\displaystyle=\sqcup_{i=0}^{k}\frac{P(\widetilde{E}(\lambda^{i}))}{\sim_{c/r_{k}}}\cong\sqcup_{i=0}^{k}\frac{E(\lambda^{i})}{G(c_{i}/r_{k})}.

The second part follows from WGr(d,n)πc(P(X~m))\mbox{WGr}(d,n)\cong\pi_{c}^{\prime}(P(\widetilde{X}_{m})) and by applying Lemma 3.3 successively for every k{1,2,,m}k\in\{1,2,\dots,m\}. ∎

We show that two weighted Grassmann orbifolds are weakly equivariantly homeomorphic if the associated weight vectors are differed by a permutation σSn\sigma\in S_{n}. Let X,YX,Y be two GG-spaces. A map f:XYf\colon X\to Y is called a weakly equivariant homeomorphism if ff is a homeomorphism and f(gx)=η(g)f(x)f(gx)=\eta(g)f(x) for some ηAut(G)\eta\in{\rm Aut}(G) and for all (g,x)G×X(g,x)\in G\times X. If η\eta is identity, then ff is called an equivariant homeomorphism.

Theorem 3.5.

Let W=(w1,w2,,wn)(0)nW=(w_{1},w_{2},\dots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n}, 0<a0<a\in\mathbb{Z} and σW:=(wσ1,wσ2,,wσn)\sigma W:=(w_{\sigma_{1}},w_{\sigma_{2}},\dots,w_{\sigma_{n}}) for some σSn\sigma\in S_{n}. If WGr(d,n){\rm WGr}(d,n) and WGr(d,n){\rm WGr}^{\prime}(d,n) are the weighted Grassmann orbifolds associated to (W,a)(W,a) and (σW,a)(\sigma W,a) respectively, then WGr(d,n){\rm WGr}(d,n) is weakly equivariantly homeomorphic to WGr(d,n){\rm WGr}^{\prime}(d,n). Moreover, this may induce different qq-cell structures on WGr(d,n){\rm WGr}(d,n) for different σ\sigma.

Proof.

The matrix A=(aij)Md(n,d)A=(a_{ij})\in M_{d}(n,d) if and only if σA=(aσij)Md(n,d)\sigma A=(a_{\sigma_{i}j})\in M_{d}(n,d). Thus the natural weakly equivariant homeomorphism f¯σ:Md(n,d)Md(n,d)\bar{f}_{\sigma}\colon M_{d}(n,d)\to M_{d}(n,d) defined by f¯σ(A)=σA\bar{f}_{\sigma}(A)=\sigma A induces the following commutative diagram.

(3.1) Md(n,d){M_{d}(n,d)}Md(n,d){M_{d}(n,d)}WGr(d,n){\mbox{WGr}(d,n)}WGr(d,n).{{\rm WGr}^{\prime}(d,n).}f¯σ\scriptstyle{\bar{f}_{\sigma}}πw\scriptstyle{\pi_{w}}πσw\scriptstyle{\pi_{\sigma w}}fσ\scriptstyle{f_{\sigma}}

Here πw\pi_{w} is the quotient map defined in Definition 2.1. Thus, (3.1) induces a weakly equivariant homeomorphism fσ:WGr(d,n)WGr(d,n)f_{\sigma}\colon\mbox{WGr}(d,n)\to{\rm WGr}^{\prime}(d,n), where ()n(\mathbb{C}^{*})^{n}-action is differed by the permutation σ\sigma. Note that fσ([A]w)=[σA]w{f}_{\sigma}([A]_{\sim w})=[\sigma A]_{\sim w}.

We discuss the effects of the permutation σ\sigma on the qq-cell structure on WGr(d,n)\mbox{WGr}(d,n). Consider i={(x1,x2,,xn)n|xj=0forj>i}\mathbb{C}^{i}=\{(x_{1},x_{2},\dots,x_{n})\in\mathbb{C}^{n}~{}|~{}x_{j}=0~{}\text{for}~{}j>i\}. For σSn\sigma\in S_{n}, define

σn:={(xσ1,xσ2,,xσn)}\sigma\mathbb{C}^{n}:=\{(x_{\sigma_{1}},x_{\sigma_{2}},\dots,x_{\sigma_{n}})\}

and

σi:={(xσ1,xσ2,,xσn)σn|xσj=0forσj>i}.\sigma\mathbb{C}^{i}:=\{(x_{\sigma_{1}},x_{\sigma_{2}},\dots,x_{\sigma_{n}})\in\sigma\mathbb{C}^{n}~{}|~{}x_{\sigma_{j}}=0~{}\text{for}~{}\sigma_{j}>i\}.

Let λ=(λ1,,λd)\lambda=(\lambda_{1},\ldots,\lambda_{d}) be a Schubert symbol for d<nd<n. Then

σE(λ)\displaystyle\sigma E(\lambda) ={σY|YE(λ)}\displaystyle=\{\sigma Y~{}|~{}Y\in E(\lambda)\}
={XGr(d,n)|dim(Xσλi)=i,dim(Xσλi1)=i1,i[d]}\displaystyle=\{X\in\mbox{Gr}(d,n)~{}|~{}\text{dim}(X\cap\sigma\mathbb{C}^{\lambda_{i}})=i,~{}\text{dim}(X\cap\sigma\mathbb{C}^{\lambda_{i}-1})=i-1,i\in[d]\}

where [d]={1,2,,d}[d]=\{1,2,\ldots,d\}. Then E(λ)σE(λ)E(\lambda)\cong\sigma E(\lambda) and dim(σE(λ))=(λ)(\sigma E(\lambda))=\ell(\lambda).

So the permutation of the coordinates in n\mathbb{C}^{n} determines another cell structure for Gr(d,n)\mbox{Gr}(d,n) given by Gr(d,n)=σGr(d,n)=i=0mσE(λi)\mbox{Gr}(d,n)=\sigma\mbox{Gr}(d,n)=\sqcup_{i=0}^{m}\sigma E(\lambda^{i}). This cell structure of Gr(d,n)\mbox{Gr}(d,n) induces the following decomposition of Md(n,d)M_{d}(n,d) which is similar to (2.14).

Md(n,d)=i=0mσE~(λi)andP(Md(n,d))=i=0mP(σE~(λi)).M_{d}(n,d)=\sqcup_{i=0}^{m}\sigma\widetilde{E}(\lambda^{i})~{}\mbox{and}~{}P(M_{d}(n,d))=\sqcup_{i=0}^{m}P(\sigma\widetilde{E}(\lambda^{i})).

Recall that λi=(λ1i,,λdi)\lambda^{i}=(\lambda_{1}^{i},\ldots,\lambda_{d}^{i}) is a Schubert symbol and cic_{i} is defined in (2.5) for i=0,,mi=0,\ldots,m. Then σλi:=(σ(λi1i),,σ(λidi))\sigma\lambda^{i}:=(\sigma(\lambda_{i_{1}}^{i}),\ldots,\sigma(\lambda_{i_{d}}^{i})), where i1,,id{1,,d}i_{1},\dots,i_{d}\in\{1,\dots,d\} such that σ(λi1i)<σ(λi2i)<<σ(λidi)\sigma(\lambda_{i_{1}}^{i})<\sigma(\lambda_{i_{2}}^{i})<\cdots<\sigma(\lambda_{i_{d}}^{i}). Let

(3.2) σci:=a+j=1dwσ(λiji).\sigma c_{i}:=a+\sum_{j=1}^{d}w_{\sigma(\lambda_{i_{j}}^{i})}.

Now from the commutativity of the diagram (2.12), we have the following.

πw(σ(E~(λi)))Plw(πw(σE~(λi)))=P(σE~(λi)) weighted -action σE(λi)G(σci).\pi_{w}(\sigma(\widetilde{E}(\lambda^{i})))\cong Pl_{w}(\pi_{w}(\sigma\widetilde{E}(\lambda^{i})))=\frac{P(\sigma\widetilde{E}(\lambda^{i}))}{\mbox{ weighted }\mathbb{C}^{*}\mbox{-action }}\cong\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i})}.

Then we get a qq-cell structure of the weighted Grassmann orbifold WGr(d,n)\mbox{WGr}(d,n) given by

WGr(d,n)σE(λ0)G(σc0)σE(λ1)G(σc1)σE(λm)G(σcm).\mbox{WGr}(d,n)\cong\frac{\sigma E(\lambda^{0})}{G(\sigma c_{0})}\sqcup\frac{\sigma E(\lambda^{1})}{G(\sigma c_{1})}\sqcup\dots\sqcup\frac{\sigma E(\lambda^{m})}{G(\sigma c_{m})}.

Remark 3.6.

Applying the permutation σ\sigma on the rows of the matrices in E(λ)E(\lambda), we get the matrices of σE(λ)\sigma E(\lambda). That is,

(v1v2vn)E(λ)(vσ1vσ2vσn)σE(λ).\begin{pmatrix}v_{1}\\ v_{2}\\ \vdots\\ v_{n}\end{pmatrix}\in E(\lambda)\iff\begin{pmatrix}v_{\sigma_{1}}\\ v_{\sigma_{2}}\\ \vdots\\ v_{\sigma_{n}}\end{pmatrix}\in\sigma E(\lambda).
Proposition 3.7.

[BNSS21, Theorem 1.1] Let XX be a qq-CW complex with no odd dimensional qq-cells and pp a prime number. Let {pt}=X0X1Xs=X\{pt\}=X_{0}\subseteq X_{1}\subseteq\dots\subseteq X_{s}=X is a filtration of XX such that XiX_{i} is obtained by attaching the qq-cell 2ki/Gi\mathbb{R}^{2k_{i}}/G_{i} to Xi1X_{i-1} for all i{1,2,,s}i\in\{1,2,\dots,s\}. If gcd{p,|Gi|}=1\gcd\{p,|G_{i}|\}=1 for all i{1,2,,s}i\in\{1,2,\dots,s\}, then H(X;)H^{*}(X;\mathbb{Z}) has no pp-torsion and Hodd(X;p)H^{odd}(X;\mathbb{Z}_{p}) is trivial.

Recall the definition of σci\sigma c_{i} from (3.2) for WGr(d,n)\mbox{WGr}(d,n) associated to the weight W=(w1,,wn)(0)nW=(w_{1},\dots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n} and 1a1\leq a\in\mathbb{Z}.

Definition 3.8.

A permutation σSn\sigma\in S_{n} is called admissible for a prime pp and WGr(d,n){\rm WGr}(d,n) if

gcd{p,σcidi}=1\gcd\{p,\frac{\sigma c_{i}}{d_{i}}\}=1

where σci\sigma c_{i} is defined in (3.2) and di=gcd{σc0,σc1,,σci}d_{i}=\gcd\{\sigma c_{0},\sigma c_{1},\dots,\sigma c_{i}\}, for i{1,2,,m}i\in\{1,2,\dots,m\}.

Some examples of admissible permutations are discussed in Example 3.12.

Remark 3.9.

There may not always exist an admissible permutation σSn\sigma\in S_{n} for a prime pp and WGr(d,n){\rm WGr}(d,n). However if d=1d=1, then m=n1m=n-1 and there always exists an admissible permutation σSn\sigma\in S_{n} for every prime pp. The admissible permutation σSn\sigma\in S_{n} may not be unique.

Now we prove the following result which says when the integral cohomology of WGr(d,n)\mbox{WGr}(d,n) have no pp-torsion.

Theorem 3.10.

If there exists an admissible permutation σSn\sigma\in S_{n} for a prime pp and WGr(d,n){\rm WGr}(d,n), then H(WGr(d,n);)H^{*}({\rm WGr}(d,n);\mathbb{Z}) has no pp-torsion and Hodd(WGr(d,n);p)H^{odd}({\rm WGr}(d,n);\mathbb{Z}_{p}) is trivial.

Proof.

Suppose σSn\sigma\in S_{n} be the admissible permutation for pp and WGr(d,n)\mbox{WGr}(d,n). Then

gcd{p,σcidi}=1\gcd\{p,\frac{\sigma c_{i}}{d_{i}}\}=1

by Definition 3.8, where di=gcd{σc0,σc1,,σci}d_{i}=\gcd\{\sigma c_{0},\sigma c_{1},\dots,\sigma c_{i}\} for all i{1,2,,m}i\in\{1,2,\dots,m\}. By Theorem 3.5, we have the following qq-cell structure

WGr(d,n)σE(λ0)G(σc0)σE(λ1)G(σc1)σE(λm)G(σcm),\mbox{WGr}(d,n)\cong\frac{\sigma E(\lambda^{0})}{G(\sigma c_{0})}\sqcup\frac{\sigma E(\lambda^{1})}{G(\sigma c_{1})}\sqcup\dots\sqcup\frac{\sigma E(\lambda^{m})}{G(\sigma c_{m})},

where σE(λi)E(λi)(λi)\sigma E(\lambda^{i})\cong E(\lambda^{i})\cong\mathbb{C}^{\ell(\lambda^{i})}. Let

σXk=i=0kσE(λi)G(σci)WGr(d,n) for k=0,1,,m.\sigma X_{k}=\sqcup_{i=0}^{k}\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i})}\subseteq\mbox{WGr}(d,n)\text{ for }k=0,1,\ldots,m.

Then σXk\sigma X_{k} is a subcomplex of WGr(d,n)\mbox{WGr}(d,n) for k=0,1,,mk=0,1,\dots,m and σXm=WGr(d,n)\sigma X_{m}=\mbox{WGr}(d,n). This gives a filtration

{pt}=σX0σX1σXm=WGr(d,n)\{pt\}=\sigma X_{0}\subset\sigma X_{1}\subset\dots\subset\sigma X_{m}=\mbox{WGr}(d,n)

such that σXiσXi1\sigma X_{i}-\sigma X_{i-1} is homeomorphic to σE(λi)G(σci)\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i})}.

Using Lemma 3.3, σE(λi)G(σci)σE(λi)G(σcidi)\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i})}\cong\frac{\sigma E(\lambda^{i})}{G(\frac{\sigma c_{i}}{d_{i}})}. That is σXiσXi1\sigma X_{i}-\sigma X_{i-1} is homeomorphic to (λi)G(σcidi)\frac{\mathbb{C}^{\ell(\lambda^{i})}}{G(\frac{\sigma c_{i}}{d_{i}})} for all i=1,2,,mi=1,2,\dots,m. Therefore, by Proposition 3.7, H(WGr(d,n);)H^{*}(\mbox{WGr}(d,n);\mathbb{Z}) has no pp-torsion and the group Hodd(WGr(d,n);p)H^{odd}(\mbox{WGr}(d,n);\mathbb{Z}_{p}) is trivial. This completes the proof. ∎

Corollary 3.11.

[Kaw73] H(𝕎P(c0,c1,,cm);)H^{*}(\mathbb{W}P(c_{0},c_{1},\dots,c_{m});\mathbb{Z}) has no torsion.

Proof.

This follows from Theorem 3.10, Remark 2.3 and 3.9. ∎

Example 3.12.

Consider the weighted Grassmann orbifold WGr(2,4)\mbox{WGr}(2,4) for W=(1,1,3,4)W=(1,1,3,4) and a=2a=2. Here n=4,d=2,(nd)=6,m=(nd)1=5n=4,~{}d=2,~{}{n\choose d}=6,~{}m={n\choose d}-1=5. So in this case, we have 66 Schubert symbols which are

λ0=(1,2)<λ1=(1,3)<λ2=(1,4)<λ3=(2,3)<λ4=(2,4)<λ5=(3,4)\lambda^{0}=(1,2)<\lambda^{1}=(1,3)<\lambda^{2}=(1,4)<\lambda^{3}=(2,3)<\lambda^{4}=(2,4)<\lambda^{5}=(3,4)

in the ordering as in Definition 2.4. For the prime p=3p=3, consider the permutation σS4\sigma\in S_{4} defined by

σ1=3,σ2=4,σ3=1,σ4=2.\sigma_{1}=3,\sigma_{2}=4,\sigma_{3}=1,\sigma_{4}=2.

Then

σc0=9,σc1=6,σc2=6,σc3=7,σc4=7 and σc5=4\sigma c_{0}=9,\sigma c_{1}=6,\sigma c_{2}=6,\sigma c_{3}=7,\sigma c_{4}=7\text{ and }\sigma c_{5}=4

using (3.2). This σ\sigma is admissible for p=3p=3 and WGr(2,4)\mbox{WGr}(2,4). Thus H(WGr(2,4);)H^{*}(\mbox{WGr}(2,4);\mathbb{Z}) has no 33-torsion by Theorem 3.10.

For the prime p=7p=7, consider the permutation σS4\sigma\in S_{4} defined by

σ1=4,σ2=2,σ3=1,σ4=3.\sigma_{1}=4,\sigma_{2}=2,\sigma_{3}=1,\sigma_{4}=3.

Then

σc0=7,σc1=7,σc2=9,σc3=4,σc4=6 and σc5=6\sigma c_{0}=7,\sigma c_{1}=7,\sigma c_{2}=9,\sigma c_{3}=4,\sigma c_{4}=6\text{ and }\sigma c_{5}=6

using (3.2). This σ\sigma is admissible for p=7p=7 and WGr(2,4)\mbox{WGr}(2,4). Thus H(WGr(2,4);)H^{*}(\mbox{WGr}(2,4);\mathbb{Z}) has no 77-torsion by Theorem 3.10.

To compute that it has no 2-torsion, we need to consider a different total order on the Schubert symbols given by

λ0=(1,2)<λ1=(1,3)<λ2=(2,3)<λ3=(1,4)<λ4=(2,4)<λ5=(3,4)\lambda^{0}=(1,2)<\lambda^{1}=(1,3)<\lambda^{2}=(2,3)<\lambda^{3}=(1,4)<\lambda^{4}=(2,4)<\lambda^{5}=(3,4)

which preserves the partial order in (2.1). In this case,

c0=4,c1=6,c2=6,c3=7,c4=7 and c5=9c_{0}=4,c_{1}=6,c_{2}=6,c_{3}=7,c_{4}=7\text{ and }c_{5}=9

using (2.5). The identity permutation in S4S_{4} is admissible for p=2p=2 and this WGr(2,4)\mbox{WGr}(2,4). Then H(WGr(2,4);)H^{*}(\mbox{WGr}(2,4);\mathbb{Z}) has no 22-torsion by Theorem 3.10.

The only primes which divides the orders of the orbifold singularities of this WGr(2,4)\mbox{WGr}(2,4) are 2,32,3 and 77. Hence the integral cohomology of WGr(2,4)\mbox{WGr}(2,4) of this example has no torsion. ∎

Remark 3.13.

Considering the total order given in Definition 2.4 on the Schubert symbols, there may not exist an admissible permutation σ\sigma for a prime. However, one can take another total order on the Schubert symbols for which one can find σ\sigma satisfying the hypothesis in Theorem 3.10 for this prime.

The qq-cell structure in Theorem 3.4 leads us to introduce the following definition which generalizes the concept of divisive weighted projective spaces of [HHRW16].

Definition 3.14.

A weighted Grassmann orbifold WGr(d,n){\rm WGr}(d,n) is called divisive if there exists σSn\sigma\in S_{n} such that σci{\sigma c_{i}} divides σci1{\sigma c_{i-1}} for i=1,2,,mi=1,2,\dots,m where σci\sigma c_{i} is defined in (3.2).

Example 3.15.

Consider the weighted Grassmann orbifold WGr(2,4)\mbox{WGr}(2,4) for the weight W=(1,6,1,1)W=(1,6,1,1) and a=3a=3. We have the ordering on the 66 Schubert symbols given by

λ0=(1,2)<λ1=(1,3)<λ2=(1,4)<λ3=(2,3)<λ4=(2,4)<λ5=(3,4).\lambda^{0}=(1,2)<\lambda^{1}=(1,3)<\lambda^{2}=(1,4)<\lambda^{3}=(2,3)<\lambda^{4}=(2,4)<\lambda^{5}=(3,4).

Consider the permutation σS4\sigma\in S_{4} defined by

σ1=2,σ2=1,σ3=3,σ4=4.\sigma_{1}=2,\sigma_{2}=1,\sigma_{3}=3,\sigma_{4}=4.

Then

σc0=10,σc1=10,σc2=10,σc3=5,σc4=5,σc5=5\sigma c_{0}=10,\sigma c_{1}=10,\sigma c_{2}=10,\sigma c_{3}=5,\sigma c_{4}=5,\sigma c_{5}=5

using (3.2). Thus σci\sigma c_{i} divides σci1\sigma c_{i-1} for i=1,2,,5i=1,2,\dots,5. So WGr(2,4)\mbox{WGr}(2,4) of this example is divisive. ∎

Example 3.16.

Let α\alpha and γ\gamma be any two non-negative integers and β\beta be any positive integer such that β>dα\beta>d\alpha. Let WGr(d,n)\mbox{WGr}(d,n) be the corresponding weighted Grassmann orbifold for W=(α+γβ,α,,α)(0)nW=(\alpha+\gamma\beta,\alpha,\dots,\alpha)\in(\mathbb{Z}_{\geq 0})^{n} and a=βdα>0a=\beta-d\alpha>0. Consider the total order {λ0,λ1,,λm}\{\lambda^{0},\lambda^{1},\dots,\lambda^{m}\} on the Schubert symbol induced by the dictionary order. Then

ci={(γ+1)βifi=0,1,,(n1d1)1βifi=(n1d1),,m.\displaystyle c_{i}=\begin{cases}(\gamma+1)\beta&\quad\text{if}~{}i=0,1,\dots,{n-1\choose d-1}-1\\ \beta&\quad\text{if}~{}i={n-1\choose d-1},\dots,m.\end{cases}

Then cic_{i} divides ci1c_{i-1} for all i=1,2,,mi=1,2,\dots,m. Therefore this WGr(d,n)\mbox{WGr}(d,n) is a divisive weighted Grassmann orbifold. ∎

Definition 3.17.

Let λ\lambda be a Schubert symbol for d<nd<n. Then a reversal of λ\lambda is a pair (k,k)(k,k^{\prime}) such that kλk\in\lambda, kλk^{\prime}\notin\lambda and k<kk^{\prime}<k. We denote the set of all reversals of λ\lambda by rev(λ)\rm{rev}(\lambda). If (k,k)rev(λ)(k,k^{\prime})\in\rm{rev}(\lambda) then (k,k)λ(k,k^{\prime})\lambda is the Schubert symbol obtained by replacing kk by kk^{\prime} in λ\lambda and ordering the later set.

Remark 3.18.

If (k,k)rev(λ)(k,k^{\prime})\in\text{rev}(\lambda) then (k,k)λλ(k,k^{\prime})\lambda\prec\lambda and (λ)\ell(\lambda) is the cardinality of the set rev(λ)\text{rev}(\lambda) where (λ)\ell(\lambda) is the length of λ\lambda. In [KT03, AM15] the authors defined an inversion of a Schubert symbol λ\lambda is a pair (k,k)(k,k^{\prime}) such that kλ,kλk\in\lambda,k^{\prime}\notin\lambda and k<kk<k^{\prime}. In some sense, our definition of reversal is dual to the definition of inversion. If inv(λ)\rm{inv}(\lambda) be the set of all inversions of λ\lambda and (λ)\ell^{{}^{\prime}}(\lambda) is the cardinality of the set inv(λ)\rm{inv}(\lambda) then (λ)+(λ)=d(nd)\ell(\lambda)+\ell^{{}^{\prime}}(\lambda)=d(n-d). Also If (k,k)rev(λ)(k,k^{\prime})\in\rm{rev}(\lambda) and (k,k)λ=μ(k,k^{\prime})\lambda=\mu then (k,k)inv(μ)(k^{\prime},k)\in\rm{inv}(\mu) and and (k,k)μ=λ(k^{\prime},k)\mu=\lambda.

Next we discuss ()n(\mathbb{C}^{*})^{n}-action on some cell structure of a divisive weighted Grassmann orbifold. Recall the ()n(\mathbb{C}^{*})^{n}-action on WGr(d,n)\mbox{WGr}(d,n) which is induced from (2.7). We adhere the notation from Section 2.

Theorem 3.19.

If WGr(d,n){\rm WGr}(d,n) is a divisive weighted Grassmann orbifold then it has a ()n(\mathbb{C}^{*})^{n}-invariant cell structure with only even dimensional cells.

Proof.

Let WGr(d,n)\mbox{WGr}(d,n) be a divisive weighted Grassmann orbifold corresponding to W=(w1,,wn)(0)nW=(w_{1},\ldots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n} and 1a1\leq a\in\mathbb{Z}. Then there exists σSn\sigma\in S_{n} such that σci\sigma c_{i} divides σci1\sigma c_{i-1} for all i=1,2,,mi=1,2,\dots,m. Let us assume σ=Id\sigma=\text{Id} (the identity permutation in SnS_{n}). Then cic_{i} divides ci1c_{i-1} for all i=1,2,,mi=1,2,\dots,m. Then gcd{c0,c1,,ci}=ci\gcd\{c_{0},c_{1},\dots,c_{i}\}=c_{i} for all i{1,2,,m}i\in\{1,2,\dots,m\}. Thus

πw(E~(λi))E(λi)G(ci)E(λi)G(ci/ci)E(λi) for all i=1,2,,m\pi_{w}(\widetilde{E}(\lambda^{i}))\cong\frac{E(\lambda^{i})}{G(c_{i})}\cong\frac{E(\lambda^{i})}{G(c_{i}/c_{i})}\cong E(\lambda^{i})\text{ for all }i=1,2,\dots,m

by Lemma 3.3. Thus each element of πw(E~(λi))\pi_{w}(\widetilde{E}(\lambda^{i})) can be represented uniquely by the equivalence class of an n×dn\times d matrix defined in (2.13).

Let λi=(λ1,,λd)\lambda^{i}=(\lambda_{1},\ldots,\lambda_{d}) be a Schubert symbol for d<nd<n and 𝐳(λi){\bf z}\in\mathbb{C}^{\ell(\lambda^{i})}. Since (λi)=(λ11)+(λ22)++(λdd)\ell(\lambda^{i})=(\lambda_{1}-1)+(\lambda_{2}-2)+\dots+(\lambda_{d}-d), we can write

𝐳=(𝐳1,𝐳2,,𝐳d){\bf z}=({\bf z}_{1},{\bf z}_{2},\ldots,{\bf z}_{d})

where 𝐳l=(z1l,z2l,,zλ1l^,,zλ2l^,,zλl1l^,,zλl1l){\bf z}_{l}=(z^{l}_{1},z^{l}_{2},\dots,\widehat{z^{l}_{\lambda_{1}}},\dots,\widehat{z^{l}_{\lambda_{2}}},\dots,\widehat{z^{l}_{\lambda_{l-1}}},\dots,z^{l}_{\lambda_{l}-1}) for l=1,,dl=1,\ldots,d.

For (t1,,tn)()n(t_{1},\ldots,t_{n})\in(\mathbb{C}^{*})^{n}, we define ss\in\mathbb{C}^{*} such that sci=tλ1tλds^{c_{i}}=t_{\lambda_{1}}\cdots t_{\lambda_{d}}. Define TGL(d,)T\in\mbox{GL}(d,\mathbb{C}) by

T=diag((tλ1swλ1),(tλ2swλ2),,(tλdswλd)).T=\mbox{diag}((\dfrac{t_{\lambda_{1}}}{s^{w_{\lambda_{1}}}}),(\dfrac{t_{\lambda_{2}}}{s^{w_{\lambda_{2}}}}),\dots,(\dfrac{t_{\lambda_{d}}}{s^{w_{\lambda_{d}}}})).

Then det(T)=sa\det(T)=s^{a}.

Define gλi:(λi)πw(E~(λi))g_{\lambda^{i}}\colon\mathbb{C}^{\ell(\lambda^{i})}\to\pi_{w}(\widetilde{E}(\lambda^{i})) by

gλi(𝐳):=[z11z12z1dzλ111zλ112zλ11d1000zλ1+12zλ1+1d0zλ212zλ21d01000zλ2+1d00zλd1d001000000].g_{\lambda^{i}}({\bf z}):=\begin{bmatrix}z^{1}_{1}&z^{2}_{1}&\dots&z^{d}_{1}\\ \vdots&\vdots&&\vdots\\ z^{1}_{\lambda_{1}-1}&z^{2}_{\lambda_{1}-1}&\dots&z^{d}_{\lambda_{1}-1}\\ 1&0&\dots&0\\ 0&z^{2}_{\lambda_{1}+1}&\dots&z^{d}_{\lambda_{1}+1}\\ \vdots&\vdots&&\vdots\\ 0&z^{2}_{\lambda_{2}-1}&\dots&z^{d}_{\lambda_{2}-1}\\ 0&1&\dots&0\\ 0&0&\dots&z^{d}_{\lambda_{2}+1}\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&z^{d}_{\lambda_{d}-1}\\ 0&0&\dots&1\\ 0&0&\dots&0\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&0\\ \end{bmatrix}.

Then gλig_{\lambda^{i}} is a homeomorphism. Now we have

(t1,t2,,tn)gλi(𝐳)=[t1z11t1z12t1z1dtλ11zλ111tλ11zλ112tλ11zλ11dtλ1000tλ1+1zλ1+12tλ1+1zλ1+1d0tλ21zλ212tλ21zλ21d0tλ2000tλ2+1zλ2+1d00tλd1zλd1d00tλd000000].(t_{1},t_{2},\dots,t_{n})g_{\lambda^{i}}({\bf z})=\begin{bmatrix}t_{1}z^{1}_{1}&t_{1}z^{2}_{1}&\dots&t_{1}z^{d}_{1}\\ \vdots&\vdots&&\vdots\\ t_{\lambda_{1}-1}z^{1}_{\lambda_{1}-1}&t_{\lambda_{1}-1}z^{2}_{\lambda_{1}-1}&\dots&t_{\lambda_{1}-1}z^{d}_{\lambda_{1}-1}\\ t_{\lambda_{1}}&0&\dots&0\\ 0&t_{\lambda_{1}+1}z^{2}_{\lambda_{1}+1}&\dots&t_{\lambda_{1}+1}z^{d}_{\lambda_{1}+1}\\ \vdots&\vdots&&\vdots\\ 0&t_{\lambda_{2}-1}z^{2}_{\lambda_{2}-1}&\dots&t_{\lambda_{2}-1}z^{d}_{\lambda_{2}-1}\\ 0&t_{\lambda_{2}}&\dots&0\\ 0&0&\dots&t_{\lambda_{2}+1}z^{d}_{\lambda_{2}+1}\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&t_{\lambda_{d}-1}z^{d}_{\lambda_{d}-1}\\ 0&0&\dots&t_{\lambda_{d}}\\ 0&0&\dots&0\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&0\\ \end{bmatrix}.

Then

(t1,t2,,tn)gλi(𝐳)=[swλ1tλ1t1z11swλ2tλ2t1z12swλdtλdt1z1dswλ1tλ1tλ11zλ111swλ2tλ2tλ11zλ112swλdtλdtλ11zλ11dswλ1tλ1tλ1000swλ2tλ2tλ1+1zλ1+12swλdtλdtλ1+1zλ1+1d0swλ2tλ2tλ21zλ212swλdtλdtλ21zλ21d0swλ2tλ2tλ2000swλdtλdtλ2+1zλ2+1d00swλdtλdtλd1zλd1d00swλdtλdtλd000000]×T.(t_{1},t_{2},\dots,t_{n})g_{\lambda^{i}}({\bf z})=\begin{bmatrix}\frac{s^{w_{\lambda_{1}}}}{t_{\lambda_{1}}}t_{1}z^{1}_{1}&\frac{s^{w_{\lambda_{2}}}}{t_{\lambda_{2}}}t_{1}z^{2}_{1}&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}}t_{1}z^{d}_{1}\\ \vdots&\vdots&&\vdots\\ \frac{s^{w_{\lambda_{1}}}}{t_{\lambda_{1}}}t_{\lambda_{1}-1}z^{1}_{\lambda_{1}-1}&\frac{s^{w_{\lambda_{2}}}}{t_{\lambda_{2}}}t_{\lambda_{1}-1}z^{2}_{\lambda_{1}-1}&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}}t_{\lambda_{1}-1}z^{d}_{\lambda_{1}-1}\\ \frac{s^{w_{\lambda_{1}}}}{t_{\lambda_{1}}}t_{\lambda_{1}}&0&\dots&0\\ 0&\frac{s^{w_{\lambda_{2}}}}{t_{\lambda_{2}}}t_{\lambda_{1}+1}z^{2}_{\lambda_{1}+1}&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}}t_{\lambda_{1}+1}z^{d}_{\lambda_{1}+1}\\ \vdots&\vdots&&\vdots\\ 0&\frac{s^{w_{\lambda_{2}}}}{t_{\lambda_{2}}}t_{\lambda_{2}-1}z^{2}_{\lambda_{2}-1}&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}}t_{\lambda_{2}-1}z^{d}_{\lambda_{2}-1}\\ 0&\frac{s^{w_{\lambda_{2}}}}{t_{\lambda_{2}}}t_{\lambda_{2}}&\dots&0\\ 0&0&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}}t_{\lambda_{2}+1}z^{d}_{\lambda_{2}+1}\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}}t_{\lambda_{d}-1}z^{d}_{\lambda_{d}-1}\\ 0&0&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}}t_{\lambda_{d}}\\ 0&0&\dots&0\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&0\\ \end{bmatrix}\times T.
Thus, (t1,t2,,tn)gλi(𝐳)\displaystyle\text{ Thus, }(t_{1},t_{2},\dots,t_{n})g_{\lambda^{i}}({\bf z})
=D×[swλ1tλ1sw1t1z11swλ2tλ2sw1t1z12swλdtλdsw1t1z1dswλ1swλ11tλ1tλ11zλ111swλ2swλ11tλ2tλ11zλ112swλdswλ11tλdtλ11zλ11d1000swλ2swλ1+1tλ2tλ1+1zλ1+12swλdswλ1+1tλdtλ1+1zλ1+1d0swλ2swλ21tλ2tλ21zλ212swλdswλ21tλdtλ21zλ21d01000swλdswλ2+1tλdtλ2+1zλ2+1d00swλdswλd1tλdtλd1zλd1d001000000]×T\displaystyle=D\times\begin{bmatrix}\frac{s^{w_{\lambda_{1}}}}{t_{\lambda_{1}}s^{w_{1}}}t_{1}z^{1}_{1}&\frac{s^{w_{\lambda_{2}}}}{t_{\lambda_{2}}s^{w_{1}}}t_{1}z^{2}_{1}&\dots&\frac{s^{w_{\lambda_{d}}}}{t_{\lambda_{d}}s^{w_{1}}}t_{1}z^{d}_{1}\\ \vdots&\vdots&&\vdots\\ \frac{s^{w_{\lambda_{1}}}}{s^{w_{\lambda_{1}-1}}t_{\lambda_{1}}}t_{\lambda_{1}-1}z^{1}_{\lambda_{1}-1}&\frac{s^{w_{\lambda_{2}}}}{s^{w_{\lambda_{1}-1}}t_{\lambda_{2}}}t_{\lambda_{1}-1}z^{2}_{\lambda_{1}-1}&\dots&\frac{s^{w_{\lambda_{d}}}}{s^{w_{\lambda_{1}-1}}t_{\lambda_{d}}}t_{\lambda_{1}-1}z^{d}_{\lambda_{1}-1}\\ 1&0&\dots&0\\ 0&\frac{s^{w_{\lambda_{2}}}}{s^{w_{\lambda_{1}+1}}t_{\lambda_{2}}}t_{\lambda_{1}+1}z^{2}_{\lambda_{1}+1}&\dots&\frac{s^{w_{\lambda_{d}}}}{s^{w_{\lambda_{1}+1}}t_{\lambda_{d}}}t_{\lambda_{1}+1}z^{d}_{\lambda_{1}+1}\\ \vdots&\vdots&&\vdots\\ 0&\frac{s^{w_{\lambda_{2}}}}{s^{w_{\lambda_{2}-1}}t_{\lambda_{2}}}t_{\lambda_{2}-1}z^{2}_{\lambda_{2}-1}&\dots&\frac{s^{w_{\lambda_{d}}}}{s^{w_{\lambda_{2}-1}}t_{\lambda_{d}}}t_{\lambda_{2}-1}z^{d}_{\lambda_{2}-1}\\ 0&1&\dots&0\\ 0&0&\dots&\frac{s^{w_{\lambda_{d}}}}{s^{w_{\lambda_{2}+1}}t_{\lambda_{d}}}t_{\lambda_{2}+1}z^{d}_{\lambda_{2}+1}\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&\frac{s^{w_{\lambda_{d}}}}{s^{w_{\lambda_{d}-1}}t_{\lambda_{d}}}t_{\lambda_{d}-1}z^{d}_{\lambda_{d}-1}\\ 0&0&\dots&1\\ 0&0&\dots&0\\ \vdots&\vdots&&\vdots\\ 0&0&\dots&0\\ \end{bmatrix}\times T
=DMT,\displaystyle=DMT,

where D=diag(sw1,,swn)D=\mbox{diag}(s^{w_{1}},\ldots,s^{w_{n}}) is a diagonal matrix. So by the equivalence relation w\sim_{w} as in Definition 2.1, we get

(t1,t2,,tn)gλi(𝐳)=Mπw(E~(λi))WGr(d,n).(t_{1},t_{2},\dots,t_{n})g_{\lambda^{i}}({\bf z})=M\in\pi_{w}(\widetilde{E}(\lambda^{i}))\subset\mbox{WGr}(d,n).

Let akla_{kl} be the coefficient of zklz^{l}_{k} in the matrix MM for 1ld1\leq l\leq d, 1kλl11\leq k\leq\lambda_{l}-1, kλ1,λ2,,λl1k\neq\lambda_{1},\lambda_{2},\dots,\lambda_{l-1}. Then

akl=swλltkswktλl.a_{kl}=\frac{s^{w_{\lambda_{l}}}t_{k}}{s^{w_{k}}t_{\lambda_{l}}}.

Now for 1kλl11\leq k\leq\lambda_{l}-1, kλ1,λ2,,λl1k\neq\lambda_{1},\lambda_{2},\dots,\lambda_{l-1} we have (λl,k)rev(λi)(\lambda_{l},k)\in\rm{rev}(\lambda^{i}). Let λj=(λl,k)λi\lambda^{j}=(\lambda_{l},k)\lambda^{i}. Note that λj<λi\lambda^{j}<\lambda^{i}. Recall cic_{i} from (2.5). So

tkswλlswktλl=tλjtλiswλlwk=tλjtλiscicj=tλjtλitλicicjci=tλj(tλi)cjci,\frac{t_{k}s^{w_{\lambda_{l}}}}{s^{w_{k}}t_{\lambda_{l}}}=\frac{t_{\lambda^{j}}}{t_{\lambda^{i}}}s^{w_{\lambda_{l}}-w_{k}}=\frac{t_{\lambda^{j}}}{t_{\lambda^{i}}}s^{c_{i}-c_{j}}=\frac{t_{\lambda^{j}}}{t_{\lambda^{i}}}t_{\lambda^{i}}^{\frac{c_{i}-c_{j}}{c_{i}}}=t_{\lambda^{j}}(t_{\lambda^{i}})^{-\frac{c_{j}}{c_{i}}},

since sci=tλ1tλd=tλis^{c_{i}}=t_{\lambda_{1}}\cdots t_{\lambda_{d}}=t_{\lambda^{i}} and tλj=tλ1tλl1tktλl+1tλdt_{\lambda^{j}}=t_{\lambda_{1}}\cdots t_{\lambda_{l-1}}t_{k}t_{\lambda_{l+1}}\cdots t_{\lambda_{d}}. Since WGr(d,n)\mbox{WGr}(d,n) is divisive and λj<λi\lambda^{j}<\lambda^{i}, we have cic_{i} divides cjc_{j}.

Define a ()n(\mathbb{C}^{*})^{n}-action on (λi)\mathbb{C}^{\ell(\lambda^{i})} by

(t1,t2,,tn)(zkl)=(tλj(tλi)cjcizkl)(t_{1},t_{2},\dots,t_{n})(z^{l}_{k})=(t_{\lambda^{j}}(t_{\lambda^{i}})^{-\frac{c_{j}}{c_{i}}}z^{l}_{k})

for 1ld;1kλl1;kλ1,λ2,,λl11\leq l\leq d;1\leq k\leq\lambda_{l}-1;k\neq\lambda_{1},\lambda_{2},\dots,\lambda_{l-1}. With this action of ()n(\mathbb{C}^{*})^{n} on (λi)\mathbb{C}^{\ell(\lambda^{i})}, the map gλig_{\lambda^{i}} becomes ()n(\mathbb{C}^{*})^{n}-equivariant.

If σId\sigma\neq\mbox{Id}, then consider the cell

πw(σE~(λi))σE(λi)G(σci)σE(λi)G(σci/σci)σE(λi), for all i=1,2,,m\pi_{w}(\sigma\widetilde{E}(\lambda^{i}))\cong\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i})}\cong\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i}/\sigma c_{i})}\cong\sigma E(\lambda^{i}),\text{ for all }i=1,2,\dots,m

by Lemma 3.3. Thus we get the map σgλi:(λi)πw(σE~(λi))\sigma g_{\lambda^{i}}\colon\mathbb{C}^{\ell(\lambda^{i})}\to\pi_{w}(\sigma\widetilde{E}(\lambda^{i})) defined by 𝐳σgλi(𝐳){\bf z}\to\sigma g_{\lambda^{i}}({\bf z}). Then by similar arguments, we get the ()n(\mathbb{C}^{*})^{n}-action on (λi)\mathbb{C}^{\ell(\lambda^{i})} defined by

(3.3) (t1,t2,,tn)(zkl)=(tσλj(tσλi)σcjσcizkl).(t_{1},t_{2},\dots,t_{n})(z^{l}_{k})=(t_{\sigma\lambda^{j}}(t_{\sigma\lambda^{i}})^{-\frac{\sigma c_{j}}{\sigma c_{i}}}z^{l}_{k}).

Corollary 3.20.

If WGr(d,n){\rm WGr}(d,n) is divisive, then H(WGr(d,n);)H^{*}({\rm WGr}(d,n);\mathbb{Z}) has no torsion and is concentrated in even degrees.

We remark that Corollary 3.20 also follows from the proof of Theorem 3.10 and Definition 3.14. However, Theorem 3.19 describes the representation of the ()n(\mathbb{C}^{*})^{n}-action on each invariant cell explicitly. We also get that a divisive weighted Grassmann orbifold is integrally equivariantly formal.

4. Equivariant cohomology, cobordism and KK-theory of weighted Grassmann orbifolds

In this section, first we compute the equivariant KK-theory ring of any weighted Grassmann orbifold with rational coefficients. Then we compute the equivariant cohomology ring, equivariant KK-theory ring and equivariant cobordism ring of a divisive weighted Grassmann orbifold with integer coefficients. We discuss the computation of the equivariant Euler classes for some line bundles on a point. We also compute the integral equivariant cohomology ring of some non-divisive weighted Grassmann orbifolds. We adhere the notations of previous sections.

We recall the ()n(\mathbb{C}^{*})^{n}-action on WGr(d,n)\mbox{WGr}(d,n) which is induced by (2.7). Consider the standard torus Tn=(S1)n()nT^{n}=(S^{1})^{n}\subset(\mathbb{C}^{*})^{n}. So we have the restricted TnT^{n}-action on WGr(d,n)\mbox{WGr}(d,n). For each Schubert symbol λ=(λ1,λ2,,λd)\lambda=(\lambda_{1},\lambda_{2},\dots,\lambda_{d}), let C(λ)Md(n,d)C(\lambda)\in M_{d}(n,d) with column vectors given by eλ1,eλ2,,eλde_{\lambda_{1}},e_{\lambda_{2}},\dots,e_{\lambda_{d}} where {e1,e2,,en}\{e_{1},e_{2},\dots,e_{n}\} is the standard basis for n\mathbb{C}^{n}. Therefore [C(λ)]WGr(d,n)[C(\lambda)]\in\mbox{WGr}(d,n) and it is a fixed point of the TnT^{n}-action on WGr(d,n)\mbox{WGr}(d,n).

Proposition 4.1.

Let WGr(d,n){\rm WGr}(d,n) be a weighted Grassmann orbifold corresponding to W=(w1,w2,,wn)(0)nanda1W=(w_{1},w_{2},\dots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n}~{}\mbox{and}~{}a\geq 1. Then there is a ()n(\mathbb{C}^{*})^{n}-invariant stratification

{pt}=X0X1X2Xm=WGr(d,n)\{pt\}=X_{0}\subset X_{1}\subset X_{2}\subset\cdots\subset X_{m}={\rm WGr}(d,n)

such that the quotient Xi/Xi1X_{i}/X_{i-1} is homeomorphic to the Thom space Th(ξi)Th(\xi^{i}) of an orbifold ()n(\mathbb{C}^{*})^{n}-vector bundle

(4.1) ξi:(λi)/G(ci)[C(λi)],\xi^{i}\colon\mathbb{C}^{\ell(\lambda^{i})}/G(c_{i})\to[C(\lambda^{i})],

where G(ci)G(c_{i}) is the cyclic group of the cic_{i}-th roots of unity, for i=1,,mi=1,\dots,m.

Proof.

Recall the ()n(\mathbb{C}^{*})^{n}-invariant stratification

{pt}=X0X1X2Xm=WGr(d,n)\{pt\}=X_{0}\subset X_{1}\subset X_{2}\subset\cdots\subset X_{m}=\mbox{WGr}(d,n)

from (2.16) which is obtained from the qq-cell structure of WGr(d,n)\mbox{WGr}(d,n) as in Lemma 2.6. Note that Xi/Xi1X_{i}/X_{i-1} is the one point compactification of E(λi)G(ci)\frac{E(\lambda^{i})}{G(c_{i})} which is the Thom space of the orbifold ()n(\mathbb{C}^{*})^{n}-vector bundle

E(λi)G(ci)[C(λi)]\frac{E(\lambda^{i})}{G(c_{i})}\to[C(\lambda^{i})]

where [C(λi)][C(\lambda^{i})] is the ()n(\mathbb{C}^{*})^{n}-fixed point corresponding to the Schubert symbol λi\lambda^{i} for i=1,,mi=1,\ldots,m. It remains to note that E(λi)E(\lambda^{i}) is ()n(\mathbb{C}^{*})^{n}-equivariantly homeomorphic to (λi)\mathbb{C}^{\ell(\lambda^{i})}, see (2.13). ∎

Now corresponding to rev(λi)\text{rev}(\lambda^{i}), one can define a subset of Schubert symbols as follows

(4.2) R(λi):={λj|λj=(k,k)λi for (k,k)rev(λi)}.R(\lambda^{i}):=\{\lambda^{j}~{}|~{}\lambda^{j}=(k,k^{\prime})\lambda^{i}\text{ for }(k,k^{\prime})\in\text{rev}(\lambda^{i})\}.

Then the cardinality of the set R(λi)R(\lambda^{i}) is (λi)\ell(\lambda^{i}) for every i{0,1,,m}i\in\{0,1,\dots,m\}. Note that the bundle in (4.1) is also an orbifold TnT^{n}-bundle.

Proposition 4.2.

The orbifold TnT^{n}-bundle in (4.1) has a decomposition

ξi:(λi)G(ci)[C(λi)]j:λjR(λi)(ξij:ijG(cij)[C(λi)]).\xi^{i}\colon\frac{\mathbb{C}^{\ell(\lambda^{i})}}{G(c_{i})}\to[C(\lambda^{i})]\cong\bigoplus_{j:\lambda^{j}\in R(\lambda^{i})}(\xi^{ij}\colon\frac{\mathbb{C}_{ij}}{G(c_{ij})}\to[C(\lambda^{i})]).
Proof.

Observe that XiXi1=E(λi)G(ci)(λi)G(ci)X_{i}\setminus X_{i-1}=\frac{E(\lambda^{i})}{G(c_{i})}\cong\frac{\mathbb{C}^{\ell(\lambda^{i})}}{G(c_{i})}. Since TnT^{n} is abelian, the TnT^{n} action on E(λi)(λi)E(\lambda^{i})\cong\mathbb{C}^{\ell(\lambda^{i})} determines the following decomposition

E(λi)j:λjR(λi)ijE(\lambda^{i})\cong\bigoplus_{j:\lambda^{j}\in R(\lambda^{i})}\mathbb{C}_{ij}

for some irreducible representation ij\mathbb{C}_{ij} of TnT^{n}. By [GGKRW18, Proposition 2.8] there exists a finite covering map q:TnTnq\colon T^{n}\to T^{n} such that the projection map ϕ:E(λi)E(λi)G(ci)\phi\colon{E(\lambda^{i})}\to\frac{E(\lambda^{i})}{G(c_{i})} is equivariant via the map qq (i.e., ϕ(tx)=q(t)ϕ(x)\phi(tx)=q(t)\phi(x)). Therefore,

E(λi)G(ci)j:λjR(λi)ijG(cij)\frac{E(\lambda^{i})}{G(c_{i})}\cong\bigoplus_{j:\lambda^{j}\in R(\lambda^{i})}\frac{\mathbb{C}_{ij}}{G(c_{ij})}

for some positive integers cijc_{ij} which divides cic_{i}. Hence the proof follows. ∎

Remark 4.3.
  1. (1)

    The attaching map ηi:S(ξi)Xi1\eta_{i}\colon S(\xi^{i})\to X_{i-1} for the qq-cell structure in (2.16) satisfies ηi|S(ξij)=fij.ξij\eta_{i}|_{S(\xi^{ij})}=f_{ij}.\xi^{ij} where fij:[C(λi)][C(λj)]f_{ij}\colon[C(\lambda^{i})]\to[C(\lambda^{j})] is the constant map.

  2. (2)

    The equivariant Euler classes {eTn(ξij)|j<i}\{e_{T^{n}}(\xi^{ij})~{}|~{}j<i\} are non zero divisors. They are pairwise prime by [HHH05, Lemma 5.2] and the TnT^{n}-action on E(λi)E(\lambda^{i}) discussed in the proof of Theorem 3.19.

Theorem 4.4.

Let WGr(d,n){\rm WGr}(d,n) be a weighted Grassmann orbifold for d<nd<n corresponding to W=(w1,w2,,wn)(0)nanda1W=(w_{1},w_{2},\dots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n}~{}\mbox{and}~{}a\geq 1. Then the generalized TnT^{n}-equivariant cohomology Tn(WGr(d,n);)\mathcal{E}^{\ast}_{T^{n}}({\rm WGr}(d,n);\mathbb{Q}) can be given by

{(fi)i=0mTn(pt;)|eTn(ξij)dividesfifjfor j<iand|λjλi|=d1}\Big{\{}(f_{i})\in\bigoplus_{i=0}^{m}\mathcal{E}^{*}_{T^{n}}(pt;\mathbb{Q})~{}\big{|}~{}e_{T^{n}}(\xi^{ij})~{}\mbox{divides}~{}f_{i}-f_{j}~{}\mbox{for }~{}j<i~{}\mbox{and}~{}|\lambda^{j}\cap\lambda^{i}|=d-1\Big{\}}

for Tn=KTn\mathcal{E}^{\ast}_{T^{n}}=K^{*}_{T^{n}}, HTnH^{*}_{T^{n}}, where eTn(ξij)e_{T^{n}}(\xi^{ij}) represents the equivariant Euler class of ξij\xi^{ij}.

Proof.

This follows from [SS21, Proposition 2.3] using Proposition 4.1, 4.2 and Remark 4.3. ∎

We note that equivariant cohomology ring of WGr(d,n)\mbox{WGr}(d,n) with rational coefficients is discussed in [AM15]. In the rest, we give a description of the equivariant cohomology ring, equivariant KK-theory ring and equivariant cobordism ring of a divisive weighted Grassmann orbifold with integer coefficients.

Proposition 4.5.

Let WGr(d,n){\rm WGr}(d,n) be a divisive weighted Grassmann orbifold for d<nd<n corresponding to W=(w1,w2,,wn)(0)nW=(w_{1},w_{2},\dots,w_{n})\in(\mathbb{Z}_{\geq 0})^{n} and a1a\geq 1. Then there is a TnT^{n}-invariant stratification

{pt}=X0X1Xm=WGr(d,n)\{pt\}=X_{0}\subset X_{1}\subset\cdots\subset X_{m}={\rm WGr}(d,n)

such that the quotient Xi/Xi1X_{i}/X_{i-1} is homeomorphic to the Thom space Th(ξi)Th(\xi^{i}) of the TnT^{n}-vector bundle

ξi:(λi)[C(λi)],\xi^{i}\colon\mathbb{C}^{\ell(\lambda^{i})}\to[C(\lambda^{i})],

for i=1,,mi=1,\dots,m.

Proof.

Since WGr(d,n)\mbox{WGr}(d,n) is divisive, there exists σSn\sigma\in S_{n} such that σci\sigma c_{i} divides σci1\sigma c_{i-1} for i=1,2,,mi=1,2,\dots,m. Then gcd{σc0,σc1,,σci}=σci\gcd\{\sigma c_{0},\sigma c_{1},\dots,\sigma c_{i}\}=\sigma c_{i} for all ii. By Theorem 3.5, one can write WGr(d,n)=i=0mσE(λi)G(σci)\mbox{WGr}(d,n)=\sqcup_{i=0}^{m}\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i})}. By Lemma 3.3, the qq-cell σE(λi)/G(σci)\sigma E(\lambda^{i})/G(\sigma c_{i}) is homeomorphic to σE(λi)/G(σciσci)(λi)\sigma E(\lambda^{i})/G(\frac{\sigma c_{i}}{\sigma c_{i}})\cong\mathbb{C}^{\ell(\lambda^{i})} for i=1,,mi=1,\ldots,m. Let Xk=i=0kσE(λi)G(σci)X_{k}=\sqcup_{i=0}^{k}\frac{\sigma E(\lambda^{i})}{G(\sigma c_{i})} for i=0,1,,mi=0,1,\ldots,m. Remaining follows from the proof of Proposition 4.1. ∎

Remark 4.6.

For a divisive weighted Grassmann orbifold, Proposition 4.2 and Remark 4.3 hold with cij=1c_{ij}=1 for every j<ij<i.

Theorem 4.7.

Let WGr(d,n){\rm WGr}(d,n) be a divisive weighted Grassmann orbifold for d<nd<n. Then the generalized TnT^{n}-equivariant cohomology Tn(WGr(d,n);)\mathcal{E}^{\ast}_{T^{n}}({\rm WGr}(d,n);\mathbb{Z}) can be given by

{(fi)i=0mTn(pt;)|eTn(ξij)dividesfifjfor j<iand|λjλi|=d1}\Big{\{}(f_{i})\in\bigoplus_{i=0}^{m}\mathcal{E}^{*}_{T^{n}}(pt;\mathbb{Z})~{}\big{|}~{}e_{T^{n}}(\xi^{ij})~{}\mbox{divides}~{}f_{i}-f_{j}~{}\mbox{for }~{}j<i~{}\mbox{and}~{}|\lambda^{j}\cap\lambda^{i}|=d-1\Big{\}}

for Tn=HTn,KTn\mathcal{E}^{\ast}_{T^{n}}=H^{\ast}_{T^{n}},K^{*}_{T^{n}} and MUTnMU^{*}_{T^{n}}.

Proof.

This follows from Proposition 4.5, Remark 4.6 and [HHH05, Theorem 2.3]. ∎

Remark 4.8.

Let λi\lambda^{i} and λj\lambda^{j} be two Schubert symbols with j<ij<i. If WGr(d,n){\rm WGr}(d,n) is a divisive weighted Grassmann orbifold then there exists a permutation σSn\sigma\in S_{n} such that σci\sigma c_{i} divides σcj\sigma c_{j}. We denote σdij:=σcjσci\sigma d_{ij}:=\frac{\sigma c_{j}}{\sigma c_{i}}\in\mathbb{Z}.

Next we discuss how to compute eTn(ξij)e_{T^{n}}(\xi^{ij}). We recall that

HTn(pt;)=H(BTn;)[y1,y2,,yn]H_{T^{n}}^{*}(pt;\mathbb{Z})=H^{*}(BT^{n};\mathbb{Z})\cong\mathbb{Z}[y_{1},y_{2},\dots,y_{n}]

where y1,y2,,yny_{1},y_{2},\dots,y_{n} be the standard basis of H2(BTn;)H^{2}(BT^{n};\mathbb{Z}). Using (3.3) the character of the one-dimensional representation for the bundle ξij\xi^{ij} is given by

(4.3) (t1,t2,,tn)tσλj(tσλi)σcjσci.(t_{1},t_{2},\dots,t_{n})\to t_{\sigma\lambda^{j}}(t_{\sigma\lambda^{i}})^{-\frac{\sigma c_{j}}{\sigma c_{i}}}.

Also

KTn(pt)R(Tn)[z,z1]K^{*}_{T^{n}}(pt)\cong R(T^{n})[z,z^{-1}]

where R(Tn)R(T^{n}) is the complex representation ring of TnT^{n} and zz is the Bott element in K2(pt)K^{-2}(pt). Note that R(Tn)R(T^{n}) is isomorphic to the ring of Laurent polynomials with nn-variables, that is R(Tn)[α1,,αn](α1αn)R(T^{n})\cong\mathbb{Z}[\alpha_{1},\ldots,\alpha_{n}]_{(\alpha_{1}\cdots\alpha_{n})}, where αi\alpha_{i} is the irreducible representation corresponding to the projection on the ii-th factor, see [Hus94]. Therefore, using (4.3) one has the following.

(4.4) eTn(ξij)={1ασλjασλiσdijinKTn0eTn(ασλjασλiσdij)inMUTn2YσλjσdijYσλiinHTn2\displaystyle e_{T^{n}}(\xi^{ij})=\begin{cases}1-\alpha_{\sigma\lambda^{j}}\alpha_{\sigma\lambda^{i}}^{-\sigma d_{ij}}&\quad\text{in}~{}K^{0}_{T^{n}}\\ e_{T^{n}}(\alpha_{\sigma\lambda^{j}}\alpha_{\sigma\lambda^{i}}^{-\sigma d_{ij}})&\quad\text{in}~{}MU^{2}_{T^{n}}\\ Y_{\sigma\lambda^{j}}-\sigma d_{ij}Y_{\sigma\lambda^{i}}&\quad\text{in}~{}H^{2}_{T^{n}}\end{cases}

for j<ij<i and |λjλi|=d1|\lambda^{j}\cap\lambda^{i}|=d-1, where Yλ:=i=1dyλiY_{\lambda}:=\sum_{i=1}^{d}y_{\lambda_{i}} and αλ=αλ1αλd\alpha_{\lambda}=\alpha_{\lambda_{1}}\cdots\alpha_{\lambda_{d}} for a Schubert symbol λ=(λ1,,λd)\lambda=(\lambda_{1},\ldots,\lambda_{d}).

We remark that the structure of MUTn(pt)MU_{T^{n}}^{\ast}(pt) is unknown, however it is referred as the ring of TnT^{n}-cobordism forms in [HHRW16].

Example 4.9.

Consider the weighted Grassmann orbifold WGr(2,4)\mbox{WGr}(2,4) for W=(12,2,2,2)W=(12,2,2,2) and a=6a=6. We have the ordering on the 66 Schubert symbols given by

λ0=(1,2)<λ1=(1,3)<λ2=(1,4)<λ3=(2,3)<λ4=(2,4)<λ5=(3,4).\lambda^{0}=(1,2)<\lambda^{1}=(1,3)<\lambda^{2}=(1,4)<\lambda^{3}=(2,3)<\lambda^{4}=(2,4)<\lambda^{5}=(3,4).

Then c0=20,c1=20,c2=20,c3=10,c4=10,c5=10c_{0}=20,c_{1}=20,c_{2}=20,c_{3}=10,c_{4}=10,c_{5}=10 from (2.5). Here cic_{i} divides ci1c_{i-1} for all i=1,2,3,4,5i=1,2,3,4,5. Thus WGr(2,4)\mbox{WGr}(2,4) is divisive for the identity permutation in S4S_{4}. Then dij=cjcid_{ij}=\frac{c_{j}}{c_{i}} in Remark 4.8 gives

dij={1if j<i and bothi,j{0,1,2}or,{3,4,5}2ifj{0,1,2}andi{3,4,5}.\displaystyle d_{ij}=\begin{cases}1&\quad\text{if }j<i\text{ and both}~{}i,j\in\{0,1,2\}~{}\mbox{or},~{}\{3,4,5\}\\ 2&\quad\mbox{if}~{}j\in\{0,1,2\}~{}\mbox{and}~{}i\in\{3,4,5\}.\end{cases}

Then one can calculate the equivariant Euler class eTn(ξij)e_{T^{n}}(\xi^{ij}) from (4.4). The generalized integral equivariant cohomology ring Tn(WGr(2,4);)\mathcal{E}_{T^{n}}^{*}(\mbox{WGr}(2,4);\mathbb{Z}) of this divisive weighted Grassmann orbifold WGr(2,4)\mbox{WGr}(2,4) can be described by Theorem 4.7. ∎

The following result gives equivariant cohomology ring of some non-divisive weighted Grassmann orbifolds with integer coefficients.

Theorem 4.10.

Let WGr(d,n){\rm WGr}(d,n) be a weighted Grassmann orbifold corresponding to the order λ0<<λm\lambda^{0}<\cdots<\lambda^{m} such that ci|ckc_{i}|c_{k} for kik\leq i and i2i\geq 2 but c1c_{1} does not divide c0c_{0}. Then the integral equivariant cohomology ring of WGr(d,n){\rm WGr}(d,n) is given by

HTn(WGr(d,n);)\displaystyle H_{T^{n}}^{*}({\rm WGr}(d,n);\mathbb{Z})
={(fi)i=0m[y1,y2,,yn]|(YλjdijYλi)divides(fifj) if j<i,\displaystyle=\{(f_{i})\in\bigoplus_{i=0}^{m}\mathbb{Z}[y_{1},y_{2},\dots,y_{n}]~{}\big{|}~{}(Y_{\lambda^{j}}-d_{ij}Y_{\lambda^{i}})~{}\mbox{divides}~{}(f_{i}-f_{j})\text{ if }j<i,
|λjλi|=d1;(i,j)(0,1)andc1Yλ0c0Yλ1divides(f1f0)}.\displaystyle|\lambda^{j}\cap\lambda^{i}|=d-1;(i,j)\neq(0,1)~{}\mbox{and}~{}c_{1}Y_{\lambda^{0}}-c_{0}Y_{\lambda^{1}}~{}\text{divides}~{}(f_{1}-f_{0})\}.
Proof.

By the given condition gcd{c0,c1,,ci}=ci\gcd\{c_{0},c_{1},\dots,c_{i}\}=c_{i} for i2i\geq 2. So, by Lemma 3.3, we have E(λi)/G(ci)E(\lambda^{i})/G(c_{i}) is homeomorphic to E(λi)/G(ci/ci)(λi)E(\lambda^{i})/G(c_{i}/c_{i})\cong\mathbb{C}^{\ell(\lambda^{i})} for i=1,,mi=1,\ldots,m. When i=1,i=1, we have X1X_{1} is equivariantly homeomorphic to 𝕎P(c0,c1)\mathbb{W}P(c_{0},c_{1}). Therefore it has a TnT^{n}-invariant cell structure. Thus by [DKS22, Theorem 2.9], we get the result. ∎

Next we discuss the equivariant cohomology ring of the weighted projective space 𝕎P(b0,b1,bm)\mathbb{W}P(b_{0},b_{1}\ldots,b_{m}) where (b0,b1,bm)(0)m+1(b_{0},b_{1}\ldots,b_{m})\in(\mathbb{Z}_{\geq 0})^{m+1} for several torus actions. By remark 2.3, 𝕎P(b0,b1,bm)=WGr(1,m+1)\mathbb{W}P(b_{0},b_{1}\ldots,b_{m})=\mbox{WGr}(1,m+1) associated to the weight W=(b01,,bm1)W=(b_{0}-1,\dots,b_{m}-1) and a=1a=1. The Schubert symbols for 1<m+11<m+1 are {1},,{m} and {m+1}\{1\},\ldots,\{m\}\text{ and }\{m+1\}. Assume that WGr(1,m+1)\mbox{WGr}(1,m+1) is divisive corresponding to this order (i.e bib_{i} divides bi1b_{i-1}, for i=1,2,,mi=1,2,\dots,m). Then

E(i+1){[(u0,u1,,ui1,1,0,,0)]𝕎P(b0,b1,bm)}iE(i+1)\cong\{[(u_{0},u_{1},\dots,u_{i-1},1,0,\dots,0)]\in\mathbb{W}P(b_{0},b_{1}\ldots,b_{m})\}\cong\mathbb{C}^{i}

for i=0,1,,mi=0,1,\dots,m.

Let (n,d)(n,d) be the pair such that d<nd<n and (nd)=m+1{n\choose{d}}=m+1. Then (2.8) gives a TnT^{n}- action on 𝕎P(b0,b1,bm)\mathbb{W}P(b_{0},b_{1}\ldots,b_{m}). Recall tλit_{\lambda^{i}} from (2.8) for the Schubert symbols λ0,λ1,,λm\lambda^{0},\lambda^{1},\dots,\lambda^{m} corresponding to d<nd<n. Therefore we have the following,

(t1,t2,,tn)[(u0,u1,,ui1,1,0,,0)]\displaystyle(t_{1},t_{2},\dots,t_{n})[(u_{0},u_{1},\dots,u_{i-1},1,0,\dots,0)]
=[(tλ0u0,tλ1u1,,tλi1ui1,tλi,0,,0)]\displaystyle=[(t_{\lambda^{0}}u_{0},t_{\lambda^{1}}u_{1},\dots,t_{\lambda^{i-1}}u_{i-1},{t_{\lambda^{i}}},0,\dots,0)]
=[((tλi)b0bitλ0u0,(tλi)b1bitλ1u1,,(tλi)bi1bitλi1ui1,1,0,,0)].\displaystyle=[((t_{\lambda^{i}})^{-\dfrac{b_{0}}{b_{i}}}t_{\lambda^{0}}u_{0},(t_{\lambda^{i}})^{-\dfrac{b_{1}}{b_{i}}}t_{\lambda^{1}}u_{1},\dots,(t_{\lambda^{i}})^{-\dfrac{b_{i-1}}{b_{i}}}t_{\lambda^{i-1}}u_{i-1},1,0,\dots,0)].

Then E(i+1)E(i+1) is TnT^{n}-invariant as well as Tm+1T^{m+1}-invariant. Let

Xi:=[(u0,u1,,ui,0,,0)]𝕎P(b0,b1,bm)}.X_{i}:=[(u_{0},u_{1},\dots,{u_{i}},0,\dots,0)]\in\mathbb{W}P(b_{0},b_{1}\ldots,b_{m})\}.

Then XiX_{i} gives a filtration

(4.5) {pt}=X0X1Xm=𝕎P(b0,b1,,bm).\{pt\}=X_{0}\subset X_{1}\subset\cdots\subset X_{m}=\mathbb{W}P(b_{0},b_{1},\dots,b_{m}).

Note that the filtration in (4.5) satisfies Proposition 4.5 and Remark 4.6. Thus in this case

ξi:E(i+1)[ei+1]j=0i(ξij:ij[ei+1])\xi^{i}\colon E(i+1)\to[e_{i+1}]\cong\bigoplus_{j=0}^{i}(\xi^{ij}:\mathbb{C}_{ij}\to[e_{i+1}])

for some irreducible representation ij\mathbb{C}_{ij}. Then one can get the following result using the proof of [HHRW16, Theorem 2.3].

Theorem 4.11.

If 𝕎P(b0,,bm)\mathbb{W}P(b_{0},\ldots,b_{m}) is divisive, then the generalized TnT^{n}-equivariant cohomology Tn(𝕎P(b0,,bm);)\mathcal{E}^{\ast}_{T^{n}}(\mathbb{W}P(b_{0},\ldots,b_{m});\mathbb{Z}) can be given by

{(fi)i=0mTn(pt;)|eTn(ξij)dividesfifjfor all j<i}\Big{\{}(f_{i})\in\bigoplus_{i=0}^{m}\mathcal{E}^{*}_{T^{n}}(pt;\mathbb{Z})~{}\big{|}~{}e_{T^{n}}(\xi^{ij})~{}\mbox{divides}~{}f_{i}-f_{j}~{}\mbox{for all }~{}j<i\Big{\}}

for Tn=HTn,KTn\mathcal{E}^{\ast}_{T^{n}}=H^{\ast}_{T^{n}},K^{*}_{T^{n}} and MUTnMU^{*}_{T^{n}}.

We note that there are several pairs (n,d)(n,d) such that d<nd<n and (nd)=m+1>2{n\choose d}=m+1>2. Now we discuss how to calculate the equivariant Euler class eTn(ξij)e_{T^{n}}(\xi^{ij}) in Theorem 4.11. The corresponding one dimensional representation on the bundle ξij\xi^{ij} for j<ij<i is determined by the character

(t1,,tn)(tλi)bjbitλj.(t_{1},\ldots,t_{n})\to(t_{\lambda^{i}})^{-\dfrac{b_{j}}{b_{i}}}t_{\lambda^{j}}.

Thus, similar to, (4.4) one can calculate the equivariant Euler class eTn(ξij)e_{T^{n}}(\xi^{ij}) of the bundle ξij\xi^{ij} for j<ij<i.

Example 4.12.

For m=2m=2, we have (31)=(32)=3{3\choose 1}={3\choose 2}=3. Thus corresponding to two different pairs (3,1)(3,1) and (3,2)(3,2) we have two different T3T^{3} action on 𝕎P(b0,b1,b2)\mathbb{W}P(b_{0},b_{1},b_{2}). The map f:T3T3f\colon T^{3}\to T^{3} defined by (t1,t2,t3)(t1t2,t1t3,t2t3)(t_{1},t_{2},t_{3})\to(t_{1}t_{2},t_{1}t_{3},t_{2}t_{3}) is not an automorphism. So these actions are not equivalent. However, using Theorem 4.11, one can calculate the equivariant cohomology of 𝕎P(b0,b1,b2)\mathbb{W}P(b_{0},b_{1},b_{2}) for both the actions if bib_{i} divides bi1b_{i-1} for i=1,2i=1,2. ∎

5. Equivariant Schubert calculus for divisive weighted Grassmann orbifolds

In this section, we show that there exist equivariant Schubert classes which form a basis for the equivariant cohomology ring of a divisive weighted Grassmann orbifold with integer coefficients. Moreover, we compute the weighted structure constants corresponding to this equivariant Schubert basis with integer coefficients.

For xHTn(WGr(d,n);)\text{x}\in H_{T^{n}}^{*}(\mbox{WGr}(d,n);\mathbb{Z}), the support of x denoted by ‘supp(x)\rm{supp}(x)’ is the set of all Schubert symbols λi\lambda^{i} such that x|λi0x|_{\lambda^{i}}\neq 0. Recall the partial order ‘\preceq’ on the Schubert symbols defined in (2.1). We follow this partial order ‘\preceq’ and we call an element xHTn(WGr(d,n);)x\in H_{T^{n}}^{*}(\mbox{WGr}(d,n);\mathbb{Z}) is supported above by λi\lambda^{i} if λiλk\lambda^{i}\preceq\lambda^{k} for all λksupp(x)\lambda^{k}\in\rm{supp}(x).

Let WGr(d,n)\mbox{WGr}(d,n) be a divisive weighted Grassmann orbifold. Then there exists σSn\sigma\in S_{n} such that

(5.1) σcidivides σci1fori=1,2,,m.\sigma c_{i}~{}\text{divides }\sigma c_{i-1}~{}\text{for}~{}i=1,2,\dots,m.

Using Theorem 3.5, it is sufficient to consider σ=Id\sigma=\rm{Id} (the identity permutation on SnS_{n}). For σ=Id\sigma=\rm{Id}, (5.1) transforms to

cidivides ci1fori=1,2,,m.c_{i}~{}\text{divides }c_{i-1}~{}\text{for}~{}i=1,2,\dots,m.

Recall the definition of R(λi)R(\lambda^{i}) from (4.2). We introduce the following definition.

Definition 5.1.

An element xHTn(WGr(d,n);)x\in H_{T^{n}}^{*}({\rm WGr}(d,n);\mathbb{Z}) is said to be an equivariant Schubert class corresponding to a Schubert symbol λi\lambda^{i} if the following conditions are satisfied.

  1. (1)

    x|λk0λiλkx|_{\lambda^{k}}\neq 0\implies\lambda^{i}\preceq\lambda^{k} (say that xx is supported above λi\lambda^{i}).

  2. (2)

    x|λi=λjR(λi)(YλjcjciYλi)x|_{\lambda^{i}}=\prod_{\lambda^{j}\in R(\lambda^{i})}(Y_{\lambda^{j}}-\dfrac{c_{j}}{c_{i}}Y_{\lambda^{i}}).

  3. (3)

    x|λkx|_{\lambda^{k}} is a homogeneous polynomial of y1,y2,,yny_{1},y_{2},\dots,y_{n} of degree (λi)\ell({\lambda^{i}}).

Proposition 5.2 (Uniqueness).

For each Schubert symbol λi\lambda^{i}, there is at most one equivariant Schubert class xx corresponding to λi\lambda^{i}.

Proof.

Suppose that there were two distinct equivariant Schubert classes x,xx,x^{{}^{\prime}} corresponding to λi\lambda^{i}. Let λj\lambda^{j} be the minimal Schubert symbol such that (xx)|λj0{(x-x^{{}^{\prime}})|_{\lambda^{j}}\neq 0}. By Definition 5.1 (1) and (2), we get λiλj\lambda^{i}\prec\lambda^{j}. Then from the condition in the expression of the equivariant cohomology ring in Theorem 4.7, we get that (xx)|λj(x-x^{{}^{\prime}})|_{\lambda^{j}} is a multiple of λkR(λj)(YλkckcjYλj)\prod_{\lambda^{k}\in R(\lambda^{j})}(Y_{\lambda^{k}}-\dfrac{c_{k}}{c_{j}}Y_{\lambda^{j}}) which is of degree (λj)\ell(\lambda^{j}). This contradicts the fact that xxx-x^{{}^{\prime}} is homogeneous of degree (λi)<(λj)\ell(\lambda^{i})<\ell(\lambda^{j}). ∎

Let us denote the equivariant Schubert class corresponding to the Schubert symbol λi\lambda^{i} by wS~λiw\widetilde{S}_{\lambda^{i}} for i=0,1,,mi=0,1,\dots,m. We remark that the existence of wS~λiw\widetilde{S}_{\lambda^{i}} follows from [HHH05, Proposition 4.3] and Theorem 4.7. Using the arguments in the proof of [KT03, Proposition 1], one gets the following.

Proposition 5.3.

The equivariant Schubert classes {wS~λi}i=0m\{w\widetilde{S}_{\lambda^{i}}\}_{i=0}^{m} is a basis for the module HTn(WGr(d,n);)H^{*}_{T^{n}}({\rm WGr}(d,n);\mathbb{Z}) over HTn(pt;)H_{T^{n}}^{*}(pt;\mathbb{Z}). Moreover, any xHTn(WGr(d,n);)x\in H_{T^{n}}^{*}({\rm WGr}(d,n);\mathbb{Z}) can be written uniquely as an HTn(pt;)H_{T^{n}}^{*}(pt;\mathbb{Z}) linear combination of wS~λiw\widetilde{S}_{\lambda^{i}} using only those λi\lambda^{i} such that λjλi\lambda^{j}\preceq\lambda^{i} for some λjsupp(x)\lambda^{j}\in\mbox{supp}(x).

Example 5.4.

In Figure 1, we compute the equivariant Schubert class wS~(2,3)HT4(WGr(2,4);)w\widetilde{S}_{(2,3)}\in H_{T^{4}}^{*}(\mbox{WGr}(2,4);\mathbb{Z}) where WGr(2,4)\mbox{WGr}(2,4) is a divisive weighted Grassmann orbifold for some W=(α+γβ,α,α,α)(0)4W=(\alpha+\gamma\beta,\alpha,\alpha,\alpha)\in(\mathbb{Z}_{\geq 0})^{4} and a=β2α>0a=\beta-2\alpha\in\mathbb{Z}_{>0}. Figure 1(a) is the lattice of the Schubert symbols for 2<42<4. Figure 1(b) gives the equivariant Schubert class corresponding to the Schubert symbol (2,3)(2,3).

(1,3)(1,3)(1,2)(1,2)(a)(a)(1,4)(1,4)(2,3)(2,3)(2,4)(2,4)(3,4)(3,4)00(b)(b)0(Y(1,3)(γ+1)Y(2,3))(Y(1,2)(γ+1)Y(2,3))(Y_{(1,3)}-(\gamma+1)Y_{(2,3)})(Y_{(1,2)}-(\gamma+1)Y_{(2,3)})(Y(1,4)(γ+1)Y(2,4))(Y(1,2)(γ+1)Y(2,4))(Y_{(1,4)}-(\gamma+1)Y_{(2,4)})(Y_{(1,2)}-(\gamma+1)Y_{(2,4)})(Y(1,4)(γ+1)Y(3,4))(Y(1,3)(γ+1)Y(3,4))(Y_{(1,4)}-(\gamma+1)Y_{(3,4)})(Y_{(1,3)}-(\gamma+1)Y_{(3,4)})
Figure 1.

In the rest of this section, we compute the weighted structure constants for the equivariant cohomology of a divisive weighted Grassmann orbifold. Since the set {wS~λi}i=0m\{w\widetilde{S}_{\lambda^{i}}\}_{i=0}^{m} form a HTn({pt};)H_{T^{n}}^{*}(\{pt\};\mathbb{Z})-basis for HTn(WGr(d,n);)H_{T^{n}}^{*}(\mbox{WGr}(d,n);\mathbb{Z}), for any two λi\lambda^{i} and λj\lambda^{j}, one has the following

(5.2) wS~λiwS~λj=λkwcijkwS~λkw\widetilde{S}_{\lambda^{i}}~{}w\widetilde{S}_{\lambda^{j}}=\sum_{\lambda^{k}}wc_{ij}^{k}~{}w\widetilde{S}_{\lambda^{k}}

where λk{λ0,λ1,,λm}\lambda^{k}\in\{\lambda^{0},\lambda^{1},\dots,\lambda^{m}\}. The constant wcijkHTn(pt;)wc_{ij}^{k}\in H_{T^{n}}^{*}(pt;\mathbb{Z}) in the formula is called ‘weighted structure constant’.

Lemma 5.5.

The weighted structure constant wcijkwc_{ij}^{k} have the following properties.

  1. (1)

    The weighted structure constant wcijkwc_{ij}^{k} has degree (λi)+(λj)(λk)\ell(\lambda^{i})+\ell(\lambda^{j})-\ell(\lambda^{k}).

  2. (2)

    wcijk=0wc_{ij}^{k}=0 unless (λk)(λi)+(λj)\ell(\lambda^{k})\leq\ell(\lambda^{i})+\ell(\lambda^{j}) and λkλi,λj\lambda^{k}\succeq\lambda^{i},\lambda^{j}.

  3. (3)

    When i=ki=k we have wciji=wS~λj|λiwc_{ij}^{i}={w\widetilde{S}_{\lambda^{j}}}|_{\lambda^{i}}.

Proof.

(1) The degree of wS~λiw\widetilde{S}_{\lambda^{i}} is (λi)\ell(\lambda^{i}). So the degree of the weighted structure constant wcijkwc_{ij}^{k} is given by

deg(wcijk)\displaystyle\text{deg}(wc_{ij}^{k}) =deg(wS~λi)+deg(wS~λj)deg(wS~λk)\displaystyle=\text{deg}(w\widetilde{S}_{\lambda^{i}})+\text{deg}(w\widetilde{S}_{\lambda^{j}})-\text{deg}(w\widetilde{S}_{\lambda^{k}})
=(λi)+(λj)(λk).\displaystyle=\ell(\lambda^{i})+\ell(\lambda^{j})-\ell(\lambda^{k}).

(2) The weighted structure constant wcijk=0wc_{ij}^{k}=0 if (λi)+(λj)(λk)<0\ell(\lambda^{i})+\ell(\lambda^{j})-\ell(\lambda^{k})<0. Also

(wS~λiwS~λj)|λs0λsλi,λj.{(w\widetilde{S}_{\lambda^{i}}~{}w\widetilde{S}_{\lambda^{j}})}|_{\lambda^{s}}\neq 0\implies\lambda^{s}\succeq\lambda^{i},\lambda^{j}.

Thus by Proposition 5.3, wcijk0λkλi,λjwc_{ij}^{k}\neq 0\implies\lambda^{k}\succeq\lambda^{i},\lambda^{j}.

(3) Compare the λi\lambda^{i}-th component of the both side in (5.2) we get

wS~λi|λiwS~λj|λi=wcijiwS~λi|λi+kiwcijkwS~λk|λi.{w\widetilde{S}_{\lambda^{i}}}|_{\lambda^{i}}~{}{w\widetilde{S}_{\lambda^{j}}}|_{\lambda^{i}}=wc_{ij}^{i}~{}{w\widetilde{S}_{\lambda^{i}}}|_{\lambda^{i}}+\sum_{k\neq i}wc_{ij}^{k}~{}{w\widetilde{S}_{\lambda^{k}}}|_{\lambda^{i}}.

Since, wcijk=0wc_{ij}^{k}=0 unless λkλi\lambda^{k}\succeq\lambda^{i}. But wS~λk|λi=0{w\widetilde{S}_{\lambda^{k}}}|_{\lambda^{i}}=0 for λkλi\lambda^{k}\succeq\lambda^{i}, and λkλi\lambda^{k}\neq\lambda^{i}. Thus all the terms in the summation vanish. So the claim follows, since wS~λi|λi0{w\widetilde{S}_{\lambda^{i}}}|_{\lambda^{i}}\neq 0. ∎

Now we introduce equivariant Schubert divisor class. Note that (λi)=0\ell(\lambda^{i})=0 if and only if i=0i=0 and (λi)=1\ell(\lambda^{i})=1 if and only if i=1i=1.

Lemma 5.6.

The equivariant Schubert divisor class wS~λ1HTn(WGr(d,n);)w\widetilde{S}_{\lambda^{1}}\in H_{T^{n}}^{*}({\rm WGr}(d,n);\mathbb{Z}) is given by

wS~λ1|λi=Yλ0c0ciYλi.{w\widetilde{S}_{\lambda^{1}}}{|_{\lambda^{i}}}=Y_{\lambda^{0}}-\dfrac{c_{0}}{c_{i}}Y_{\lambda^{i}}.
Proof.

Note that wS~λ1|λ1=Yλ0c0c1Yλ1{w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{1}}=Y_{\lambda^{0}}-\dfrac{c_{0}}{c_{1}}Y_{\lambda^{1}}. For other Schubert symbol λi\lambda^{i}, it follows from Definition 5.1. ∎

Let λi\lambda^{i} and λj\lambda^{j} be two Schubert symbols such that λjλi\lambda^{j}\leq\lambda^{i}. Then Lemma 5.6 gives

wS~λ1|λiwS~λ1|λj=c0cj(YλjcjciYλi).{w\widetilde{S}_{\lambda^{1}}}{|_{\lambda^{i}}}-{w\widetilde{S}_{\lambda^{1}}}{|_{\lambda^{j}}}=\dfrac{c_{0}}{c_{j}}(Y_{\lambda^{j}}-\dfrac{c_{j}}{c_{i}}Y_{\lambda^{i}}).

For any two Schubert symbol λi\lambda^{i} and λj\lambda^{j} we denote λiλj\lambda^{i}\to\lambda^{j} if (λi)=(λj)+1\ell(\lambda^{i})=\ell(\lambda^{j})+1 and λjλi\lambda^{j}\preceq\lambda^{i}.

Proposition 5.7 (Weighted Pieri rule).
wS~λ1wS~λj=(wS~λ1|λj)wS~λj+λiλjc0cjwS~λi.w\widetilde{S}_{\lambda^{1}}~{}w\widetilde{S}_{\lambda^{j}}=({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{j}})~{}w\widetilde{S}_{\lambda^{j}}+\sum_{\lambda^{i}\to\lambda^{j}}\dfrac{c_{0}}{c_{j}}~{}w\widetilde{S}_{\lambda^{i}}.
Proof.

Using the fact that deg(wS~λ1)=1(w\widetilde{S}_{\lambda^{1}})=1, we have

wS~λ1wS~λj=(wc1jj)wS~λj+λiλj(wc1ji)wS~λi.w\widetilde{S}_{\lambda^{1}}~{}w\widetilde{S}_{\lambda^{j}}=(wc^{j}_{1j})~{}w\widetilde{S}_{\lambda^{j}}+\sum_{\lambda^{i}\to\lambda^{j}}(wc^{i}_{1j})~{}w\widetilde{S}_{\lambda^{i}}.

From Lemma 5.5, we get wc1jj=wS~λ1|λjwc^{j}_{1j}={w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{j}}. Fix λi\lambda^{i} such that λiλj\lambda^{i}\to\lambda^{j} and compare λi\lambda^{i}-th component of both side we get

wS~λ1|λiwS~λj|λi\displaystyle{w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}}~{}{w\widetilde{S}_{\lambda^{j}}}|_{\lambda^{i}} =(wc1jj)wS~λj|λi+(wc1ji)wS~λi|λi\displaystyle=(wc^{j}_{1j})~{}{w\widetilde{S}_{\lambda^{j}}}|_{\lambda^{i}}+(wc^{i}_{1j})~{}{w\widetilde{S}_{\lambda^{i}}}|_{\lambda^{i}}
(wc1ji)wS~λi|λi\displaystyle\implies(wc^{i}_{1j})~{}{w\widetilde{S}_{\lambda^{i}}}|_{\lambda^{i}} =(wS~λ1|λiwS~λ1|λj)wS~λj|λi\displaystyle=({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}}-{w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{j}})~{}{w\widetilde{S}_{\lambda^{j}}}|_{\lambda^{i}}
(wc1ji)wS~λi|λi\displaystyle\implies(wc^{i}_{1j})~{}{w\widetilde{S}_{\lambda^{i}}}|_{\lambda^{i}} =c0cj(YλjcjciYλi)wS~λj|λi.\displaystyle=\dfrac{c_{0}}{c_{j}}~{}(Y_{\lambda^{j}}-\dfrac{c_{j}}{c_{i}}Y_{\lambda^{i}})~{}{w\widetilde{S}_{\lambda^{j}}}|_{\lambda^{i}}.

Thus wc1ji=c0cjwc^{i}_{1j}=\dfrac{c_{0}}{c_{j}}, if λiλj\lambda^{i}\to\lambda^{j}. Hence we get the proof. ∎

By applying Proposition 5.7 repeatedly we can compute the following as well as the higher products.

(wS~λ1)2wS~λj\displaystyle(w\widetilde{S}_{\lambda^{1}})^{2}~{}w\widetilde{S}_{\lambda^{j}} =wS~λ1((wS~λ1|λj)wS~λj+λiλjc0cjwS~λi)\displaystyle=w\widetilde{S}_{\lambda^{1}}~{}(({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{j}})~{}w\widetilde{S}_{\lambda^{j}}+\sum_{\lambda^{i}\to\lambda^{j}}\dfrac{c_{0}}{c_{j}}~{}w\widetilde{S}_{\lambda^{i}})
=(wS~λ1|λj)2wS~λj+λiλj(wS~λ1|λj)c0cjwS~λi\displaystyle=({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{j}})^{2}~{}w\widetilde{S}_{\lambda^{j}}+\sum_{\lambda^{i}\to\lambda^{j}}({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{j}})~{}\dfrac{c_{0}}{c_{j}}~{}w\widetilde{S}_{\lambda^{i}}
+λiλjc0cj(wS~λ1|λi)wS~λi+λkλiλjc0cjc0ciwS~λk.\displaystyle\quad+\sum_{\lambda^{i}\to\lambda^{j}}\dfrac{c_{0}}{c_{j}}~{}({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}})~{}w\widetilde{S}_{\lambda^{i}}+\sum_{\lambda^{k}\to\lambda^{i}\to\lambda^{j}}\dfrac{c_{0}}{c_{j}}\dfrac{c_{0}}{c_{i}}~{}w\widetilde{S}_{\lambda^{k}}.
Proposition 5.8.

For any three Schubert symbols λi,λj and λk\lambda^{i},\lambda^{j}\text{ and }\lambda^{k}, we have the following recurrence relation

(wS~λ1|λkwS~λ1|λi)wcijk=(λsλic0ciwcsjkλkλtc0ctwcijt).({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{k}}-{w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}})wc_{ij}^{k}=(\sum_{\lambda^{s}\to\lambda^{i}}\dfrac{c_{0}}{c_{i}}wc_{sj}^{k}-\sum_{\lambda^{k}\to\lambda^{t}}\dfrac{c_{0}}{c_{t}}wc_{ij}^{t}).
Proof.

We use the associativity of the multiplication in HTn(WGr(d,n);)H_{T^{n}}^{*}(\mbox{WGr}(d,n);\mathbb{Z}) and weighted Pieri rule to expand wS~λ1wS~λiwS~λjw\widetilde{S}_{\lambda^{1}}w\widetilde{S}_{\lambda^{i}}w\widetilde{S}_{\lambda^{j}} in two different ways.

(5.3) (wS~λ1wS~λi)wS~λj\displaystyle(w\widetilde{S}_{\lambda^{1}}w\widetilde{S}_{\lambda^{i}})w\widetilde{S}_{\lambda^{j}} =((wS~λ1|λi)wS~λi+λsλic0ciwS~λs)wS~λj\displaystyle=(({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}})w\widetilde{S}_{\lambda^{i}}+\sum_{\lambda^{s}\to\lambda^{i}}\dfrac{c_{0}}{c_{i}}w\widetilde{S}_{\lambda^{s}})w\widetilde{S}_{\lambda^{j}}
=(wS~λ1|λi)λlwcijlwS~λl+λsλic0ciλlwcsjlwS~λl.\displaystyle=({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}})\sum_{\lambda^{l}}wc_{ij}^{l}w\widetilde{S}_{\lambda^{l}}+\sum_{\lambda^{s}\to\lambda^{i}}\dfrac{c_{0}}{c_{i}}\sum_{\lambda^{l}}wc_{sj}^{l}w\widetilde{S}_{\lambda^{l}}.
(5.4) wS~λ1(wS~λiwS~λj)\displaystyle w\widetilde{S}_{\lambda^{1}}(w\widetilde{S}_{\lambda^{i}}w\widetilde{S}_{\lambda^{j}}) =wS~λ1λlwcijlwS~λl\displaystyle=w\widetilde{S}_{\lambda^{1}}\sum_{\lambda^{l}}wc_{ij}^{l}w\widetilde{S}_{\lambda^{l}}
=λlwcijl((wS~λ1|λl)wS~λl+λrλlc0clwS~λr).\displaystyle=\sum_{\lambda^{l}}wc_{ij}^{l}(({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{l}})w\widetilde{S}_{\lambda^{l}}+\sum_{\lambda^{r}\to\lambda^{l}}\dfrac{c_{0}}{c_{l}}w\widetilde{S}_{\lambda^{r}}).

Comparing the coefficient of wS~λkw\widetilde{S}_{\lambda^{k}} in (5.3) and (5.4) we get

(wS~λ1|λi)wcijk+λsλic0ciwcsjk=wcijk(wS~λ1|λk)+λkλtc0ctwcijt.({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{i}})wc_{ij}^{k}+\sum_{\lambda^{s}\to\lambda^{i}}\dfrac{c_{0}}{c_{i}}wc_{sj}^{k}=wc_{ij}^{k}({w\widetilde{S}_{\lambda^{1}}}|_{\lambda^{k}})+\sum_{\lambda^{k}\to\lambda^{t}}\dfrac{c_{0}}{c_{t}}wc_{ij}^{t}.

Acknowledgment. The first author thanks ‘Indian Institute of Technology Madras’ for PhD fellowship. The second author thanks ‘Indian Institute of Technology Madras’ for SEED research grant.

References

  • [ALR07] Alejandro Adem, Johann Leida, and Yongbin Ruan, Orbifolds and stringy topology, Cambridge Tracts in Mathematics, vol. 171, Cambridge University Press, Cambridge, 2007. MR 2359514 (2009a:57044)
  • [AM15] Hiraku Abe and Tomoo Matsumura, Equivariant cohomology of weighted Grassmannians and weighted Schubert classes, Int. Math. Res. Not. IMRN (2015), no. 9, 2499–2524. MR 3344679
  • [Bai56] Walter L. Baily, Jr., The decomposition theorem for VV-manifolds, Amer. J. Math. 78 (1956), 862–888. MR 100103
  • [BFR09] Anthony Bahri, Matthias Franz, and Nigel Ray, The equivariant cohomology ring of weighted projective space, Math. Proc. Cambridge Philos. Soc. 146 (2009), no. 2, 395–405. MR 2475973 (2010a:57054)
  • [BNSS21] Anthony Bahri, Dietrich Notbohm, Soumen Sarkar, and Jongbaek Song, On integral cohomology of certain orbifolds, Int. Math. Res. Not. IMRN (2021), no. 6, 4140–4168. MR 4230390
  • [CR02] Alessio Corti and Miles Reid, Weighted Grassmannians, Algebraic geometry, de Gruyter, Berlin, 2002, pp. 141–163. MR 1954062
  • [CR04] W. Chen and Y. Ruan, A new cohomology theory of orbifold, Comm. Math. Phys. 248 (2004), no. 1, 1–31. MR 2104605 (2005j:57036)
  • [DKS22] Alastair Darby, Shintaro Kuroki, and Jongbaek Song, Equivariant cohomology of torus orbifolds, Canadian Journal of Mathematics 74 (2022), no. 2, 299–328.
  • [GGKRW18] Fernando Galaz-García, Martin Kerin, Marco Radeschi, and Michael Wiemeler, Torus orbifolds, slice-maximal torus actions, and rational ellipticity, Int. Math. Res. Not. IMRN (2018), no. 18, 5786–5822. MR 3862119
  • [Gor78] R. Mark Goresky, Triangulation of stratified objects, Proc. Amer. Math. Soc. 72 (1978), no. 1, 193–200. MR 0500991 (58 #18473)
  • [Han05] Bernhard Hanke, Geometric versus homotopy theoretic equivariant bordism, Math. Ann. 332 (2005), no. 3, 677–696. MR 2181767 (2006j:55003)
  • [Hat02] Allen Hatcher, Algebraic topology, Cambridge University Press, Cambridge, 2002. MR MR1867354 (2002k:55001)
  • [HHH05] Megumi Harada, André Henriques, and Tara S. Holm, Computation of generalized equivariant cohomologies of Kac-Moody flag varieties, Adv. Math. 197 (2005), no. 1, 198–221. MR 2166181 (2006h:53086)
  • [HHRW16] Megumi Harada, Tara S. Holm, Nigel Ray, and Gareth Williams, The equivariant KK-theory and cobordism rings of divisive weighted projective spaces, Tohoku Math. J. (2) 68 (2016), no. 4, 487–513. MR 3605445
  • [Hus94] Dale Husemoller, Fibre bundles, third ed., Graduate Texts in Mathematics, vol. 20, Springer-Verlag, New York, 1994. MR 1249482
  • [JP03] Xiaoxiang Jiao and Jiagui Peng, Pseudo-holomorphic curves in complex Grassmann manifolds, Trans. Amer. Math. Soc. 355 (2003), no. 9, 3715–3726. MR 1990170
  • [Kaw73] T. Kawasaki, Cohomology of twisted projective spaces and lens complexes, Math. Ann. 206 (1973), 243–248. MR 0339247 (49 #4008)
  • [KT03] Allen Knutson and Terence Tao, Puzzles and (equivariant) cohomology of Grassmannians, Duke Math. J. 119 (2003), no. 2, 221–260. MR 1997946
  • [Lak72] Dan Laksov, Algebraic cycles on Grassmann varieties, Advances in Math. 9 (1972), 267–295. MR 318145
  • [May96] J. P. May, Equivariant homotopy and cohomology theory, CBMS Regional Conference Series in Mathematics, vol. 91, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1996, With contributions by M. Cole, G. Comezaña, S. Costenoble, A. D. Elmendorf, J. P. C. Greenlees, L. G. Lewis, Jr., R. J. Piacenza, G. Triantafillou, and S. Waner. MR 1413302
  • [MS74] John W. Milnor and James D. Stasheff, Characteristic classes, Princeton University Press, Princeton, N. J., 1974, Annals of Mathematics Studies, No. 76. MR 0440554 (55 #13428)
  • [Muk15] Amiya Mukherjee, Differential topology, second ed., Hindustan Book Agency, New Delhi; Birkhäuser/Springer, Cham, 2015. MR 3379695
  • [PS10] Mainak Poddar and Soumen Sarkar, On quasitoric orbifolds, Osaka J. Math. 47 (2010), no. 4, 1055–1076. MR 2791564 (2012e:57058)
  • [Sat56] I. Satake, On a generalization of the notion of manifold, Proc. Nat. Acad. Sci. U.S.A. 42 (1956), 359–363. MR 79769
  • [Sat57] Ichirô Satake, The Gauss-Bonnet theorem for VV-manifolds, J. Math. Soc. Japan 9 (1957), 464–492. MR 95520
  • [Seg68] Graeme Segal, Equivariant KK-theory, Inst. Hautes Études Sci. Publ. Math. (1968), no. 34, 129–151. MR 0234452
  • [Sin01] Dev P. Sinha, Computations of complex equivariant bordism rings, Amer. J. Math. 123 (2001), no. 4, 577–605. MR 1844571 (2002g:55008)
  • [SS21] Soumen Sarkar and Jongbaek Song, GKM theory for orbifold stratified spaces and application to singular toric varieties, Topology Appl. 288 (2021), 107472, 15. MR 4186079
  • [tD70] Tammo tom Dieck, Bordism of GG-manifolds and integrality theorems, Topology 9 (1970), 345–358. MR 0266241 (42 #1148)
  • [Thu80] William P. Thurston, The geometry and topology of three-manifolds, Princeton Lecture Notes, 1980, www.msri.org/publications/books/gt3m/.