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

Quantum integrable model for the quantum cohomology/K-theory of flag varieties and the double β\beta-Grothendieck polynomials

Jirui Guo School of Mathematical Sciences,
Institute for Advanced Study,
Key Laboratory of Intelligent Computing and Applications,
Tongji University, Shanghai 200092, China
(jrkwok@tongji.edu.cn)
Abstract

A GL(n)(n) quantum integrable system generalizing the asymmetric five vertex spin chain is shown to encode the ring relations of the equivariant quantum cohomology and equivariant quantum K-theory ring of flag varieties. We also show that the Bethe ansatz states of this system generate the double β\beta-Grothendieck polynomials.

1 Introduction

The mathematical structure and exact solutions of quantum integrable systems have proved to be useful tools in providing information about physical phenomena, as well as revealing certain algebraic results. For example, in [2], it was shown that the underlying Frobenius algebra of the asymmetric six vertex model, which depends on a parameter β\beta, is exactly the equivariant quantum cohomology of the Grassmannians when β=0\beta=0, and the equivariant quantum K-theory ring of the Grassmannians when β=1\beta=-1. The Bethe/Gauge correspondence of this model was later studied in [3].

This work aims at generalizing the idea of [2], establishing a relationship between a GL(n)(n) integrable model and the equivariant quantum cohomology/quantum K-theory ring of general flag varieties. We will also show that the Bethe ansatz states of this model generate the double β\beta-Grothendieck polynomials. When β=0\beta=0 these polynomials reduce to the double Schubert polynomials representing Schubert classes in the equivariant cohomology ring, and when β=1\beta=-1 they reduce to the double Grothendieck polynomials representing Schubert classes in the equivariant K-theory ring.

The relationships between the quantum cohomology/K-theory of flag varieties and Toda lattices were discussed in [4] and [5]. Moreover, it was shown that quantum K-theory ring of the cotangent bundle of flag varieties can be realized by the XXZ model, and in certain limit it reduces to the quantum K-theory ring of flag varieties [6]. Double β\beta-Grothendieck polynomials can also be realized by the partition functions of certain 2d lattice models [7, 8, 9].

Our main results are as follows:

  • The Bethe ansatz equations of the GL(n)(n) quantum integrable system defined by the R-matrix (2.3)

    R(n)(x,y)=(100000In100(1+β(xy))In1(xy)In10000R(n1)(x,y)),R^{(n)}(x,y)=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&0&I_{n-1}&0\\ 0&(1+\beta(x\ominus y))I_{n-1}&(x\ominus y)I_{n-1}&0\\ 0&0&0&R^{(n-1)}(x,y)\end{array}\right),

    where

    R(2)(x,y)=(1000001001+β(xy)xy00001),R^{(2)}(x,y)=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&0&1&0\\ 0&1+\beta(x\ominus y)&x\ominus y&0\\ 0&0&0&1\end{array}\right),

    are derived (Theorem 3.1).

  • It is shown that the Bathe ansatz equations give rise to the Whitney type ring relations of the equivariant quantum cohomology ring of the partial flag variety Fl(kn1,,k2,k1;N)\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N) when β=0\beta=0 in terms of the Chern classes (Eq.(3.22)):

    c(𝒮i)c(𝒮i+1/𝒮i)=c(𝒮i+1)+(1)kni1kniqnic(𝒮i1),c(\mathcal{S}_{i})\cdot c(\mathcal{S}_{i+1}/\mathcal{S}_{i})=c(\mathcal{S}_{i+1})+(-1)^{k_{n-i-1}-k_{n-i}}q_{n-i}c(\mathcal{S}_{i-1}),

    and give rise to the ring relations of the equivariant quantum K-theory ring of Fl(kn1,,k2,k1;N)\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N) when β=1\beta=-1 in terms of the λy\lambda_{y} classes (Eq.(3.27)):

    λy(𝒮i)λy(𝒮i+1/𝒮i)=λy(𝒮i+1)+qni1qniykni1kni(λy(𝒮i1)λy(𝒮i))det(𝒮i+1/𝒮i),\lambda_{y}(\mathcal{S}_{i})*\lambda_{y}(\mathcal{S}_{i+1}/\mathcal{S}_{i})=\lambda_{y}(\mathcal{S}_{i+1})+\frac{q_{n-i}}{1-q_{n-i}}y^{k_{n-i-1}-k_{n-i}}\left(\lambda_{y}(\mathcal{S}_{i-1})-\lambda_{y}(\mathcal{S}_{i})\right)*\det(\mathcal{S}_{i+1}/\mathcal{S}_{i}),

    where 𝒮i\mathcal{S}_{i} is the iith tautological bundle of the flag variety.

  • Given the matrix elements Bi(u)B_{i}(u) of the monodromy matrix as in Eq.(2.22), and the pseudo vacuum state |Ω(0)\ket{\Omega^{(0)}} as in (3.2), the expansion coefficients in the natural basis of the quantum state BN1(σ1)B1(σN1)|Ω(0)B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}} are exactly the double β\beta-Grothendieck polynomials 𝒢π(β)\mathcal{G}^{(\beta)}_{\pi} (Theorem 4.6):

    BN1(σ1)B1(σN1)|Ω(0)=πSN𝒢π(β)(σ1,,σN1;t1,,tN1)|ωNπ1,B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}}=\sum_{\pi\in S_{N}}\mathcal{G}^{(\beta)}_{\pi}(\sigma_{1},\cdots,\sigma_{N-1};\ominus t_{1},\cdots,\ominus t_{N-1})\ket{\omega_{N}\pi^{-1}},

    where ωN\omega_{N} is the permutation with maximal length in the permutation group SNS_{N}.

  • When the elementary symmetric polynomial el(σ1(i),,σNi(i))e_{l}(\sigma^{(i)}_{1},\cdots,\sigma^{(i)}_{N-i}) is identified with el(x1,,xNi)e_{l}(x_{1},\cdots,x_{N-i}) for all i=1,,N1i=1,\cdots,N-1 and l=1,,Nil=1,\cdots,N-i, the Bethe ansatz state (3.4) (in the full flag case) has the expansion (Theorem 4.11):

    |ψ(0)\displaystyle\ket{\psi^{(0)}} =i1,,ik1ψi1ik1(1)Bi1(σ1(1))Bik1(σk1(1))|Ω(0)\displaystyle=\sum_{i_{1},\cdots,i_{k_{1}}}\psi_{i_{1}\cdots i_{k_{1}}}^{(1)}B_{i_{1}}(\sigma_{1}^{(1)})\cdots B_{i_{k_{1}}}(\sigma_{k_{1}}^{(1)})\ket{\Omega^{(0)}}
    =wSN𝒢w(β)(x1,,xN1;t1,,tN1)|ωNw1,\displaystyle=\sum_{w\in S_{N}}\mathcal{G}^{(\beta)}_{w}(x_{1},\cdots,x_{N-1};\ominus t_{1},\cdots,\ominus t_{N-1})\ket{\omega_{N}w^{-1}},

    where xix_{i} can be interpreted as the first Chern class of the line bundle 𝒮i/𝒮i1\mathcal{S}_{i}/\mathcal{S}_{i-1} when β=0\beta=0, and 1xi1-x_{i} can be interpreted as the K-theory class of 𝒮i/𝒮i1\mathcal{S}_{i}/\mathcal{S}_{i-1} when β=1\beta=-1 (This means σa(Ni)\sigma^{(N-i)}_{a} (β=0\beta=0) or 1σa(Ni)1-\sigma^{(N-i)}_{a} (β=1\beta=-1), a=1,,ia=1,\cdots,i, should be interpreted as the Chern roots of 𝒮i\mathcal{S}_{i}).

This paper is organized as follows. In Sec. 2, the basic ingredients of the GL(n)(n) quantum integrable system generalizing the asymmetric five vertex spin chain are introduced. In Sec. 3, the Bethe ansatz equations are derived, which are shown to give rise to the ring relations of the equivariant quantum cohomology/K-theory ring of the flag varieties. Sec. 4 consists of a series of propositions leading to our main results on the relationship between the quantum integrable model under study and the double β\beta-Grothendieck polynomials. In the appendix, we present some background of the quantum cohomology/K-theory of flag varieties and the double β\beta-Grothendieck polynomials.

2 The GL(n)(n) asymmetric five vertex model

In this section, we present the basic ingredients of the GL(n)(n) asymmetric five vertex model. Let V=nV=\mathbb{C}^{n}, we define the R-matrix R(n)(x,y)End(VV)R^{(n)}(x,y)\in\mathrm{End}(V\otimes V) as follows:
For n=2n=2, the R-matrix is defined as

R(2)(x,y)=(1000001001+β(xy)xy00001),R^{(2)}(x,y)=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&0&1&0\\ 0&1+\beta(x\ominus y)&x\ominus y&0\\ 0&0&0&1\end{array}\right), (2.1)

where β\beta is a formal variable, and

xyxy1+βy.x\ominus y\equiv\frac{x-y}{1+\beta y}. (2.2)

For n>2n>2, under the decomposition Vn1V\cong\mathbb{C}\oplus\mathbb{C}^{n-1}, R(n)(x,y)R^{(n)}(x,y) is defined as follows:

R(n)(x,y)=(100000In100(1+β(xy))In1(xy)In10000R(n1)(x,y)),R^{(n)}(x,y)=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&0&I_{n-1}&0\\ 0&(1+\beta(x\ominus y))I_{n-1}&(x\ominus y)I_{n-1}&0\\ 0&0&0&R^{(n-1)}(x,y)\end{array}\right), (2.3)

where IrI_{r} stands for the identity matrix of dimension rr. The entries of the R-matrix can be written as

Rij,kl(n)(x,y)=(1+δk<lβ(xy))δilδjk+δk>l(xy)δikδjl,R^{(n)}_{ij,kl}(x,y)=(1+\delta_{k<l}\beta(x\ominus y))\delta_{il}\delta_{jk}+\delta_{k>l}(x\ominus y)\delta_{ik}\delta_{jl}, (2.4)

where i,ji,j are row indices, k,lk,l are colomn indices, {i,k}\{i,k\} and {j,l}\{j,l\} are indices for the first and second factor of VVV\otimes V respectively. We also define

δi<j={1,ifi<j,0,ifij,δi>j={1,ifi>j,0,ifij.\delta_{i<j}=\left\{\begin{array}[]{ll}1,&~{}\mathrm{if}~{}i<j,\\ 0,&~{}\mathrm{if}~{}i\geq j,\end{array}\right.\quad\delta_{i>j}=\left\{\begin{array}[]{ll}1,&~{}\mathrm{if}~{}i>j,\\ 0,&~{}\mathrm{if}~{}i\leq j.\end{array}\right.

The R-matrix (2.3) appeared in [9] as the q0q\rightarrow 0 limit of a Drinfeld twist of the R-matrix for an evaluation module of Uq(sl^n+1)U_{q}(\widehat{\mathrm{sl}}_{n+1}), but here we show that R(n)(x,y)R^{(n)}(x,y) satisfies the Yang-Baxter equation by a direct computation:

Theorem 2.1.

The R-matrix defined by Eq.(2.3) satisfies the Yang-Baxter equation

R12(n)(x,y)R13(n)(x,z)R23(n)(y,z)=R23(n)(y,z)R13(n)(x,z)R12(n)(x,y)R^{(n)}_{12}(x,y)R^{(n)}_{13}(x,z)R^{(n)}_{23}(y,z)=R^{(n)}_{23}(y,z)R^{(n)}_{13}(x,z)R^{(n)}_{12}(x,y) (2.5)

for all n>1n>1, where Rab(n)R^{(n)}_{ab} acts on the aath and bbth factor of VVVV\otimes V\otimes V.

Proof.

A direct computation shows that Eq.(2.5) holds for n=2n=2. Now suppose R(n1)R^{(n-1)} satisfies the Yang-Baxter equation for some n>2n>2. Let iabi_{ab} denote the identity map from the aath factor to the bbth factor of VVVV\otimes V\otimes V. Plugging (2.3) into Eq.(2.5), we find that Eq.(2.5) decomposes into the following identities:

i21i32=i31,i13=i23i12,\displaystyle i_{21}i_{32}=i_{31},\quad i_{13}=i_{23}i_{12}, (2.6)
i12i31i23=i32i13i21,i21i13i32=i23i31i12,\displaystyle i_{12}i_{31}i_{23}=i_{32}i_{13}i_{21},\quad i_{21}i_{13}i_{32}=i_{23}i_{31}i_{12}, (2.7)
(xy)i31i23+(yz)i21=i21(xz),(xz)i12=(xy)i32i13+(yz)i12,\displaystyle(x-y)i_{31}i_{23}+(y-z)i_{21}=i_{21}(x-z),\quad(x-z)i_{12}=(x-y)i_{32}i_{13}+(y-z)i_{12}, (2.8)
R12(n1)(x,y)i31R23(n1)(y,z)=i32R13(n1)(x,z)i21,\displaystyle R^{(n-1)}_{12}(x,y)i_{31}R^{(n-1)}_{23}(y,z)=i_{32}R^{(n-1)}_{13}(x,z)i_{21}, (2.9)
i12R13(n1)(x,z)i23=R23(n1)(y,z)i13R12(n1)(x,y),\displaystyle i_{12}R^{(n-1)}_{13}(x,z)i_{23}=R_{23}^{(n-1)}(y,z)i_{13}R_{12}^{(n-1)}(x,y), (2.10)
(xz)(1+βy)R12(n1)(x,y)i32=(yz)(1+βx)i31i12+(xy)(1+βz)i32R13(n1)(x,z),\displaystyle(x-z)(1+\beta y)R_{12}^{(n-1)}(x,y)i_{32}=(y-z)(1+\beta x)i_{31}i_{12}+(x-y)(1+\beta z)i_{32}R_{13}^{(n-1)}(x,z), (2.11)
(yz)(1+βx)i21i13+(xy)(1+βz)R13(n1)(x,z)i23=(xz)(1+βy)i23R12(n1)(x,y),\displaystyle(y-z)(1+\beta x)i_{21}i_{13}+(x-y)(1+\beta z)R_{13}^{(n-1)}(x,z)i_{23}=(x-z)(1+\beta y)i_{23}R_{12}^{(n-1)}(x,y), (2.12)
R12(n1)(x,y)R13(n1)(x,z)R23(n1)(y,z)=R23(n1)(y,z)R13(n1)(x,z)R12(n1)(x,y).\displaystyle R^{(n-1)}_{12}(x,y)R^{(n-1)}_{13}(x,z)R^{(n-1)}_{23}(y,z)=R^{(n-1)}_{23}(y,z)R^{(n-1)}_{13}(x,z)R^{(n-1)}_{12}(x,y). (2.13)

Eq.(2.6)-(2.8) are direct consequences of the definition of iabi_{ab}. Eq.(2.9) and (2.10) can be easily derived from the identities

k,lRij,lk(m)(x,y)Rkl,st(m)(y,z)=Rij,st(m)(x,z)\sum_{k,l}R^{(m)}_{ij,lk}(x,y)R^{(m)}_{kl,st}(y,z)=R^{(m)}_{ij,st}(x,z) (2.14)

and

k,lRij,lk(m)(y,z)Rkl,st(m)(x,y)=Rij,st(m)(x,z)\sum_{k,l}R^{(m)}_{ij,lk}(y,z)R^{(m)}_{kl,st}(x,y)=R^{(m)}_{ij,st}(x,z) (2.15)

respectively, while Eq.(2.11) and (2.12) can be easily derived from

(yz)(1+βx)R(m)(x,x)+(xy)(1+βz)R(m)(x,z)=(xz)(1+βy)R(m)(x,y),(y-z)(1+\beta x)R^{(m)}(x,x)+(x-y)(1+\beta z)R^{(m)}(x,z)=(x-z)(1+\beta y)R^{(m)}(x,y), (2.16)

where m2m\geq 2. Eq.(2.14)-(2.16) can all be proved by simple induction on mm. Finally, Eq.(2.13) is the induction hypothesis. Therefore, by induction, the Yang-Baxter equation holds for all n2n\geq 2. ∎

To construct an integrable system, we take NN copies of VnV\cong\mathbb{C}^{n} and label them by Vi,i=1,,NV_{i},i=1,\cdots,N. Let VaVV_{a}\cong V be an auxiliary space. The monodromy matrix is defined by

Ta(n)(u)=RaN(u,tN)Ra,N1(u,tN1)Ra2(u,t2)Ra1(u,t1),T^{(n)}_{a}(u)=R_{aN}(u,t_{N})R_{a,N-1}(u,t_{N-1})\cdots R_{a2}(u,t_{2})R_{a1}(u,t_{1}), (2.17)

where RaiR_{ai} acts on VaViV_{a}\otimes V_{i}. The Yang-Baxter equation (2.5) leads to the following RTT relation:

Rab(n)(x,y)Ta(n)(x)Tb(n)(y)=Tb(n)(y)Ta(n)(x)Rab(n)(x,y).R^{(n)}_{ab}(x,y)T^{(n)}_{a}(x)T^{(n)}_{b}(y)=T^{(n)}_{b}(y)T^{(n)}_{a}(x)R^{(n)}_{ab}(x,y). (2.18)

We define the transfer matrix to be

t(n)(u)=TrVa(MnTa(n)(u)),t^{(n)}(u)=\mathrm{Tr}_{V_{a}}(M_{n}T^{(n)}_{a}(u)), (2.19)

where Mn=diag(b0,b1,,bn1)M_{n}=\mathrm{diag}(b_{0},b_{1},\cdots,b_{n-1}) acts on VaV_{a}, imposing a twisted periodic boundary condition. In the following, we also need

Ml:=diag(bnl,,bn1),l=1,,n.M_{l}:=\mathrm{diag}(b_{n-l},\cdots,b_{n-1}),\quad l=1,\cdots,n. (2.20)

Eq.(2.18) implies that111Notice that for any diagonal matrix MM acting on the auxiliary space, we have Rab(n)(x,y)MaMb=MbMaRab(n)(x,y)R^{(n)}_{ab}(x,y)M_{a}M_{b}=M_{b}M_{a}R^{(n)}_{ab}(x,y).

[t(u),t(v)]=0.[t(u),t(v)]=0. (2.21)

Under the decomposition Va=n1V_{a}=\mathbb{C}\oplus\mathbb{C}^{n-1}, the monodromy matrix can be written in the following form:

Ta(n)(u)=(A(u)B1(u)Bn1(u)C1(u)D11(u)D1,n1(u)Cn1(u)Dn1,1(u)Dn1,n1(u))T^{(n)}_{a}(u)=\left(\begin{array}[]{cccc}A(u)&B_{1}(u)&\cdots&B_{n-1}(u)\\ C_{1}(u)&D_{11}(u)&\cdots&D_{1,n-1}(u)\\ \vdots&\vdots&\ddots&\vdots\\ C_{n-1}(u)&D_{n-1,1}(u)&\cdots&D_{n-1,n-1}(u)\end{array}\right) (2.22)

with A(u),Bi(u),Ci(u),Dij(u)A(u),B_{i}(u),C_{i}(u),D_{ij}(u) being operators acting on the Hilbert space =V1V2VN\mathcal{H}=V_{1}\otimes V_{2}\otimes\cdots\otimes V_{N}. The RTT relation (2.18) yields the following commutation ralations:

A(x)Bi(y)\displaystyle A(x)B_{i}(y) =1yx[Bi(y)A(x)Bi(x)A(y)],\displaystyle=\frac{1}{y\ominus x}\left[B_{i}(y)A(x)-B_{i}(x)A(y)\right], (2.23)
Dij(x)Bk(y)\displaystyle D_{ij}(x)B_{k}(y) =s,l[1xyBs(y)Dil(x)Rls,jk(n1)(x,y)]+1yxBj(x)Dik(y),\displaystyle=\sum_{s,l}\left[\frac{1}{x\ominus y}B_{s}(y)D_{il}(x)R^{(n-1)}_{ls,jk}(x,y)\right]+\frac{1}{y\ominus x}B_{j}(x)D_{ik}(y), (2.24)
Bi(x)Bj(y)\displaystyle B_{i}(x)B_{j}(y) =r,sBr(y)Bs(x)Rsr,ij(n1)(x,y).\displaystyle=\sum_{r,s}B_{r}(y)B_{s}(x)R^{(n-1)}_{sr,ij}(x,y). (2.25)

3 Bethe ansatz equations and the equivariant quantum
cohomology/K-theory ring of flag varieties

In this section, we derive the Bethe ansatz equations of the GL(n)(n) five vertex model introduced in the last section following the method proposed in [10], and discuss their relationship with the equivariant quantum cohomology and quantum K-theory ring of flag varieties.

Let V=nV=\mathbb{C}^{n} be spanned by v0,v1,,vn1v_{0},v_{1},\cdots,v_{n-1}, where vi:=|iv_{i}:=\ket{i} is the (i+1)(i+1)th unit vector of n\mathbb{C}^{n}. Define

|n1,n2,,nN:=vn1vn2vnN,\ket{n_{1},n_{2},\cdots,n_{N}}:=v_{n_{1}}\otimes v_{n_{2}}\otimes\cdots\otimes v_{n_{N}},

then a basis of the Hilbert space =V1V2VN\mathcal{H}=V_{1}\otimes V_{2}\otimes\cdots V_{N} can be chosen to be

{|n1,n2,,nN|0nin1}.\{\ket{n_{1},n_{2},\cdots,n_{N}}|~{}0\leq n_{i}\leq n-1~{}\}. (3.1)

We call the basis (3.1) the natural basis of the Hilbert space \mathcal{H}. The pseudo vacuum is taken to be

|Ω(0)=|0,0,,0V1V2VN.\ket{\Omega^{(0)}}=\ket{0,0,\cdots,0}\in V_{1}\otimes V_{2}\otimes\cdots\otimes V_{N}. (3.2)

From Eq.(2.3) and (2.22), it is easy to compute

A(u)|Ω(0)=|Ω(0),Dij(u)|Ω(0)=k=1N(utk)δij|Ω(0).A(u)\ket{\Omega^{(0)}}=\ket{\Omega^{(0)}},\quad D_{ij}(u)\ket{\Omega^{(0)}}=\prod_{k=1}^{N}(u\ominus t_{k})\delta_{ij}\ket{\Omega^{(0)}}. (3.3)

The Bethe ansatz state is taken to be of the form

|ψ(0)=i1,,ik1ψi1ik1(1)Bi1(σ1(1))Bik1(σk1(1))|Ω(0),\ket{\psi^{(0)}}=\sum_{i_{1},\cdots,i_{k_{1}}}\psi_{i_{1}\cdots i_{k_{1}}}^{(1)}B_{i_{1}}(\sigma_{1}^{(1)})\cdots B_{i_{k_{1}}}(\sigma_{k_{1}}^{(1)})\ket{\Omega^{(0)}}, (3.4)

where σα1(1)\sigma_{\alpha_{1}}^{(1)}, α1=1,,k1\alpha_{1}=1,\cdots,k_{1}, are parameters to be determined, and the expansion coefficients ψi1ik1(1)\psi_{i_{1}\cdots i_{k_{1}}}^{(1)} depend on n2n-2 sets of parameters

σα2(2),,σαn1(n1)\sigma^{(2)}_{\alpha_{2}},\cdots,\sigma^{(n-1)}_{\alpha_{n-1}}

with αs=1,2,,ks\alpha_{s}=1,2,\cdots,k_{s}, where σa(s)σb(s)\sigma^{(s)}_{a}\neq\sigma^{(s)}_{b} for aba\neq b and 1sn11\leq s\leq n-1. Now we can prove the following

Theorem 3.1.

If the parameters σα1(1),,σαn1(n1)\sigma^{(1)}_{\alpha_{1}},\cdots,\sigma^{(n-1)}_{\alpha_{n-1}} satisfy the following Bethe ansatz equations:

a=1k1(1+β(σa(1)σα1(1)))i=1N(σα1(1)ti)=(1)k11q1b=1k2(σb(2)σα1(1)),\displaystyle\prod_{a=1}^{k_{1}}(1+\beta(\sigma^{(1)}_{a}\ominus\sigma^{(1)}_{\alpha_{1}}))\prod_{i=1}^{N}(\sigma^{(1)}_{\alpha_{1}}\ominus t_{i})=(-1)^{k_{1}-1}q_{1}\prod_{b=1}^{k_{2}}(\sigma^{(2)}_{b}\ominus\sigma^{(1)}_{\alpha_{1}}), (3.5)
a=1km(1+β(σa(m)σαm(m)))i=1km1(σαm(m)σi(m1))=(1)km1qmb=1km+1(σb(m+1)σαm(m)),\displaystyle\prod_{a=1}^{k_{m}}(1+\beta(\sigma^{(m)}_{a}\ominus\sigma^{(m)}_{\alpha_{m}}))\prod_{i=1}^{k_{m-1}}(\sigma^{(m)}_{\alpha_{m}}\ominus\sigma^{(m-1)}_{i})=(-1)^{k_{m}-1}q_{m}\prod_{b=1}^{k_{m+1}}(\sigma^{(m+1)}_{b}\ominus\sigma^{(m)}_{\alpha_{m}}), (3.6)
a=1kn1(1+β(σa(n1)σαn1(n1)))i=1kn2(σαn1(n1)σi(n2))=(1)kn11qn1,\displaystyle\prod_{a=1}^{k_{n-1}}(1+\beta(\sigma^{(n-1)}_{a}\ominus\sigma^{(n-1)}_{\alpha_{n-1}}))\prod_{i=1}^{k_{n-2}}(\sigma^{(n-1)}_{\alpha_{n-1}}\ominus\sigma^{(n-2)}_{i})=(-1)^{k_{n-1}-1}q_{n-1}, (3.7)

where m=2,,n2,αs=1,,ks,qs:=bs1/bs,s=1,,n1m=2,\cdots,n-2,\quad\alpha_{s}=1,\cdots,k_{s},\quad q_{s}:=b_{s-1}/b_{s},~{}s=1,\cdots,n-1, then the state (3.4) is a common eigenstate of the transfer matrices t(u)t(u) defined by (2.19).

Proof.

From the commutation relations (2.23) and (2.24), one can compute

A(u)|ψ(0)=a=1k1(σa(1)u)1|ψ(0)+|ϕA,\displaystyle A(u)\ket{\psi^{(0)}}=\prod_{a=1}^{k_{1}}(\sigma_{a}^{(1)}\ominus u)^{-1}\ket{\psi^{(0)}}+\ket{\phi_{A}}, (3.8)
Dij(u)|ψ(0)=l=1N(utl)a=1k1(uσa(1))\displaystyle D_{ij}(u)\ket{\psi^{(0)}}=\frac{\prod_{l=1}^{N}(u\ominus t_{l})}{\prod_{a=1}^{k_{1}}(u\ominus\sigma^{(1)}_{a})}
×i1,,ik1j1,,jk1{[Tij(n1)(u;σ(1))]i1ik1,j1jk1ψj1jk1(1)Bi1(σ1(1))Bik1(σk1(1))}|Ω(0)+|ϕD,\displaystyle\times\sum_{i_{1},\cdots,i_{k_{1}}}\sum_{j_{1},\cdots,j_{k_{1}}}\left\{[T_{ij}^{(n-1)}(u;\sigma^{(1)})]_{i_{1}\cdots i_{k_{1}},j_{1}\cdots j_{k_{1}}}\psi_{j_{1}\cdots j_{k_{1}}}^{(1)}B_{i_{1}}(\sigma_{1}^{(1)})\cdots B_{i_{k_{1}}}(\sigma_{k_{1}}^{(1)})\right\}\ket{\Omega^{(0)}}+\ket{\phi_{D}},

where |ϕA\ket{\phi_{A}} and |ϕD\ket{\phi_{D}} contain all the unwanted terms that are not in the subspace spanned by states of the form Bi1(σ1(1))Bik1(σk1(1))|Ω(0)B_{i_{1}}(\sigma_{1}^{(1)})\cdots B_{i_{k_{1}}}(\sigma_{k_{1}}^{(1)})\ket{\Omega^{(0)}}, and

[Tij(n1)(u;σ(1))]i1ik1,j1jk1\displaystyle[T_{ij}^{(n-1)}(u;\sigma^{(1)})]_{i_{1}\cdots i_{k_{1}},j_{1}\cdots j_{k_{1}}} (3.9)
=μ1μk11Riik1,μk11jk1(n1)(u,σk1(1))Rμ2i2,μ1j2(n1)(u,σ2(1))Rμ1i1,jj1(n1)(u,σ1(1)).\displaystyle=\sum_{\mu_{1}\cdots\mu_{k_{1}-1}}R^{(n-1)}_{ii_{k_{1}},\mu_{k_{1}-1}j_{k_{1}}}(u,\sigma^{(1)}_{k_{1}})\cdots R^{(n-1)}_{\mu_{2}i_{2},\mu_{1}j_{2}}(u,\sigma^{(1)}_{2})R^{(n-1)}_{\mu_{1}i_{1},jj_{1}}(u,\sigma^{(1)}_{1}).

Eq.(3.9) suggests that T(n1)T^{(n-1)} can be viewed as the monodromy matrix of the GL(n1)(n-1) asymmetric five vertex model with k1k_{1} sites and R-matrix R(n1)R^{(n-1)}, where σa(1),a=1,,k1\sigma^{(1)}_{a},a=1,\cdots,k_{1} are the equivariant parameters. Correspondingly, the transfer matrix can be defined as

t(n1)(u;σ(1))=Tr(Mn1T(n1)(u;σ(1))),t^{(n-1)}(u;\sigma^{(1)})=\mathrm{Tr}(M_{n-1}T^{(n-1)}(u;\sigma^{(1)})),

where MlM_{l} is defined by (2.20). Let

ψi1ik1(1)=i1i2ik1|ψ(1),is=1,,n1.\psi_{i_{1}\cdots i_{k_{1}}}^{(1)}=\braket{i_{1}i_{2}\cdots i_{k_{1}}}{\psi^{(1)}},\quad i_{s}=1,\cdots,n-1.

If |ψ(1)\ket{\psi^{(1)}} is an eigenstate of t(n1)(u;σ(1))t^{(n-1)}(u;\sigma^{(1)}) with eigenvalue Λ(1)(u)\Lambda^{(1)}(u), i.e.

t(n1)(u;σ(1))|ψ(1)=Λ(1)(u)|ψ(1),t^{(n-1)}(u;\sigma^{(1)})\ket{\psi^{(1)}}=\Lambda^{(1)}(u)\ket{\psi^{(1)}}, (3.10)

then, as a consequence of Eq.(3.8)

Tr(Mn1D(u))|ψ(0)=i=1N(uti)a=1k1(uσa(1))Λ(1)(u)|ψ(0)+|ϕ~D,\mathrm{Tr}(M_{n-1}D(u))\ket{\psi^{(0)}}=\frac{\prod_{i=1}^{N}(u\ominus t_{i})}{\prod_{a=1}^{k_{1}}(u\ominus\sigma^{(1)}_{a})}\Lambda^{(1)}(u)\ket{\psi^{(0)}}+\ket{\tilde{\phi}_{D}},

where |ϕ~D\ket{\tilde{\phi}_{D}} is the sum of the unwanted terms. Then

t(n)(u)|ψ(0)=[b0A(u)+Tr(Mn1D(u))]|ψ(0)=Λ(0)(u)|ψ(0)+|ϕ(n),t^{(n)}(u)\ket{\psi^{(0)}}=\left[b_{0}A(u)+\mathrm{Tr}(M_{n-1}D(u))\right]\ket{\psi^{(0)}}=\Lambda^{(0)}(u)\ket{\psi^{(0)}}+\ket{\phi^{(n)}},

where

Λ(0)(u)=b0a=1k1(σa(1)u)1+i=1N(uti)a=1k1(uσa(1))Λ(1)(u),\Lambda^{(0)}(u)=b_{0}\prod_{a=1}^{k_{1}}(\sigma_{a}^{(1)}\ominus u)^{-1}+\frac{\prod_{i=1}^{N}(u\ominus t_{i})}{\prod_{a=1}^{k_{1}}(u\ominus\sigma^{(1)}_{a})}\Lambda^{(1)}(u), (3.11)

and |ϕ(n)\ket{\phi^{(n)}} is the sum of the unwanted terms. For |ψ(0)\ket{\psi^{(0)}} to be an eigenstate of t(n)(u)t^{(n)}(u) for all uu, the unwanted terms must vanish. As in the usual algebraic Bethe ansatz method, vanishing of the unwanted terms |ϕ(n)\ket{\phi^{(n)}} amounts to Λ(0)(u)\Lambda^{(0)}(u) being holomorphic, i.e.

Res(Λ(0)(u),σa(1))=0\mathrm{Res}(\Lambda^{(0)}(u),\sigma^{(1)}_{a})=0 (3.12)

for all a=1,,k1a=1,\cdots,k_{1}. Eq.(3.12) gives us the equations for σa(1)\sigma_{a}^{(1)}:

α=1k1(1+βσα(1))i=1N(σa(1)ti)Λ(1)(σa(1))=(1)k11b0(1+βσa(1))k1,a=1,,k1.\prod_{\alpha=1}^{k_{1}}(1+\beta\sigma_{\alpha}^{(1)})\prod_{i=1}^{N}(\sigma_{a}^{(1)}\ominus t_{i})\Lambda^{(1)}(\sigma_{a}^{(1)})=(-1)^{k_{1}-1}b_{0}(1+\beta\sigma_{a}^{(1)})^{k_{1}},\quad a=1,\cdots,k_{1}. (3.13)

Thus we have reduced the eigenvalue problem for the GL(n)(n) system to the eigenvalue problem (3.10) for a GL(n1)(n-1) system. Clearly, we can continue to reduce the rank in exactly the same way. At each step, a set of parameters σαm(m),αm=1,,km\sigma^{(m)}_{\alpha_{m}},\alpha_{m}=1,\cdots,k_{m}, has to be introduced to construct the Bethe ansatz state

|ψ(m1)=i1,,ikmψi1,,ikm(m)Bi1(m1)(σ1(m))Bikm(m1)(σkm(m))|Ω(m1),\ket{\psi^{(m-1)}}=\sum_{i_{1},\cdots,i_{k_{m}}}\psi^{(m)}_{i_{1},\cdots,i_{k_{m}}}B^{(m-1)}_{i_{1}}(\sigma^{(m)}_{1})\cdots B^{(m-1)}_{i_{k_{m}}}(\sigma^{(m)}_{k_{m}})\ket{\Omega^{(m-1)}}, (3.14)

where

ψi1,,ikm(m)=i1,,ikm|ψ(m),\psi^{(m)}_{i_{1},\cdots,i_{k_{m}}}=\braket{i_{1},\cdots,i_{k_{m}}}{\psi^{(m)}}, (3.15)

and Bi(m1)B^{(m-1)}_{i}’s are the BiB_{i} operators defined as matrix elements of the monodromy matrix in the corresponding step as in (2.22) and σa(m1),a=1,,km1\sigma^{(m-1)}_{a},a=1,\cdots,k_{m-1}, serve as the equivariant parameters. As in the m=1m=1 case, for (3.14) to be an eigenstate, the following equations have to be satisfied:

b=1km(1+βσb(m))a=1km1(σαm(m)σa(m1))Λ(m)(σαm(m))=(1)km1bm1(1+βσαm(m))km,αm=1,,km,\prod_{b=1}^{k_{m}}(1+\beta\sigma_{b}^{(m)})\prod_{a=1}^{k_{m-1}}(\sigma_{\alpha_{m}}^{(m)}\ominus\sigma_{a}^{(m-1)})\Lambda^{(m)}(\sigma_{\alpha_{m}}^{(m)})=(-1)^{k_{m}-1}b_{m-1}(1+\beta\sigma_{\alpha_{m}}^{(m)})^{k_{m}},\quad\alpha_{m}=1,\cdots,k_{m}, (3.16)

accordingly, the eigenvalues satisfy

Λ(m1)(u)=bm1α=1km(σα(m)u)1+a=1km1(uσa(m1))b=1km(uσb(m))Λ(m)(u)\Lambda^{(m-1)}(u)=b_{m-1}\prod_{\alpha=1}^{k_{m}}(\sigma_{\alpha}^{(m)}\ominus u)^{-1}+\frac{\prod_{a=1}^{k_{m-1}}(u\ominus\sigma_{a}^{(m-1)})}{\prod_{b=1}^{k_{m}}(u\ominus\sigma^{(m)}_{b})}\Lambda^{(m)}(u) (3.17)

for m=2,,n2m=2,\cdots,n-2.

This procedure continues until we get a GL(2)(2) system with R-matrix given by Eq.(2.1). This GL(2)(2) system can be solved exactly by the algebraic Bethe ansatz method as for other spin chain models, leading to the Bethe ansatz equations Eq.(3.7) and

Λ(n2)(u)=bn2α=1kn1(σα(n1)u)1+bn1a=1kn2(uσa(n2))b=1kn1(uσb(n1)).\Lambda^{(n-2)}(u)=b_{n-2}\prod_{\alpha=1}^{k_{n-1}}(\sigma_{\alpha}^{(n-1)}\ominus u)^{-1}+b_{n-1}\frac{\prod_{a=1}^{k_{n-2}}(u\ominus\sigma_{a}^{(n-2)})}{\prod_{b=1}^{k_{n-1}}(u\ominus\sigma^{(n-1)}_{b})}. (3.18)

Eq.(3.18) then allows us to solve all the eigenvalues Λ(m),m=0,1,,n3\Lambda^{(m)},m=0,1,\cdots,n-3, from Eq.(3.17) and (3.11) recursively. Plugging these eigenvalues into Eq.(3.16) and (3.13), we get the equations Eq.(3.6) and (3.5). ∎

Now let us restrict to the cases in which kn1<<k2<k1<Nk_{n-1}<\cdots<k_{2}<k_{1}<N and compare Eq.(3.5)-(3.7) with the ring relations of the equivariant quantum cohomology and equivariant quantum K-theory ring of flag varieties. The reader may refer to the appendix and references therein for a review of the quantum cohomology and quantum K-theory ring of flag varieties.

When β=0\beta=0, we can identify σa(ni)\sigma_{a}^{(n-i)} with the Chern roots of 𝒮i\mathcal{S}_{i}, the iith tautological bundle of the flag variety Fl(kn1,,k2,k1;N)\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N), i.e. the jjth elementary symmetric polynomial ej(σ1(ni),,σkni(ni))e_{j}(\sigma_{1}^{(n-i)},\cdots,\sigma_{k_{n-i}}^{(n-i)}) is identified with the jjth Chern class of 𝒮i\mathcal{S}_{i}:

ej(σ1(ni),,σkni(ni))=cj(𝒮i),j=1,,kni.e_{j}(\sigma_{1}^{(n-i)},\cdots,\sigma_{k_{n-i}}^{(n-i)})=c_{j}(\mathcal{S}_{i}),\quad j=1,\cdots,k_{n-i}.

Under this identification, it was checked in [11, 12] that Eq.(3.5)-(3.7) specialized to β=0\beta=0 are exactly the ring relations of the equivariant quantum cohomology of Fl(kn1,,k2,k1;N)\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N) with the equivariant parameters given by (t1,,tN)(t_{1},\cdots,t_{N}).

When β=1\beta=-1, we can identify 1σa(ni)1-\sigma_{a}^{(n-i)} with the Chern roots of 𝒮i\mathcal{S}_{i}, i.e. the jjth elementary symmetric polynomial ej(1σ1(i),,1σkni(i))e_{j}(1-\sigma_{1}^{(i)},\cdots,1-\sigma_{k_{n-i}}^{(i)}) is identified with j𝒮i\wedge^{j}\mathcal{S}_{i} in the Grothendieck group of the flag variety Fl(kn1,,k2,k1;N)\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N):

ej(1σ1(ni),,1σkni(ni))=j𝒮i,j=1,,kni.e_{j}(1-\sigma_{1}^{(n-i)},\cdots,1-\sigma_{k_{n-i}}^{(n-i)})=\wedge^{j}\mathcal{S}_{i},\quad j=1,\cdots,k_{n-i}.

It was argued in [12] that Eq.(3.5)-(3.7) specialized to β=1\beta=-1 are the ring relations of the (predicted) equivariant quantum K-theory ring of Fl(kn1,,k2,k1;N)\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N) with the equivariant parameters given by (1t1,,1tN)(1-t_{1},\cdots,1-t_{N}).

Indeed, as in [12], one can use Vieta’s formula222Vieta’s formula: If r1,r2,,rnr_{1},r_{2},\cdots,r_{n} are roots of the polynomial anxn+an1xn1++a1x+a0a_{n}x^{n}+a_{n-1}x^{n-1}+\cdots+a_{1}x+a_{0}, then el(r1,,rn)=(1)lanl/ane_{l}(r_{1},\cdots,r_{n})=(-1)^{l}a_{n-l}/a_{n}. to convert Eq.(3.5)-(3.7) into the ring relations of the equivariant quantum cohomology ring (β=0\beta=0) or the equivariant K-theory ring (β=1\beta=-1) of the flag variety Fl(kn1,,k2,k1;N)\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N).

When β=0\beta=0, Eq.(3.5)-(3.7) reduces to

α=1km1(σa(m)σα(m1))=(1)km1qmγ=1km+1(σγ(m+1)σa(m))\prod_{\alpha=1}^{k_{m-1}}(\sigma^{(m)}_{a}-\sigma^{(m-1)}_{\alpha})=(-1)^{k_{m}-1}q_{m}\prod_{\gamma=1}^{k_{m+1}}(\sigma^{(m+1)}_{\gamma}-\sigma^{(m)}_{a}) (3.19)

for a=1,,km,m=1,2,,n1a=1,\cdots,k_{m},m=1,2,\cdots,n-1, where we have defined σi(0)=ti,k0=N,kn=0\sigma^{(0)}_{i}=t_{i},k_{0}=N,k_{n}=0. Eq.(3.19) tells us that σa(m),a=1,,km\sigma^{(m)}_{a},a=1,\cdots,k_{m}, are roots of the polynomial

i=0km1(1)ixkm1iei(σ(m1))+(1)kmkm+1qmj=0km+1(1)jxkm+1jej(σ(m+1)),\sum_{i=0}^{k_{m-1}}(-1)^{i}x^{k_{m-1}-i}e_{i}(\sigma^{(m-1)})+(-1)^{k_{m}-k_{m+1}}q_{m}\sum_{j=0}^{k_{m+1}}(-1)^{j}x^{k_{m+1}-j}e_{j}(\sigma^{(m+1)}), (3.20)

where ei(σ(s))e_{i}(\sigma^{(s)}) is the iith elementary symmetric polynomial in the variables σa(s)\sigma^{(s)}_{a}, a=1,,ksa=1,\cdots,k_{s}. Notice that the coefficient of xkm1lx^{k_{m-1}-l} in (3.20) is333We set el(x1,,xs)=0e_{l}(x_{1},\cdots,x_{s})=0 for l<0l<0 or l>sl>s.

(1)lel(σ(m1))+(1)km1km+lqmekm+1km1+l(σ(m+1)).(-1)^{l}e_{l}(\sigma^{(m-1)})+(-1)^{k_{m-1}-k_{m}+l}q_{m}e_{k_{m+1}-k_{m-1}+l}(\sigma^{(m+1)}).

Assume the other km1kmk_{m-1}-k_{m} roots of (3.20) are ηb(m),b=1,,km1km\eta^{(m)}_{b},b=1,\cdots,k_{m-1}-k_{m}, then Vieta’s formula yields

i=0lei(σ(m))eli(η(m))=el(σ(m1))+(1)km1kmqmekm+1km1+l(σ(m+1)).\sum_{i=0}^{l}e_{i}(\sigma^{(m)})e_{l-i}(\eta^{(m)})=e_{l}(\sigma^{(m-1)})+(-1)^{k_{m-1}-k_{m}}q_{m}e_{k_{m+1}-k_{m-1}+l}(\sigma^{(m+1)}). (3.21)

If we interpret σa(m),a=1,,km,\sigma^{(m)}_{a},a=1,\cdots,k_{m}, as Chern roots of 𝒮nm\mathcal{S}_{n-m} and ηb(m),b=1,,km1km,\eta^{(m)}_{b},b=1,\cdots,k_{m-1}-k_{m}, as Chern roots of 𝒮nm+1/𝒮nm\mathcal{S}_{n-m+1}/\mathcal{S}_{n-m}, i.e. ei(σ(m))=ci(𝒮nm)e_{i}(\sigma^{(m)})=c_{i}(\mathcal{S}_{n-m}), ei(η(m))=ci(𝒮nm+1/𝒮nm)e_{i}(\eta^{(m)})=c_{i}(\mathcal{S}_{n-m+1}/\mathcal{S}_{n-m}), then Eq.(3.21) becomes444𝒮0\mathcal{S}_{0} is the trivial vector bundle of rank zero, so c(𝒮0)=1c({\mathcal{S}_{0}})=1. 𝒮n\mathcal{S}_{n} is the trivial vector bundle of rank NN, but since it carries the TT-action, we have c(𝒮n)=i=0Nei(t1,,tN)c(\mathcal{S}_{n})=\sum_{i=0}^{N}e_{i}(t_{1},\cdots,t_{N}).

c(𝒮i)c(𝒮i+1/𝒮i)=c(𝒮i+1)+(1)kni1kniqnic(𝒮i1),c(\mathcal{S}_{i})\cdot c(\mathcal{S}_{i+1}/\mathcal{S}_{i})=c(\mathcal{S}_{i+1})+(-1)^{k_{n-i-1}-k_{n-i}}q_{n-i}c(\mathcal{S}_{i-1}), (3.22)

for i=1,2,,n1i=1,2,\cdots,n-1, which are the quantum Whitney relations of the equivariant quantum cohomology ring QHT(Fl(kn1,,k2,k1;N))QH^{*}_{T}(\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N)) and are equivalent to the ring relations proved in [13].

When β0\beta\neq 0, let Xa(m)=1+βσa(m)X^{(m)}_{a}=1+\beta\sigma^{(m)}_{a}, then Eq.(3.5)-(3.7) can be written as

α=1km1(Xa(m)Xα(m1))l=1km(Xl(m))\displaystyle\prod_{\alpha=1}^{k_{m-1}}\left(X^{(m)}_{a}-X^{(m-1)}_{\alpha}\right)\prod_{l=1}^{k_{m}}\left(X^{(m)}_{l}\right) (3.23)
=\displaystyle= (1)km1qmβkm1km+1s=1km1(Xs(m1))γ=1km+1(Xγ(m+1)Xa(m))(Xa(m))kmkm+1\displaystyle(-1)^{k_{m}-1}q_{m}\beta^{k_{m-1}-k_{m+1}}\prod_{s=1}^{k_{m-1}}\left(X^{(m-1)}_{s}\right)\prod_{\gamma=1}^{k_{m+1}}\left(X^{(m+1)}_{\gamma}-X^{(m)}_{a}\right)\left(X^{(m)}_{a}\right)^{k_{m}-k_{m+1}}

for a=1,,km,m=1,2,,n1a=1,\cdots,k_{m},m=1,2,\cdots,n-1, where Xi(0)=1+βti,k0=N,kn=0X^{(0)}_{i}=1+\beta t_{i},k_{0}=N,k_{n}=0. Eq.(3.23) implies that Xa(m),a=1,,km,X^{(m)}_{a},a=1,\cdots,k_{m}, are roots of the polynomial l=0km1alxl\sum_{l=0}^{k_{m-1}}a_{l}x^{l}, where

akm1l=\displaystyle a_{k_{m-1}-l}= (1)lel(X(m1))ekm(X(m))\displaystyle(-1)^{l}e_{l}(X^{(m-1)})e_{k_{m}}(X^{(m)})
+(1)km1km+1+lqmβkm1km+1ekm1(X(m1))elkm1+km(X(m+1)).\displaystyle+(-1)^{k_{m-1}-k_{m+1}+l}q_{m}\beta^{k_{m-1}-k_{m+1}}e_{k_{m-1}}(X^{(m-1)})e_{l-k_{m-1}+k_{m}}(X^{(m+1)}).

Assume the roots of l=0km1alxl\sum_{l=0}^{k_{m-1}}a_{l}x^{l} are Xa(m),a=1,,km,X^{(m)}_{a},a=1,\cdots,k_{m}, and Yb(m),b=1,,km1kmY^{(m)}_{b},b=1,\cdots,k_{m-1}-k_{m}, then Vieta’s formula yields

ekm(X(m))i+j=lei(X(m))ej(Y(m))\displaystyle e_{k_{m}}(X^{(m)})\sum_{i+j=l}e_{i}(X^{(m)})e_{j}(Y^{(m)}) (3.24)
=\displaystyle= ekm(X(m))el(X(m1))+qm(β)km1km+1ekm1(X(m1))elkm1+km(X(m+1)).\displaystyle e_{k_{m}}(X^{(m)})e_{l}(X^{(m-1)})+q_{m}(-\beta)^{k_{m-1}-k_{m+1}}e_{k_{m-1}}(X^{(m-1)})e_{l-k_{m-1}+k_{m}}(X^{(m+1)}).

Taking l=km1l=k_{m-1} in Eq.(3.24), we get

ekm1(X(m1))=ekm(X(m))ekm1km(Y(m)).e_{k_{m-1}}(X^{(m-1)})=e_{k_{m}}(X^{(m)})e_{k_{m-1}-k_{m}}(Y^{(m)}). (3.25)

Eq.(3.24) and (3.25) together give us the relation

i+j=lei(X(m))ej(Y(m))=el(X(m1))+qm(β)km1km+1elkm1+km(X(m+1))ekm1km(Y(m)).\sum_{i+j=l}e_{i}(X^{(m)})e_{j}(Y^{(m)})=e_{l}(X^{(m-1)})+q_{m}(-\beta)^{k_{m-1}-k_{m+1}}e_{l-k_{m-1}+k_{m}}(X^{(m+1)})e_{k_{m-1}-k_{m}}(Y^{(m)}). (3.26)

In the case β=1\beta=-1, (3.26) was interpreted as the quantum Whitney relation of the equivariant quantum K-theory ring of the flag variety in [12]. The idea is to identify ei(X(m))e_{i}(X^{(m)}) with i𝒮nm\wedge^{i}\mathcal{S}_{n-m} and make the following identification

el(Y(m))={l(𝒮nm+1/𝒮nm),l<km1km,11qmdet(𝒮nm+1/𝒮nm),l=km1km.e_{l}(Y^{(m)})=\left\{\begin{array}[]{ll}\wedge^{l}(\mathcal{S}_{n-m+1}/\mathcal{S}_{n-m}),&l<k_{m-1}-k_{m},\\ \frac{1}{1-q_{m}}\det(\mathcal{S}_{n-m+1}/\mathcal{S}_{n-m}),&l=k_{m-1}-k_{m}.\end{array}\right.

Then Eq.(3.26) for β=1\beta=-1 can be written in terms of λy\lambda_{y} class as555For a vector bundle \mathcal{E}, the λy\lambda_{y} class is defined as λy()=i=0rank()yii\lambda_{y}(\mathcal{E})=\sum_{i=0}^{\mathrm{rank}(\mathcal{E})}y^{i}\wedge^{i}\mathcal{E} in the Grothendieck group.

λy(𝒮i)λy(𝒮i+1/𝒮i)=λy(𝒮i+1)+qni1qniykni1kni(λy(𝒮i1)λy(𝒮i))det(𝒮i+1/𝒮i)\lambda_{y}(\mathcal{S}_{i})*\lambda_{y}(\mathcal{S}_{i+1}/\mathcal{S}_{i})=\lambda_{y}(\mathcal{S}_{i+1})+\frac{q_{n-i}}{1-q_{n-i}}y^{k_{n-i-1}-k_{n-i}}\left(\lambda_{y}(\mathcal{S}_{i-1})-\lambda_{y}(\mathcal{S}_{i})\right)*\det(\mathcal{S}_{i+1}/\mathcal{S}_{i}) (3.27)

for i=1,2,,n1i=1,2,\cdots,n-1, which are the quantum Whitney relations of the equivariant quantum K-theory ring QKT(Fl(kn1,,k2,k1;N))QK_{T}(\mathrm{Fl}(k_{n-1},\cdots,k_{2},k_{1};N)) proposed in [12].

Remark 1.

Notice that in [12], the analogue of the Bethe ansatz equations were derived from the 3d 𝒩=2\mathcal{N}=2 gauged linear sigma model (GLSM) compactified on a circle. The GLSMs for the flag varieties are quiver gauge theories, whose quiver diagrams have the following form in our convention [14]

kn1\textstyle{k_{n-1}}kn2\textstyle{k_{n-2}}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k1\textstyle{~{}k_{1~{}}}N\textstyle{N}

where circles represent U(ki)U(k_{i}) gauge groups, square represents U(N)U(N) global symmetry and arrows represent chiral fields in bifundamental representations. The Higgs branch of the GLSM is a nonlinear sigma model with target space being the flag variety, the equations for vacuum states on the Coulomb branch are exactly the Bethe ansatz equations we obtained above in the case of β=1\beta=-1. Though ring relations of quantum K-theory ring of general partial flag varieties have not been rigorously established, it was checked in [12] that the ring relations obtained from the GLSMs match those of the special cases, such as incidence varieties and full flag varieties, in which exact results have been proved.

Remark 2.

Theorem 3.1 shows that the Bethe ansatz equations of the quantum integrable model defined by the R-matrix (2.3) and monodromy matrix (2.17) reduce to the vacuum equations on the Coulomb branch of the 2d GLSM for flag varieties when β=0\beta=0, and reduce to the vacuum equations on the Coulomb branch of the 3d GLSM for flag varieties when β=1\beta=-1. Therefore, we have established a Bethe/Gauge correspondence in the sense of [15, 16].

4 Relationship with the double β\beta-Grothendieck polynomials

In this section, we consider the case of full flag variety, i.e. n=Nn=N and kNi=ik_{N-i}=i for i=1,,N1i=1,\cdots,N-1. We will show that the Bethe ansatz states generate the double β\beta-Grothendieck polynomials (Theorem 4.6 and 4.11). The double β\beta-Grothendieck polynomials 𝒢w(β)(𝒙;𝒚)\mathcal{G}^{(\beta)}_{w}(\bm{x};\bm{y}) for wSNw\in S_{N} and 𝒙=(x1,,xN),𝒚=(y1,,yN)\bm{x}=(x_{1},\cdots,x_{N}),\bm{y}=(y_{1},\cdots,y_{N}) are defined in the appendix, see Eq.(A.1)-(A.3).

In addition to Eq.(2.2), let us also define

t:=t1+βt\ominus t:=\frac{-t}{1+\beta t} (4.1)

and 𝒕:=(t1,,tN)\ominus\bm{t}:=(\ominus t_{1},\cdots,\ominus t_{N}) for 𝒕=(t1,,tN)\bm{t}=(t_{1},\cdots,t_{N}). We use ωN\omega_{N} to denote the permutation of NN elements with maximal length, and l(w)l(w) denotes the length of the permutation ww.

From Eq.(2.4), (2.17) and the fact that Bk(u)=[Ta(n)(u)]0,kB_{k}(u)=[{T^{(n)}_{a}(u)}]_{0,k} (Eq.(2.22)), it is easy to compute

Bk(u)|n1,n2,,nN=\displaystyle B_{k}(u)\ket{n_{1},n_{2},\cdots,n_{N}}= (4.2)
=m1=0N1mN1=0N1RaN(u,tN)0,mN1Ra2(u,t2)m2,m1Ra1(u,t1)m1,k|n1,n2,,nN\displaystyle=\sum_{m_{1}=0}^{N-1}\cdots\sum_{m_{N-1}=0}^{N-1}{R_{aN}(u,t_{N})}_{0,m_{N-1}}\cdots{R_{a2}(u,t_{2})}_{m_{2},m_{1}}{R_{a1}(u,t_{1})}_{m_{1},k}\ket{n_{1},n_{2},\cdots,n_{N}}
=m1=0N1mN1=0N1i=1N[δmi,ni(1+δmi>mi1β(uti))|mi1+δmi>niδmi,mi1(uti)|ni],\displaystyle=\sum_{m_{1}=0}^{N-1}\cdots\sum_{m_{N-1}=0}^{N-1}\bigotimes_{i=1}^{N}\left[\delta_{m_{i},n_{i}}(1+\delta_{m_{i}>m_{i-1}}\beta(u\ominus t_{i}))\ket{m_{i-1}}+\delta_{m_{i}>n_{i}}\delta_{m_{i},m_{i-1}}(u\ominus t_{i})\ket{n_{i}}\right],

where m0=k,mN=0m_{0}=k,m_{N}=0. The state (4.2) vanishes unless minim_{i}\geq n_{i} for all 1iN1\leq i\leq N, therefore it vanishes when nN0n_{N}\neq 0. Also, mi=mi1m_{i}=m_{i-1} if mi>nim_{i}>n_{i}.

Refer to caption
Figure 1: The left action of Bk(u)B_{k}(u) on |n1,n2,,nN\ket{n_{1},n_{2},\cdots,n_{N}} for N=12N=12 and a specific choice of (m1,,mN1)(m_{1},\cdots,m_{N-1}) in the expansion (4.2).

The left action of Bk(u)B_{k}(u) on |n1,n2,,nN\ket{n_{1},n_{2},\cdots,n_{N}} can then be depicted as Fig.1 (for a specific (m1,,mN1)(m_{1},\cdots,m_{N-1})). We see for any 𝒎=(m1,,mN)\bm{m}=(m_{1},\cdots,m_{N}) in the expansion (4.2), the effect of the operator Bk(u)B_{k}(u) is to annihilate a |0\ket{0} state, create a |k\ket{k} state, and permute nin_{i} among the sites according to 𝒎\bm{m}. If we write the expansion (4.2) as Bk(u)|𝒏=𝒎C𝒎|ψ𝒎B_{k}(u)\ket{\bm{n}}=\sum_{\bm{m}}C_{\bm{m}}\ket{\psi_{\bm{m}}}, then for any ni<min_{i}<m_{i}, there is a factor (uti)(u\ominus t_{i}) in C𝒎C_{\bm{m}}, and for any mi>mi1m_{i}>m_{i-1}, there is a factor (1+β(uti))(1+\beta(u\ominus t_{i})) in C𝒎C_{\bm{m}}.

For wSNw\in S_{N}, let us define

|w:=|w(012N1),\ket{w}:=\ket{w(012\cdots N-1)}, (4.3)

then, from (4.2) and the discussion above, we have the following

Lemma 4.1.

For any πSN\pi\in S_{N}, we have the expansion

Bπ(1)(u1)Bπ(2)(u2)Bπ(N1)(uN1)|Ω(0)=wSNCwπ(u1,,uN1)|w,B_{\pi(1)}(u_{1})B_{\pi(2)}(u_{2})\cdots B_{\pi(N-1)}(u_{N-1})\ket{\Omega^{(0)}}=\sum_{w\in S_{N}}C^{\pi}_{w}(u_{1},\cdots,u_{N-1})\ket{w},

where Cwπ[u1,,uN1,t1,,tN]C^{\pi}_{w}\in\mathbb{C}[u_{1},\cdots,u_{N-1},t_{1},\cdots,t_{N}] and |Ω(0)\ket{\Omega^{(0)}} is defined by Eq.(3.2).

Similarly, the right action of Bk(u)B_{k}(u) on the dual state n1,n2,,nN|\bra{n_{1},n_{2},\cdots,n_{N}} is

n1,n2,,nN|Bk(u)\displaystyle\bra{n_{1},n_{2},\cdots,n_{N}}B_{k}(u) (4.4)
=m2=0N1mN=0N1n1,n2,,nN|RaN(u,tN)0,mNRa2(u,t2)m3,m2Ra1(u,t1)m2,k\displaystyle=\sum_{m_{2}=0}^{N-1}\cdots\sum_{m_{N}=0}^{N-1}\bra{n_{1},n_{2},\cdots,n_{N}}{R_{aN}(u,t_{N})}_{0,m_{N}}\cdots{R_{a2}(u,t_{2})}_{m_{3},m_{2}}{R_{a1}(u,t_{1})}_{m_{2},k}
=m2=0N1mN=0N1i=1N[δmi,ni(1+δmi<mi+1β(uti))mi+1|+δmi>niδmi,mi+1(uti)ni|],\displaystyle=\sum_{m_{2}=0}^{N-1}\cdots\sum_{m_{N}=0}^{N-1}\bigotimes_{i=1}^{N}\left[\delta_{m_{i},n_{i}}(1+\delta_{m_{i}<m_{i+1}}\beta(u\ominus t_{i}))\bra{m_{i+1}}+\delta_{m_{i}>n_{i}}\delta_{m_{i},m_{i+1}}(u\ominus t_{i})\bra{n_{i}}\right],

where m1=k,mN+1=0m_{1}=k,m_{N+1}=0. The state (4.4) vanishes unless minim_{i}\geq n_{i} for all 1iN1\leq i\leq N and there is some nan_{a} such that na=k=man_{a}=k=m_{a} and mj>njm_{j}>n_{j} for all j<aj<a. Moreover, mi=mi+1m_{i}=m_{i+1} if mi>nim_{i}>n_{i}.

The right action of Bk(u)B_{k}(u) on n1,n2,,nN|\bra{n_{1},n_{2},\cdots,n_{N}} can therefore be depicted as Fig.2 (for a specific choice of (m2,,mN)(m_{2},\cdots,m_{N})). We see for any 𝒎\bm{m} in the expansion (4.4), the effect of the operator Bk(u)B_{k}(u) is to annihilate a k|\bra{k} state, create a 0|\bra{0} state on the NNth site, and permute nin_{i} among the sites according to 𝒎\bm{m}. If we write the expansion (4.4) as 𝒏|Bk(u)=𝒎ψ𝒎|C𝒎\bra{\bm{n}}B_{k}(u)=\sum_{\bm{m}}\bra{\psi_{\bm{m}}}C_{\bm{m}}, then for any ni<min_{i}<m_{i}, there is a factor (uti)(u\ominus t_{i}) in C𝒎C_{\bm{m}}, and there is a factor (1+β(uti))(1+\beta(u\ominus t_{i})) in C𝒎C_{\bm{m}} for any mi<mi+1m_{i}<m_{i+1}.

Refer to caption
Figure 2: The right action of Bk(u)B_{k}(u) on n1,n2,,nN|\bra{n_{1},n_{2},\cdots,n_{N}} for N=12N=12 and a specific choice of (m2,,mN)(m_{2},\cdots,m_{N}) in the expansion (4.4).

For wSNw\in S_{N}, define

w|:=w(012N1)|,\bra{w}:=\bra{w(012\cdots N-1)}, (4.5)

then, from the expansion Eq.(4.4), one can prove

Lemma 4.2.

If (n1,n2,,nN1)(n_{1},n_{2},\cdots,n_{N-1}) satisfies: (i) 0ni<k0\leq n_{i}<k; (ii) nl+1nl+2nN1n_{l+1}\geq n_{l+2}\geq\cdots\geq n_{N-1} for some l0l\geq 0, then

nπ(1),nπ(2),,nπ(l),k,nl+1,,nN1|Bk(u)=i=1l(uti)nπ(1),,nπ(l),nl+1,,nN1,0|\bra{n_{\pi(1)},n_{\pi(2)},\cdots,n_{\pi(l)},k,n_{l+1},\cdots,n_{N-1}}B_{k}(u)=\prod_{i=1}^{l}(u\ominus t_{i})\bra{n_{\pi(1)},\cdots,n_{\pi(l)},n_{l+1},\cdots,n_{N-1},0}

for any πSl\pi\in S_{l}.

Proof.

Given the constraints on nin_{i}, the only possible (m2,m3,,mN)(m_{2},m_{3},\cdots,m_{N}) in the expansion (4.4) is

mi={k,2il+1,ni,l+2iN.m_{i}=\left\{\begin{array}[]{ll}k,&2\leq i\leq l+1,\\ n_{i},&l+2\leq i\leq N.\end{array}\right.

Since nπ(i)<min_{\pi(i)}<m_{i} for 1il1\leq i\leq l, we have the factor i=1l(uti)\prod_{i=1}^{l}(u\ominus t_{i}) from (4.4). ∎

Let us define

ρk:=s1s2skSN,\rho_{k}:=s_{1}s_{2}\cdots s_{k}\in S_{N}, (4.6)

where sis_{i} is the simple transposition of the iith and (i+1)(i+1)th elements, and ρ0\rho_{0} is defined to be the identity permutation. Let iβ\partial^{\beta}_{i} be the β\beta-divided difference operator defined by Eq.(A.1):

iβf(x1,,xN):=(1+βxi+1)f(x1,,xN)(1+βxi)f(x1,,xi+1,xi,,xN)xixi+1.\partial^{\beta}_{i}f(x_{1},\cdots,x_{N}):=\frac{(1+\beta x_{i+1})f(x_{1},\cdots,x_{N})-(1+\beta x_{i})f(x_{1},\cdots,x_{i+1},x_{i},\cdots,x_{N})}{x_{i}-x_{i+1}}.

Generally, for w=si1si2sil(w)w=s_{i_{1}}s_{i_{2}}\cdots s_{i_{l(w)}}, we have the definition666In the following, if a polynomial f(𝒙;𝒚)f(\bm{x};\bm{y}) depends on two sets of variables 𝒙\bm{x} and 𝒚\bm{y} as in the case of double β\beta-Grothendieck polynomials, wβ\partial^{\beta}_{w} in wβf(𝒙;𝒚)\partial^{\beta}_{w}f(\bm{x};\bm{y}) is understood to act on the first set of variables, see Eq.(A.1). The same convention applies to the operator ¯wβ\bar{\partial}^{\beta}_{w} defined by Eq.(4.31).

wβ:=i1βi2βil(w)β.\partial^{\beta}_{w}:=\partial^{\beta}_{i_{1}}\partial^{\beta}_{i_{2}}\cdots\partial^{\beta}_{i_{l(w)}}. (4.7)

Now we can use Lemma 4.2 to prove the following

Lemma 4.3.
1,2,,k,0,k+1,,N1|BN1(σ1)B1(σN1)=𝒢ρk1ωN(β)(𝝈;𝒕)Ω(0)|.\bra{1,2,\cdots,k,0,k+1,\cdots,N-1}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})=\mathcal{G}^{(\beta)}_{\rho^{-1}_{k}\omega_{N}}(\bm{\sigma};\ominus\bm{t})\bra{\Omega^{(0)}}.
Proof.

By repeated use of Lemma 4.2, one can compute

1,2,,k,0,k+1,,N1|BN1(σ1)B1(σN1)\displaystyle\bra{1,2,\cdots,k,0,k+1,\cdots,N-1}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1}) (4.8)
=\displaystyle= a=1Nk1i=1Na(σati)b=NkN2j=1Nb1(σbtj)Ω(0)|.\displaystyle\prod_{a=1}^{N-k-1}\prod_{i=1}^{N-a}(\sigma_{a}\ominus t_{i})\prod_{b=N-k}^{N-2}\prod_{j=1}^{N-b-1}(\sigma_{b}\ominus t_{j})\bra{\Omega^{(0)}}.

Notice that σt=σt1+βt=σ+t+βσ(t)\sigma\ominus t=\frac{\sigma-t}{1+\beta t}=\sigma+\ominus t+\beta\sigma(\ominus t), Eq.(A.2) then yields

i+jN(σitj)=i+jN(σi+tj+βσi(tj))=𝒢ωN(β)(𝝈;𝒕).\prod_{i+j\leq N}(\sigma_{i}\ominus t_{j})=\prod_{i+j\leq N}(\sigma_{i}+\ominus t_{j}+\beta\sigma_{i}(\ominus t_{j}))=\mathcal{G}^{(\beta)}_{\omega_{N}}(\bm{\sigma};\ominus\bm{t}).

Therefore, from (4.8) and the definition of the double β\beta-Grothendieck polynomials Eq.(A.1)-(A.3), we get

1,2,,k,0,k+1,,N1|BN1(σ1)B1(σN1)\displaystyle\bra{1,2,\cdots,k,0,k+1,\cdots,N-1}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})
=\displaystyle= 𝒢ωN(β)(𝝈;𝒕)(σN1t1)(σN2t2)(σNktk)Ω(0)|\displaystyle\frac{\mathcal{G}^{(\beta)}_{\omega_{N}}(\bm{\sigma};\ominus\bm{t})}{(\sigma_{N-1}\ominus t_{1})(\sigma_{N-2}\ominus t_{2})\cdots(\sigma_{N-k}\ominus t_{k})}\bra{\Omega^{(0)}}
=\displaystyle= N1βN2βNkβ𝒢ωN(β)(𝝈;𝒕)Ω(0)|\displaystyle\partial^{\beta}_{N-1}\partial^{\beta}_{N-2}\cdots\partial^{\beta}_{N-k}\mathcal{G}^{(\beta)}_{\omega_{N}}(\bm{\sigma};\ominus\bm{t})\bra{\Omega^{(0)}}
=\displaystyle= 𝒢ωNsNksN2sN1(β)(𝝈;𝒕)Ω(0)|\displaystyle\mathcal{G}^{(\beta)}_{\omega_{N}s_{N-k}\cdots s_{N-2}s_{N-1}}(\bm{\sigma};\ominus\bm{t})\bra{\Omega^{(0)}}
=\displaystyle= 𝒢sks2s1ωN(β)(𝝈;𝒕)Ω(0)|=𝒢ρk1ωN(β)(𝝈;𝒕)Ω(0)|,\displaystyle\mathcal{G}^{(\beta)}_{s_{k}\cdots s_{2}s_{1}\omega_{N}}(\bm{\sigma};\ominus\bm{t})\bra{\Omega^{(0)}}=\mathcal{G}^{(\beta)}_{\rho^{-1}_{k}\omega_{N}}(\bm{\sigma};\ominus\bm{t})\bra{\Omega^{(0)}},

where in the second to last step, we used ωNsNi=siωN\omega_{N}s_{N-i}=s_{i}\omega_{N}. ∎

Lemma 4.4.

If πSN\pi\in S_{N} and l(siπ)>l(π)l(s_{i}\pi)>l(\pi), then

(siπ|+βπ|)BN1(σ1)BN2(σ2)Bi+1(σNi1)Bi(σNi)Bi1(σNi+1)\displaystyle(\bra{s_{i}\pi}+\beta\bra{\pi})B_{N-1}(\sigma_{1})B_{N-2}(\sigma_{2})\cdots B_{i+1}(\sigma_{N-i-1})B_{i}(\sigma_{N-i})B_{i-1}(\sigma_{N-i+1})
=\displaystyle= π|BN1(σ1)BN2(σ2)Bi+1(σNi1)Bi1(σNi)Bi(σNi+1).\displaystyle\bra{\pi}B_{N-1}(\sigma_{1})B_{N-2}(\sigma_{2})\cdots B_{i+1}(\sigma_{N-i-1})B_{i-1}(\sigma_{N-i})B_{i}(\sigma_{N-i+1}).
Proof.

First we show, for

ψ|=n1,,na1,i1,na+1,,nb1,i,nb+1,,nN|\bra{\psi}=\bra{n_{1},\cdots,n_{a-1},i-1,n_{a+1},\cdots,n_{b-1},i,n_{b+1},\cdots,n_{N}} (4.9)

and

siψ|=n1,,na1,i,na+1,,nb1,i1,nb+1,,nN|,\bra{s_{i}\psi}=\bra{n_{1},\cdots,n_{a-1},i,n_{a+1},\cdots,n_{b-1},i-1,n_{b+1},\cdots,n_{N}}, (4.10)

where ns<i1n_{s}<i-1 for sa,bs\neq a,b, we have

(siψ|+βψ|)Bi(x)Bi1(y)=ψ|Bi1(x)Bi(y).(\bra{s_{i}\psi}+\beta\bra{\psi})B_{i}(x)B_{i-1}(y)=\bra{\psi}B_{i-1}(x)B_{i}(y). (4.11)

We have the following expansions in the natural basis according to (4.4)

ψ|Bi1(x)=𝒎Mi1C𝒎(i1)ψ𝒎(i1)|,\displaystyle\bra{\psi}B_{i-1}(x)=\sum_{\bm{m}\in M_{i-1}}C_{\bm{m}}^{(i-1)}\bra{\psi^{(i-1)}_{\bm{m}}},
ψ|Bi(x)=𝒎MiC𝒎(i)ψ𝒎(i)|,\displaystyle\bra{\psi}B_{i}(x)=\sum_{\bm{m}\in M_{i}}C_{\bm{m}}^{(i)}\bra{\psi^{(i)}_{\bm{m}}},
siψ|Bi(x)=𝒎M¯iC¯𝒎(i)ψ¯𝒎(i)|,\displaystyle\bra{s_{i}\psi}B_{i}(x)=\sum_{\bm{m}\in\bar{M}_{i}}\bar{C}^{(i)}_{\bm{m}}\bra{\bar{\psi}^{(i)}_{\bm{m}}},

where Mi1,MiM_{i-1},M_{i} and M¯i\bar{M}_{i} are the sets in which 𝒎=(m1,,mN)\bm{m}=(m_{1},\cdots,m_{N}) takes value such that the corresponding coefficients C𝒎(i1)C_{\bm{m}}^{(i-1)}, C𝒎(i)C_{\bm{m}}^{(i)} and C¯𝒎(i)\bar{C}^{(i)}_{\bm{m}} in (4.4) are nonzero, and ψ𝒎(i1)|\bra{\psi^{(i-1)}_{\bm{m}}}, ψ𝒎(i)|\bra{\psi^{(i)}_{\bm{m}}} and ψ¯𝒎(i)|\bra{\bar{\psi}^{(i)}_{\bm{m}}} are members of the natural basis (3.1). Because ma=i1,mb=im_{a}=i-1,m_{b}=i for 𝒎Mi1\bm{m}\in M_{i-1} and ma=i,mb=i1m_{a}=i,m_{b}=i-1 for 𝒎M¯i\bm{m}\in\bar{M}_{i}, there is a map ϕ:Mi1M¯i\phi:M_{i-1}\rightarrow\bar{M}_{i} defined by

ϕ(𝒎)j={i,ifmj=i1,i1,ifmj=i,mj,ifmjiandmji1.\phi(\bm{m})_{j}=\left\{\begin{array}[]{ll}i,&\mathrm{if}~{}m_{j}=i-1,\\ i-1,&\mathrm{if}~{}m_{j}=i,\\ m_{j},&\mathrm{if}~{}m_{j}\neq i~{}\mathrm{and}~{}m_{j}\neq i-1.\end{array}\right.

ϕ\phi is a bijection since ns<i1n_{s}<i-1 for sa,bs\neq a,b. Moreover, for 𝒎Mi\bm{m}\in M_{i}, we have nj<mjn_{j}<m_{j} for 1j<b1\leq j<b and mb=im_{b}=i, therefore we have an injective map τ:MiM¯i\tau:M_{i}\rightarrow\bar{M}_{i} defined by

τ(𝒎)j={i,1ja,i1,a+1jb,mj,j>b.\tau(\bm{m})_{j}=\left\{\begin{array}[]{ll}i,&1\leq j\leq a,\\ i-1,&a+1\leq j\leq b,\\ m_{j},&j>b.\end{array}\right.

For 𝒎imτ\bm{m}\in\mathrm{im}\tau, i.e. ms=i,mt=i1m_{s}=i,m_{t}=i-1 for 1sa,a+1tb1\leq s\leq a,a+1\leq t\leq b, since na=ma>ma+1n_{a}=m_{a}>m_{a+1} and ϕ1(𝒎)a<ϕ1(𝒎)a+1,τ1(𝒎)a>na\phi^{-1}(\bm{m})_{a}<\phi^{-1}(\bm{m})_{a+1},\tau^{-1}(\bm{m})_{a}>n_{a}, Eq.(4.4) implies

Cϕ1(𝒎)(i1)=(1+β(xta))C¯𝒎(i),Cτ1(𝒎)(i)=(xta)C¯𝒎(i).C^{(i-1)}_{\phi^{-1}(\bm{m})}=(1+\beta(x\ominus t_{a}))\bar{C}^{(i)}_{\bm{m}},\quad C^{(i)}_{\tau^{-1}(\bm{m})}=(x\ominus t_{a})\bar{C}^{(i)}_{\bm{m}}. (4.12)

On the other hand, if 𝒎imτ\bm{m}\not\in\mathrm{im}\tau, then ma+1<i1m_{a+1}<i-1, therefore

Cϕ1(𝒎)(i1)=C¯𝒎(i).C^{(i-1)}_{\phi^{-1}(\bm{m})}=\bar{C}^{(i)}_{\bm{m}}. (4.13)

Consequently,

ψ|Bi1(x)=ϕ1(𝒎)Mi1Cϕ1(𝒎)(i1)ψϕ1(𝒎)(i1)|\displaystyle\bra{\psi}B_{i-1}(x)=\sum_{\phi^{-1}(\bm{m})\in M_{i-1}}C_{\phi^{-1}(\bm{m})}^{(i-1)}\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}} (4.14)
=𝒎imτ(1+β(xta))C¯𝒎(i)ψϕ1(𝒎)(i1)|+𝒎imτC¯𝒎(i)ψϕ1(𝒎)(i1)|,\displaystyle\quad\quad\quad\quad\quad=\sum_{\bm{m}\in\mathrm{im}\tau}(1+\beta(x\ominus t_{a}))\bar{C}^{(i)}_{\bm{m}}\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}}+\sum_{\bm{m}\not\in\mathrm{im}\tau}\bar{C}^{(i)}_{\bm{m}}\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}},
ψ|Bi(x)=𝒎¯MiC𝒎¯(i)ψ𝒎¯(i)|=𝒎¯Mi(xta)C¯τ(𝒎¯)(i)ψ𝒎¯(i)|=𝒎imτ(xta)C¯𝒎(i)ψτ1(𝒎)(i)|,\displaystyle\bra{\psi}B_{i}(x)=\sum_{\bar{\bm{m}}\in M_{i}}C^{(i)}_{\bar{\bm{m}}}\bra{\psi^{(i)}_{\bar{\bm{m}}}}=\sum_{\bar{\bm{m}}\in M_{i}}(x\ominus t_{a})\bar{C}^{(i)}_{\tau(\bar{\bm{m}})}\bra{\psi^{(i)}_{\bar{\bm{m}}}}=\sum_{\bm{m}\in\mathrm{im}\tau}(x\ominus t_{a})\bar{C}^{(i)}_{\bm{m}}\bra{\psi^{(i)}_{\tau^{-1}(\bm{m})}},
siψ|Bi(x)=𝒎M¯iC¯𝒎(i)ψ¯𝒎(i)|=𝒎imτC¯𝒎(i)ψ¯𝒎(i)|+𝒎imτC¯𝒎(i)ψ¯𝒎(i)|,\displaystyle\bra{s_{i}\psi}B_{i}(x)=\sum_{\bm{m}\in\bar{M}_{i}}\bar{C}_{\bm{m}}^{(i)}\bra{\bar{\psi}^{(i)}_{\bm{m}}}=\sum_{\bm{m}\in\mathrm{im}\tau}\bar{C}^{(i)}_{\bm{m}}\bra{\bar{\psi}^{(i)}_{\bm{m}}}+\sum_{\bm{m}\not\in\mathrm{im}\tau}\bar{C}^{(i)}_{\bm{m}}\bra{\bar{\psi}^{(i)}_{\bm{m}}},

where, according to (4.4),

ψϕ1(𝒎)(i1)|=n1,,na1,i,na+1,,nb1,ns1,,nsNb+1|,\displaystyle\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}}=\bra{n_{1},\cdots,n_{a-1},i,n_{a+1},\cdots,n_{b-1},n^{\prime}_{s_{1}},\cdots,n^{\prime}_{s_{N-b+1}}},
ψτ1(𝒎)(i)|=n1,,na1,i1,na+1,,nb1,ns1,,nsNb+1|,\displaystyle\bra{\psi^{(i)}_{\tau^{-1}(\bm{m})}}=\bra{n_{1},\cdots,n_{a-1},i-1,n_{a+1},\cdots,n_{b-1},n^{\prime}_{s_{1}},\cdots,n^{\prime}_{s_{N-b+1}}},
ψ¯𝒎(i)|=n1,,na1,i1,na+1,,nb1,ns1,,nsNb+1|,\displaystyle\bra{\bar{\psi}^{(i)}_{\bm{m}}}=\bra{n_{1},\cdots,n_{a-1},i-1,n_{a+1},\cdots,n_{b-1},n^{\prime}_{s_{1}},\cdots,n^{\prime}_{s_{N-b+1}}},

with some ns1,,nsNb+1n^{\prime}_{s_{1}},\cdots,n^{\prime}_{s_{N-b+1}} determined by 𝒎\bm{m}. From the equations above, we see
ψτ1(𝒎)(i)|=ψ¯𝒎(i)|\bra{\psi^{(i)}_{\tau^{-1}(\bm{m})}}=\bra{\bar{\psi}^{(i)}_{\bm{m}}} and ψϕ1(𝒎)(i1)|Bi(y)=ψ¯𝒎(i)|Bi1(y)\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}}B_{i}(y)=\bra{\bar{\psi}^{(i)}_{\bm{m}}}B_{i-1}(y), from which, together with Eq.(4.14), one can compute

(siψ|+βψ|)Bi(x)Bi1(y)\displaystyle(\bra{s_{i}\psi}+\beta\bra{\psi})B_{i}(x)B_{i-1}(y)
=\displaystyle= 𝒎imτC¯𝒎(i)ψ¯𝒎(i)|Bi1(y)+𝒎imτC¯𝒎(i)ψ¯𝒎(i)|Bi1(y)+β𝒎imτ(xta)C¯𝒎(i)ψ¯𝒎(i)|Bi1(y)\displaystyle\sum_{\bm{m}\not\in\mathrm{im}\tau}\bar{C}^{(i)}_{\bm{m}}\bra{\bar{\psi}^{(i)}_{\bm{m}}}B_{i-1}(y)+\sum_{\bm{m}\in\mathrm{im}\tau}\bar{C}^{(i)}_{\bm{m}}\bra{\bar{\psi}^{(i)}_{\bm{m}}}B_{i-1}(y)+\beta\sum_{\bm{m}\in\mathrm{im}\tau}(x\ominus t_{a})\bar{C}^{(i)}_{\bm{m}}\bra{\bar{\psi}^{(i)}_{\bm{m}}}B_{i-1}(y)
=\displaystyle= 𝒎imτC¯𝒎(i)ψϕ1(𝒎)(i1)|Bi(y)+𝒎imτ(1+β(xta))C¯𝒎(i)ψϕ1(𝒎)(i1)|Bi(y)\displaystyle\sum_{\bm{m}\not\in\mathrm{im}\tau}\bar{C}^{(i)}_{\bm{m}}\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}}B_{i}(y)+\sum_{\bm{m}\in\mathrm{im}\tau}(1+\beta(x\ominus t_{a}))\bar{C}^{(i)}_{\bm{m}}\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}}B_{i}(y)
=\displaystyle= ϕ1(𝒎)Mi1Cϕ1(𝒎)(i1)ψϕ1(𝒎)(i1)|Bi(y)\displaystyle\sum_{\phi^{-1}(\bm{m})\in M_{i-1}}C^{(i-1)}_{\phi^{-1}(\bm{m})}\bra{\psi^{(i-1)}_{\phi^{-1}(\bm{m})}}B_{i}(y)
=\displaystyle= ψ|Bi1(x)Bi(y).\displaystyle\bra{\psi}B_{i-1}(x)B_{i}(y).

Thus Eq.(4.11) is proved.

Now assume, for ki+1k\geq i+1 and φ|=n1,,nN|\bra{\varphi}=\bra{n_{1},\cdots,n_{N}} with {n1,,nN}={1,2,,k1,0,,0}\{n_{1},\cdots,n_{N}\}=\{1,2,\cdots,k-1,0,\cdots,0\}, we have

(siφ|+βφ|)Bk1(σ1)Bi(σki)Bi1(σki+1)=φ|Bk1(σ1)Bi1(σki)Bi(σki+1),(\bra{s_{i}\varphi}+\beta\bra{\varphi})B_{k-1}(\sigma_{1})\cdots B_{i}(\sigma_{k-i})B_{i-1}(\sigma_{k-i+1})=\bra{\varphi}B_{k-1}(\sigma_{1})\cdots B_{i-1}(\sigma_{k-i})B_{i}(\sigma_{k-i+1}), (4.15)

where φ|\bra{\varphi} and siφ|\bra{s_{i}\varphi} are defined in the same way as Eq.(4.9) and (4.10). The k=i+1k=i+1 case of (4.15) is valid due to Eq.(4.11). Now let ψ|\bra{\psi} be a state of the form

ψ|=n1,,na1,i1,na+1,,nb1,i,nb+1,,nN|\bra{\psi}=\bra{n_{1},\cdots,n_{a-1},i-1,n_{a+1},\cdots,n_{b-1},i,n_{b+1},\cdots,n_{N}}

with {n1,,nN}={1,2,,k,0,,0}\{n_{1},\cdots,n_{N}\}=\{1,2,\cdots,k,0,\cdots,0\}, where na=i1,nb=in_{a}=i-1,n_{b}=i. We expand the following states in the natural basis according to (4.4):

ψ|Bk(u)=𝒎M1C𝒎ψ𝒎|+𝒎M2C𝒎ψ𝒎|,\displaystyle\bra{\psi}B_{k}(u)=\sum_{\bm{m}\in M_{1}}C_{\bm{m}}\bra{\psi_{\bm{m}}}+\sum_{\bm{m}\in M_{2}}C^{\prime}_{\bm{m}}\bra{\psi^{\prime}_{\bm{m}}}, (4.16)
siψ|Bk(u)=𝒎M¯1C¯𝒎ψ¯𝒎|,\displaystyle\bra{s_{i}\psi}B_{k}(u)=\sum_{\bm{m}\in\bar{M}_{1}}\bar{C}_{\bm{m}}\bra{\bar{\psi}_{\bm{m}}}, (4.17)

where M2M_{2} contains 𝒎\bm{m}’s satisfying i1=na<ma=ma+1==mb=ii-1=n_{a}<m_{a}=m_{a+1}=\cdots=m_{b}=i, and M1M_{1} consists of all the other 𝒎\bm{m}’s such that the corresponding coefficients are nonzero. One important difference between these two sets is that i1i-1 remains to the left of ii in ψ𝒎|\bra{\psi_{\bm{m}}}, while ii is moved to the left of i1i-1 in ψ𝒎|\bra{\psi^{\prime}_{\bm{m}}}. Since it is not possible to have i=na<ma=ma+1==mb=i1i=n_{a}<m_{a}=m_{a+1}=\cdots=m_{b}=i-1 in the expansion (4.17), we have a one-to-one correspondence between M1M_{1} and M¯1\bar{M}_{1} given by ϕ~:M1M¯1\tilde{\phi}:M_{1}\rightarrow\bar{M}_{1} with ϕ~\tilde{\phi} defined by

ϕ~(𝒎)j={i,ifmj=i1,i1,ifmj=i,mj,ifmjiandmji1.\tilde{\phi}(\bm{m})_{j}=\left\{\begin{array}[]{ll}i,&\mathrm{if}~{}m_{j}=i-1,\\ i-1,&\mathrm{if}~{}m_{j}=i,\\ m_{j},&\mathrm{if}~{}m_{j}\neq i~{}\mathrm{and}~{}m_{j}\neq i-1.\end{array}\right.

Since ψ¯ϕ~(𝒎)|\bra{\bar{\psi}_{\tilde{\phi}(\bm{m})}} differs from ψ𝒎|\bra{\psi_{\bm{m}}} by exchange of i1i-1 and ii, it is easy to see

ψ¯ϕ~(𝒎)|=siψ𝒎|.\bra{\bar{\psi}_{\tilde{\phi}(\bm{m})}}=\bra{s_{i}\psi_{\bm{m}}}. (4.18)

Notice that ns<i1n_{s}<i-1 or ns>in_{s}>i for sa,bs\neq a,b, Eq.(4.4) implies

C𝒎={(1+β(uta))C¯ϕ~(𝒎),ifma=i1,mb=i,ns<ms=mbfora<s<b,C¯ϕ~(𝒎),otherwise.C_{\bm{m}}=\left\{\begin{array}[]{ll}(1+\beta(u\ominus t_{a}))\bar{C}_{\tilde{\phi}(\bm{m})},&\mathrm{if}~{}m_{a}=i-1,m_{b}=i,n_{s}<m_{s}=m_{b}~{}\mathrm{for}~{}a<s<b,\\ \bar{C}_{\tilde{\phi}(\bm{m})},&\mathrm{otherwise}.\end{array}\right. (4.19)

Moreover, there is an injective map τ~:M2M1\tilde{\tau}:M_{2}\rightarrow M_{1} given by

τ~(𝒎)j={i1,ifmj=iandja,mj,otherwise.\tilde{\tau}(\bm{m})_{j}=\left\{\begin{array}[]{ll}i-1,&\mathrm{if}~{}m_{j}=i~{}\mathrm{and}~{}j\leq a,\\ m_{j},&\mathrm{otherwise}.\end{array}\right.

Notice that 𝒎imτ~\bm{m}\in\mathrm{im}\tilde{\tau} is equivalent to ma=i1,mb=i,ns<ms=mbm_{a}=i-1,m_{b}=i,n_{s}<m_{s}=m_{b} for a<s<ba<s<b, so Eq.(4.19) can be rewritten as

C𝒎={(1+β(uta))C¯ϕ~(𝒎),if𝒎imτ~,C¯ϕ~(𝒎),otherwise.C_{\bm{m}}=\left\{\begin{array}[]{ll}(1+\beta(u\ominus t_{a}))\bar{C}_{\tilde{\phi}(\bm{m})},&\mathrm{if}~{}\bm{m}\in\mathrm{im}\tilde{\tau},\\ \bar{C}_{\tilde{\phi}(\bm{m})},&\mathrm{otherwise}.\end{array}\right. (4.20)

In addition, if 𝒎imτ~\bm{m}\in\mathrm{im}\tilde{\tau}, then

ψ¯ϕ~(𝒎)|=ψτ~1(𝒎)|,\bra{\bar{\psi}_{\tilde{\phi}(\bm{m})}}=\bra{\psi^{\prime}_{\tilde{\tau}^{-1}(\bm{m})}}, (4.21)

because on both sides, i1i-1 is moved to the aath site, ii is moved to the ccth site, where c=max{l<a|nl=ml}c=\mathrm{max}\{l<a~{}|~{}n_{l}=m_{l}\}, and the other nsn_{s}’s are moved in the same way, and

Cτ~1(𝒎)=(uta)C¯ϕ~(𝒎),C^{\prime}_{\tilde{\tau}^{-1}(\bm{m})}=(u\ominus t_{a})\bar{C}_{\tilde{\phi}(\bm{m})}, (4.22)

because ϕ~(𝒎)a=na=i\tilde{\phi}(\bm{m})_{a}=n_{a}=i, τ~1(𝒎)a>na=i1\tilde{\tau}^{-1}(\bm{m})_{a}>n_{a}=i-1, and ns<i1n_{s}<i-1 or ns>in_{s}>i for sa,bs\neq a,b. Therefore

𝒎M1C¯ϕ~(𝒎)ψ¯ϕ~(𝒎)|+β𝒎M2C𝒎ψ𝒎|\displaystyle\sum_{\bm{m}\in M_{1}}\bar{C}_{\tilde{\phi}(\bm{m})}\bra{\bar{\psi}_{\tilde{\phi}(\bm{m})}}+\beta\sum_{\bm{m}\in M_{2}}C^{\prime}_{\bm{m}}\bra{\psi^{\prime}_{\bm{m}}} (4.23)
=\displaystyle= 𝒎M1\imτ~C¯ϕ~(𝒎)ψ¯ϕ~(𝒎)|+𝒎imτ~C¯ϕ~(𝒎)ψ¯ϕ~(𝒎)|+β𝒎M2C𝒎ψ𝒎|\displaystyle\sum_{\bm{m}\in M_{1}\backslash\mathrm{im}\tilde{\tau}}\bar{C}_{\tilde{\phi}(\bm{m})}\bra{\bar{\psi}_{\tilde{\phi}(\bm{m})}}+\sum_{\bm{m}\in\mathrm{im}\tilde{\tau}}\bar{C}_{\tilde{\phi}(\bm{m})}\bra{\bar{\psi}_{\tilde{\phi}(\bm{m})}}+\beta\sum_{\bm{m}\in M_{2}}C^{\prime}_{\bm{m}}\bra{\psi^{\prime}_{\bm{m}}}
=\displaystyle= 𝒎M1\imτ~C¯ϕ~(𝒎)siψ𝒎|+𝒎imτ~C¯ϕ~(𝒎)siψ𝒎|+β(uta)𝒎imτ~C¯ϕ~(𝒎)siψ𝒎|\displaystyle\sum_{\bm{m}\in M_{1}\backslash\mathrm{im}\tilde{\tau}}\bar{C}_{\tilde{\phi}(\bm{m})}\bra{s_{i}\psi_{\bm{m}}}+\sum_{\bm{m}\in\mathrm{im}\tilde{\tau}}\bar{C}_{\tilde{\phi}(\bm{m})}\bra{s_{i}\psi_{\bm{m}}}+\beta(u\ominus t_{a})\sum_{\bm{m}\in\mathrm{im}\tilde{\tau}}\bar{C}_{\tilde{\phi}(\bm{m})}\bra{s_{i}\psi_{\bm{m}}}
=\displaystyle= 𝒎M1C𝒎siψ𝒎|,\displaystyle\sum_{\bm{m}\in M_{1}}C_{\bm{m}}\bra{s_{i}\psi_{\bm{m}}},

where the second equality is due to Eq.(4.18), (4.21) and (4.22), and in the last equality we used Eq.(4.20). Consequently,

(siψ|+βψ|)Bk(u)\displaystyle(\bra{s_{i}\psi}+\beta\bra{\psi})B_{k}(u) (4.24)
=\displaystyle= 𝒎M1C¯ϕ~(𝒎)ψ¯ϕ~(𝒎)|+β𝒎M2C𝒎ψ𝒎|+β𝒎M1C𝒎ψ𝒎|\displaystyle\sum_{\bm{m}\in M_{1}}\bar{C}_{\tilde{\phi}(\bm{m})}\bra{\bar{\psi}_{\tilde{\phi}(\bm{m})}}+\beta\sum_{\bm{m}\in M_{2}}C^{\prime}_{\bm{m}}\bra{\psi^{\prime}_{\bm{m}}}+\beta\sum_{\bm{m}\in M_{1}}C_{\bm{m}}\bra{\psi_{\bm{m}}}
=\displaystyle= 𝒎M1C𝒎(siψ𝒎|+βψ𝒎|),\displaystyle\sum_{\bm{m}\in M_{1}}C_{\bm{m}}(\bra{s_{i}\psi_{\bm{m}}}+\beta\bra{\psi_{\bm{m}}}),

where the second equality is due to (4.23). From the induction hypothesis Eq.(4.15), we have

(siψ𝒎|+βψ𝒎|)Bk1(σ1)Bi(σki)Bi1(σki+1)=ψ𝒎|Bk1(σ1)Bi1(σki)Bi(σki+1),(\bra{s_{i}\psi_{\bm{m}}}+\beta\bra{\psi_{\bm{m}}})B_{k-1}(\sigma_{1})\cdots B_{i}(\sigma_{k-i})B_{i-1}(\sigma_{k-i+1})=\bra{\psi_{\bm{m}}}B_{k-1}(\sigma_{1})\cdots B_{i-1}(\sigma_{k-i})B_{i}(\sigma_{k-i+1}), (4.25)

which results in

(siψ|+βψ|)Bk(u)Bk1(σ1)Bi(σki)Bi1(σki+1)\displaystyle(\bra{s_{i}\psi}+\beta\bra{\psi})B_{k}(u)B_{k-1}(\sigma_{1})\cdots B_{i}(\sigma_{k-i})B_{i-1}(\sigma_{k-i+1}) (4.26)
=\displaystyle= 𝒎M1C𝒎(siψ𝒎|+βψ𝒎|)Bk1(σ1)Bi(σki)Bi1(σki+1)\displaystyle\sum_{\bm{m}\in M_{1}}C_{\bm{m}}(\bra{s_{i}\psi_{\bm{m}}}+\beta\bra{\psi_{\bm{m}}})B_{k-1}(\sigma_{1})\cdots B_{i}(\sigma_{k-i})B_{i-1}(\sigma_{k-i+1})
=\displaystyle= 𝒎M1C𝒎ψ𝒎|Bk1(σ1)Bi1(σki)Bi(σki+1)\displaystyle\sum_{\bm{m}\in M_{1}}C_{\bm{m}}\bra{\psi_{\bm{m}}}B_{k-1}(\sigma_{1})\cdots B_{i-1}(\sigma_{k-i})B_{i}(\sigma_{k-i+1})
=\displaystyle= ψ|Bk(u)Bk1(σ1)Bi1(σki)Bi(σki+1),\displaystyle\bra{\psi}B_{k}(u)B_{k-1}(\sigma_{1})\cdots B_{i-1}(\sigma_{k-i})B_{i}(\sigma_{k-i+1}),

where the first equality is due to (4.24), the second equality is due to (4.25), and in the last step we used Eq.(4.16) and the fact that, for 𝒎M2\bm{m}\in M_{2}, ii is to the left of i1i-1 in ψ𝒎|\bra{\psi^{\prime}_{\bm{m}}}, so (notice that BlB_{l} with l>il>i cannot move i1i-1 to the left of ii because the only way to do this is to have i<ms=mt=i1i<m_{s}=m_{t}=i-1 for some s<ts<t, which is impossible)

ψ𝒎|Bk1(σ1)Bi1(σki)Bi(σki+1)=0,𝒎M2,\bra{\psi^{\prime}_{\bm{m}}}B_{k-1}(\sigma_{1})\cdots B_{i-1}(\sigma_{k-i})B_{i}(\sigma_{k-i+1})=0,\quad\bm{m}\in M_{2},

because it is impossible to have i<ms=ms+1==mt=i1i<m_{s}=m_{s+1}=\cdots=m_{t}=i-1 when Bi1(σki)B_{i-1}(\sigma_{k-i}) acts on the state.

Combining Eq.(4.11), (4.15) and (4.26), the lemma is proved by induction. ∎

In order to relate the quantum states of the integrable system to the double β\beta-Grothendieck polynomials, we need the commutation relations among the different BiB_{i} operators. From Eq.(2.4) and (2.25), we have the commutation relations

Bi(x)Bi(y)=Bi(y)Bi(x),\displaystyle B_{i}(x)B_{i}(y)=B_{i}(y)B_{i}(x), (4.27)
Bi(x)Bj(y)=Bj(y)Bi(x)Bj(x)Bi(y)yx=(1+βx)Bj(x)Bi(y)Bj(y)Bi(x)xy,i<j.\displaystyle B_{i}(x)B_{j}(y)=\frac{B_{j}(y)B_{i}(x)-B_{j}(x)B_{i}(y)}{y\ominus x}=(1+\beta x)\frac{B_{j}(x)B_{i}(y)-B_{j}(y)B_{i}(x)}{x-y},\quad i<j. (4.28)

Let us define

¯iβ=(1+βxi)i,\bar{\partial}^{\beta}_{i}=(1+\beta x_{i})\partial_{i}, (4.29)

where i\partial_{i} is the ordinary divided difference operator defined by iβ|β=0\partial^{\beta}_{i}|_{\beta=0}. It is easy to check that

¯iβ=iβ+β.\bar{\partial}^{\beta}_{i}=\partial^{\beta}_{i}+\beta. (4.30)

Similar to Eq.(4.7), for w=si1si2sil(w)w=s_{i_{1}}s_{i_{2}}\cdots s_{i_{l(w)}}, we define

¯wβ:=¯i1β¯i2β¯il(w)β.\bar{\partial}^{\beta}_{w}:=\bar{\partial}^{\beta}_{i_{1}}\bar{\partial}^{\beta}_{i_{2}}\cdots\bar{\partial}^{\beta}_{i_{l(w)}}. (4.31)

From Eq.(4.28), (4.29) and Lemma 4.4, we can prove

Lemma 4.5.

For πSN\pi\in S_{N}, if l(siπ)>l(π)l(s_{i}\pi)>l(\pi), then

siπ|BN1(σ1)Bi(σNi)Bi1(σNi+1)=Niβπ|BN1(σ1)Bi(σNi)Bi1(σNi+1).\bra{s_{i}\pi}B_{N-1}(\sigma_{1})\cdots B_{i}(\sigma_{N-i})B_{i-1}(\sigma_{N-i+1})=\partial^{\beta}_{N-i}\bra{\pi}B_{N-1}(\sigma_{1})\cdots B_{i}(\sigma_{N-i})B_{i-1}(\sigma_{N-i+1}).
Proof.
(siπ|+βπ|)BN1(σ1)Bi(σNi)Bi1(σNi+1)\displaystyle(\bra{s_{i}\pi}+\beta\bra{\pi})B_{N-1}(\sigma_{1})\cdots B_{i}(\sigma_{N-i})B_{i-1}(\sigma_{N-i+1})
=\displaystyle= π|BN1(σ1)Bi1(σNi)Bi(σNi+1)\displaystyle\bra{\pi}B_{N-1}(\sigma_{1})\cdots B_{i-1}(\sigma_{N-i})B_{i}(\sigma_{N-i+1})
=\displaystyle= (Niβ+β)π|BN1(σ1)Bi(σNi)Bi1(σNi+1),\displaystyle(\partial^{\beta}_{N-i}+\beta)\bra{\pi}B_{N-1}(\sigma_{1})\cdots B_{i}(\sigma_{N-i})B_{i-1}(\sigma_{N-i+1}),

where the first equality is due to Lemma 4.4, and the second equality is due to Eq.(4.28), (4.29) and (4.30). ∎

Theorem 4.6.

For πSN\pi\in S_{N},

BN1(σ1)BN2(σ2)B1(σN1)|Ω(0)=wSN𝒢w(β)(𝝈;𝒕)|ωNw1,B_{N-1}(\sigma_{1})B_{N-2}(\sigma_{2})\cdots B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}}=\sum_{w\in S_{N}}\mathcal{G}^{(\beta)}_{w}(\bm{\sigma};\ominus\bm{t})\ket{\omega_{N}w^{-1}},

where ωN\omega_{N} is the permutation in SNS_{N} with maximal length.

Proof.

Lemma 4.1 suggests that we only need to show

ωNw1|BN1(σ1)B1(σN1)|Ω(0)=𝒢w(β)(𝝈;𝒕).\bra{\omega_{N}w^{-1}}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}}=\mathcal{G}^{(\beta)}_{w}(\bm{\sigma};\ominus\bm{t}). (4.32)

Lemma 4.3 shows Eq.(4.32) is valid for ωNw1=ρk,k=0,1,,N1\omega_{N}w^{-1}=\rho_{k},k=0,1,\cdots,N-1. Because any permutation wSNw\in S_{N} can be written as w=πρkw=\pi\rho_{k} for some πSN1\pi\in S_{N-1} acting on the last N1N-1 elements and some k=0,,N1k=0,\cdots,N-1, it remains to show the validity of Eq.(4.32) for ωNw1=πρk\omega_{N}w^{-1}=\pi\rho_{k} with πSN1\pi\in S_{N-1} acting on {1,2,,N1}\{1,2,\cdots,N-1\}. We show this by induction on the lenghth of π\pi. Let us assume

π~ρk|BN1(σ1)B2(σN2)B1(σN1)|Ω(0)=𝒢(π~ρk)1ωN(β)(𝝈;𝒕)\bra{\tilde{\pi}\rho_{k}}B_{N-1}(\sigma_{1})\cdots B_{2}(\sigma_{N-2})B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}}=\mathcal{G}^{(\beta)}_{(\tilde{\pi}\rho_{k})^{-1}\omega_{N}}(\bm{\sigma};\ominus\bm{t})

for π~SN1\tilde{\pi}\in S_{N-1} with l(π~)m1l(\tilde{\pi})\leq m-1 for some m1m\geq 1. Now, for πSN1\pi\in S_{N-1} with l(π)=ml(\pi)=m, there exists π~SN1\tilde{\pi}\in S_{N-1} with l(π~)=m1l(\tilde{\pi})=m-1, such that π=siπ~\pi=s_{i}\tilde{\pi}. Then

πρk|BN1(σ1)B1(σN1)|Ω(0)\displaystyle\bra{\pi\rho_{k}}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}}
=\displaystyle= siπ~ρk|BN1(σ1)B1(σN1)|Ω(0)\displaystyle\bra{s_{i}\tilde{\pi}\rho_{k}}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}}
=\displaystyle= Niβπ~ρk|BN1(σ1)B1(σN1)|Ω(0)\displaystyle\partial^{\beta}_{N-i}\bra{\tilde{\pi}\rho_{k}}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1})\ket{\Omega^{(0)}}
=\displaystyle= Niβ𝒢ρk1π~1ωN(β)(𝝈;𝒕)\displaystyle\partial^{\beta}_{N-i}\mathcal{G}^{(\beta)}_{\rho_{k}^{-1}\tilde{\pi}^{-1}\omega_{N}}(\bm{\sigma};\ominus\bm{t})
=\displaystyle= 𝒢ρk1π~1ωNsNi(β)(𝝈;𝒕)=𝒢ρk1π~1siωN(β)(𝝈;𝒕)\displaystyle\mathcal{G}^{(\beta)}_{\rho_{k}^{-1}\tilde{\pi}^{-1}\omega_{N}s_{N-i}}(\bm{\sigma};\ominus\bm{t})=\mathcal{G}^{(\beta)}_{\rho_{k}^{-1}\tilde{\pi}^{-1}s_{i}\omega_{N}}(\bm{\sigma};\ominus\bm{t})
=\displaystyle= 𝒢(πρk)1ωN(β)(𝝈;𝒕),\displaystyle\mathcal{G}^{(\beta)}_{(\pi\rho_{k})^{-1}\omega_{N}}(\bm{\sigma};\ominus\bm{t}),

where the second equality is due to Lemma 4.5. ∎

Lemma 4.7.

For any wSN1w\in S_{N-1},

Bw(1)(σ1)Bw(N1)(σN1)=¯w1ωN1βBN1(σ1)B1(σN1).B_{w(1)}(\sigma_{1})\cdots B_{w(N-1)}(\sigma_{N-1})=\bar{\partial}^{\beta}_{w^{-1}\omega_{N-1}}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1}).
Proof.

Let w=πωN1w=\pi\omega_{N-1}, then

Bw(1)(σ1)Bw(N1)(σN1)=Bπ(N1)(σ1)Bπ(1)(σN1).B_{w(1)}(\sigma_{1})\cdots B_{w(N-1)}(\sigma_{N-1})=B_{\pi(N-1)}(\sigma_{1})\cdots B_{\pi(1)}(\sigma_{N-1}). (4.33)

Let π=si1si2sip\pi=s_{i_{1}}s_{i_{2}}\cdots s_{i_{p}}, where p=l(π)p=l(\pi). Then we can apply the transpositions sNip1,,sNi21s_{N-i_{p}-1},\cdots,s_{N-i_{2}-1}, sNi11s_{N-i_{1}-1} successively to the subindices of the right hand side of (4.33) to bring them to the strictly descending order, where sjs_{j} exchanges the jjth and (j+1)(j+1)th subindices at each step. Since we need at least pp transpositions to complete the ordering, every transposition must bring a smaller subindex to the right of a neighboring larger subindex. According to (4.28), the application of sjs_{j} must be accompanied by the action of ¯jβ\bar{\partial}^{\beta}_{j} to the right hand side of (4.33) in order to keep the equality. Therefore,

Bπ(N1)(σ1)Bπ(1)(σN1)=¯Nip1β¯Ni21β¯Ni11βBN1(σ1)B1(σN1).B_{\pi(N-1)}(\sigma_{1})\cdots B_{\pi(1)}(\sigma_{N-1})=\bar{\partial}^{\beta}_{N-i_{p}-1}\cdots\bar{\partial}^{\beta}_{N-i_{2}-1}\bar{\partial}^{\beta}_{N-i_{1}-1}B_{N-1}(\sigma_{1})\cdots B_{1}(\sigma_{N-1}).

Because of the identity

sNip1sNi21sNi11=ωN1sipsi2si1ωN1=ωN1π1ωN1=w1ωN1,s_{N-i_{p}-1}\cdots s_{N-i_{2}-1}s_{N-i_{1}-1}=\omega_{N-1}s_{i_{p}}\cdots s_{i_{2}}s_{i_{1}}\omega_{N-1}=\omega_{N-1}\pi^{-1}\omega_{N-1}=w^{-1}\omega_{N-1},

the proof is thus complete. ∎

Remark 3.

Lemma 4.7 allows us to deduce the matrix elements of Bw(1)(σ1)Bw(N1)(σN1)B_{w(1)}(\sigma_{1})\cdots B_{w(N-1)}(\sigma_{N-1}) from Theorem 4.6.

Example 1.

For N=3N=3, we have, by Eq.(4.2) or Theorem 4.6,

B2(σ1)B1(σ2)|000\displaystyle B_{2}(\sigma_{1})B_{1}(\sigma_{2})\ket{000}
=\displaystyle= |210+(σ1t1)|120+[(σ1t1)+(σ2t2)+β(σ1t1)(σ2t2)]|201\displaystyle\ket{210}+(\sigma_{1}\ominus t_{1})\ket{120}+[(\sigma_{1}\ominus t_{1})+(\sigma_{2}\ominus t_{2})+\beta(\sigma_{1}\ominus t_{1})(\sigma_{2}\ominus t_{2})]\ket{201}
+(σ1t1)(σ1t2)|102+(σ1t1)(σ2t1)|021+(σ1t1)(σ1t2)(σ2t1)|012\displaystyle+(\sigma_{1}\ominus t_{1})(\sigma_{1}\ominus t_{2})\ket{102}+(\sigma_{1}\ominus t_{1})(\sigma_{2}\ominus t_{1})\ket{021}+(\sigma_{1}\ominus t_{1})(\sigma_{1}\ominus t_{2})(\sigma_{2}\ominus t_{1})\ket{012}
=\displaystyle= 𝒢123(β)(σ1,σ2;t1,t2)|210+𝒢213(β)(σ1,σ2;t1,t2)|120+𝒢132(β)(σ1,σ2;t1,t2)|201\displaystyle\mathcal{G}^{(\beta)}_{123}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})\ket{210}+\mathcal{G}^{(\beta)}_{213}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})\ket{120}+\mathcal{G}^{(\beta)}_{132}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})\ket{201}
+𝒢312(β)(σ1,σ2;t1,t2)|102+𝒢231(β)(σ1,σ2;t1,t2)|021+𝒢321(β)(σ1,σ2;t1,t2)|012,\displaystyle+\mathcal{G}^{(\beta)}_{312}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})\ket{102}+\mathcal{G}^{(\beta)}_{231}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})\ket{021}+\mathcal{G}^{(\beta)}_{321}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})\ket{012},

while

B1(σ1)B2(σ2)|000\displaystyle B_{1}(\sigma_{1})B_{2}(\sigma_{2})\ket{000} (4.34)
=\displaystyle= (1+β(σ1t1))|120+[(σ2t1)(1+β(σ1t2))+(σ1t2)(1+β(σ1t1))]|102\displaystyle(1+\beta(\sigma_{1}\ominus t_{1}))\ket{120}+[(\sigma_{2}\ominus t_{1})(1+\beta(\sigma_{1}\ominus t_{2}))+(\sigma_{1}\ominus t_{2})(1+\beta(\sigma_{1}\ominus t_{1}))]\ket{102}
+(σ1t1)(σ2t1)(1+β(σ1t2))|012.\displaystyle+(\sigma_{1}\ominus t_{1})(\sigma_{2}\ominus t_{1})(1+\beta(\sigma_{1}\ominus t_{2}))\ket{012}.

It is easy to verify that Eq.(4.34) is consistent with Lemma 4.7 since we have

¯1β𝒢213(β)(σ1,σ2;t1,t2)=1+β(σ1t1),\displaystyle\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{213}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=1+\beta(\sigma_{1}\ominus t_{1}),
¯1β𝒢321(β)(σ1,σ2;t1,t2)=(σ1t1)(σ2t1)(1+β(σ1t2)),\displaystyle\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{321}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=(\sigma_{1}\ominus t_{1})(\sigma_{2}\ominus t_{1})(1+\beta(\sigma_{1}\ominus t_{2})),
¯1β𝒢312(β)(σ1,σ2;t1,t2)=(σ2t1)(1+β(σ1t2))+(σ1t2)(1+β(σ1t1)),\displaystyle\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{312}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=(\sigma_{2}\ominus t_{1})(1+\beta(\sigma_{1}\ominus t_{2}))+(\sigma_{1}\ominus t_{2})(1+\beta(\sigma_{1}\ominus t_{1})),
¯1β𝒢231(β)(σ1,σ2;t1,t2)=¯1β𝒢132(β)(σ1,σ2;t1,t2)=¯1β𝒢123(β)(σ1,σ2;t1,t2)=0.\displaystyle\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{231}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{132}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{123}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=0.

Now we would like to apply Theorem 4.6 and Lemma 4.7 to the Bethe ansatz state (3.4). First, we show the following

Theorem 4.8.

For any wSnw\in S_{n}, we have the following form of the Cauchy identity

𝒢w(β)(𝒙;𝒕)=vSn𝒢v(β)(𝒙;𝒛)¯vβ𝒢w(β)(𝒛;𝒕).\mathcal{G}^{(\beta)}_{w}(\bm{x};\ominus\bm{t})=\sum_{v\in S_{n}}\mathcal{G}^{(\beta)}_{v}(\bm{x};\ominus\bm{z})\bar{\partial}^{\beta}_{v}\mathcal{G}^{(\beta)}_{w}(\bm{z};\ominus\bm{t}).
Proof.

The generalized Cauchy identity Eq.(A.7) can be rewritten as (notice that z=z\ominus\ominus z=z)

𝒢w(β)(𝒕;𝒙)=vSn𝒢v(β)(𝒙;𝒛)𝒢v1,w(β)(𝒕;𝒛).\mathcal{G}^{(\beta)}_{w}(\ominus\bm{t};\bm{x})=\sum_{v\in S_{n}}\mathcal{G}^{(\beta)}_{v}(\bm{x};\ominus\bm{z})\mathcal{G}^{(\beta)}_{v^{-1},w}(\ominus\bm{t};\bm{z}). (4.35)

From Eq.(A.4), (A.5) and (A.6), we get

𝒢v1,w(β)(𝒕;𝒛)=¯v,𝒛β𝒢1,w(β)(𝒕;𝒛)=¯vβ𝒢w1(β)(𝒛;𝒕).\mathcal{G}^{(\beta)}_{v^{-1},w}(\ominus\bm{t};\bm{z})=\bar{\partial}^{\beta}_{v,\bm{z}}\mathcal{G}^{(\beta)}_{1,w}(\ominus\bm{t};\bm{z})=\bar{\partial}^{\beta}_{v}\mathcal{G}^{(\beta)}_{w^{-1}}(\bm{z};\ominus\bm{t}).

Plugging the identity above in (4.35), we arrive at

𝒢w1(β)(𝒙;𝒕)=vSn𝒢v(β)(𝒙;𝒛)¯vβ𝒢w1(β)(𝒛;𝒕).\mathcal{G}^{(\beta)}_{w^{-1}}(\bm{x};\ominus\bm{t})=\sum_{v\in S_{n}}\mathcal{G}^{(\beta)}_{v}(\bm{x};\ominus\bm{z})\bar{\partial}^{\beta}_{v}\mathcal{G}^{(\beta)}_{w^{-1}}(\bm{z};\ominus\bm{t}).

Then the theorem is proved by replacing w1w^{-1} with ww. ∎

Let Ln(𝒕,β)L_{n}(\bm{t},\beta) be the submodule of (𝒕,β)[x1,,xn]\mathbb{C}(\bm{t},\beta)[x_{1},\cdots,x_{n}] defined by

Ln(𝒕,β)=span(𝒕,β){x1i1x2i2xnin|0ijnj}.L_{n}(\bm{t},\beta)=\mathrm{span}_{\mathbb{C}(\bm{t},\beta)}\{x_{1}^{i_{1}}x_{2}^{i_{2}}\cdots x_{n}^{i_{n}}~{}|~{}0\leq i_{j}\leq n-j\}. (4.36)
Corollary 4.9.

For any F(𝐱;𝐭,β)Ln(𝐭,β)F(\bm{x};\bm{t},\beta)\in L_{n}(\bm{t},\beta), we have

F(𝒙;𝒕,β)=vSn𝒢v(β)(𝒙;𝒛)¯vβF(𝒛;𝒕,β).F(\bm{x};\bm{t},\beta)=\sum_{v\in S_{n}}\mathcal{G}^{(\beta)}_{v}(\bm{x};\ominus\bm{z})\bar{\partial}^{\beta}_{v}F(\bm{z};\bm{t},\beta). (4.37)
Proof.

Since {𝒢w(β)(𝒙;𝒕)|wSn}\{\mathcal{G}^{(\beta)}_{w}(\bm{x};\ominus\bm{t})~{}|~{}w\in S_{n}\} is a (𝒕,β)\mathbb{C}(\bm{t},\beta)-basis of Ln(𝒕,β)L_{n}(\bm{t},\beta) (as are the Schubert and Grothendieck polynomials [20, Proposition 2.7]), Eq.(4.37) follows from Theorem 4.8 and linearity of ¯vβ\bar{\partial}^{\beta}_{v}. ∎

In section 3, we have seen that σa(Ni)\sigma^{(N-i)}_{a} (β=0\beta=0) or 1σa(Ni)1-\sigma^{(N-i)}_{a} (β=1\beta=-1), a=1,,ia=1,\cdots,i, can be identified with the Chern roots of 𝒮i\mathcal{S}_{i} when σa(Ni)\sigma^{(N-i)}_{a}’s satisfy the Bethe ansatz equations Eq.(3.5)-(3.7). We will show that the Bethe ansatz state (3.4) generates the double β\beta-Grothendieck polynomials when this identification is implemented. For this purpose, we need the following

Lemma 4.10.

For πSn\pi\in S_{n}, if el(σ1,,σn1)e_{l}(\sigma_{1},\cdots,\sigma_{n-1}) is identified with el(x1,,xn1)e_{l}(x_{1},\cdots,x_{n-1}) for every l=1,,n1l=1,\cdots,n-1, then

wSn1𝒢w(β)(x1,,xn2;σ1,,σn2)¯wβ𝒢π(β)(σ1,,σn1;t1,,tn1)\displaystyle\sum_{w\in S_{n-1}}\mathcal{G}^{(\beta)}_{w}(x_{1},\cdots,x_{n-2};\ominus\sigma_{1},\cdots,\ominus\sigma_{n-2})\cdot\bar{\partial}^{\beta}_{w}\mathcal{G}^{(\beta)}_{\pi}(\sigma_{1},\cdots,\sigma_{n-1};\ominus t_{1},\cdots,\ominus t_{n-1})
=\displaystyle= 𝒢π(β)(x1,,xn1;t1,,tn1).\displaystyle\mathcal{G}^{(\beta)}_{\pi}(x_{1},\cdots,x_{n-1};\ominus t_{1},\cdots,\ominus t_{n-1}).
Proof.

Let Ln(𝒕,β)L_{n}(\bm{t},\beta) be the submodule defined by (4.36). Then, since {𝒢w(β)(𝒙;𝒕)|wSn}\{\mathcal{G}^{(\beta)}_{w}(\bm{x};\ominus\bm{t})~{}|~{}w\in S_{n}\} and
{ei1(x1)ei2(x1,x2)ein1(x1,,xn1)|0ijj}\{e_{i_{1}}(x_{1})e_{i_{2}}(x_{1},x_{2})\cdots e_{i_{n-1}}(x_{1},\cdots,x_{n-1})~{}|~{}0\leq i_{j}\leq j\} are both (𝒕,β)\mathbb{C}(\bm{t},\beta)-bases of Ln(𝒕,β)L_{n}(\bm{t},\beta), we have the expansion

𝒢π(β)(σ1,,σn1;t1,,tn1)\displaystyle\mathcal{G}^{(\beta)}_{\pi}(\sigma_{1},\cdots,\sigma_{n-1};\ominus t_{1},\cdots,\ominus t_{n-1})
=\displaystyle= i1,i2,,in1Ci1in1π(𝒕,β)ei1(σ1)ei2(σ1,σ2)ein1(σ1,,σn1),\displaystyle\sum_{i_{1},i_{2},\cdots,i_{n-1}}C^{\pi}_{i_{1}\cdots i_{n-1}}(\bm{t},\beta)e_{i_{1}}(\sigma_{1})e_{i_{2}}(\sigma_{1},\sigma_{2})\cdots e_{i_{n-1}}(\sigma_{1},\cdots,\sigma_{n-1}),

where Ci1in1π(𝒕,β)(𝒕,β)C^{\pi}_{i_{1}\cdots i_{n-1}}(\bm{t},\beta)\in\mathbb{C}(\bm{t},\beta).

Therefore

wSn1𝒢w(β)(x1,,xn2;σ1,,σn2)¯wβ𝒢π(β)(σ1,,σn1;t1,,tn1)\displaystyle\sum_{w\in S_{n-1}}\mathcal{G}^{(\beta)}_{w}(x_{1},\cdots,x_{n-2};\ominus\sigma_{1},\cdots,\ominus\sigma_{n-2})\cdot\bar{\partial}^{\beta}_{w}\mathcal{G}^{(\beta)}_{\pi}(\sigma_{1},\cdots,\sigma_{n-1};\ominus t_{1},\cdots,\ominus t_{n-1})
=\displaystyle= wSn1𝒢w(β)(x1,,xn2;σ1,,σn2)\displaystyle\sum_{w\in S_{n-1}}\mathcal{G}^{(\beta)}_{w}(x_{1},\cdots,x_{n-2};\ominus\sigma_{1},\cdots,\ominus\sigma_{n-2})
¯wβ[i1,i2,,in1Ci1in1π(𝒕,β)ei1(σ1)ei2(σ1,σ2)ein1(σ1,,σn1)]\displaystyle\quad\quad\quad\cdot\bar{\partial}^{\beta}_{w}\left[\sum_{i_{1},i_{2},\cdots,i_{n-1}}C^{\pi}_{i_{1}\cdots i_{n-1}}(\bm{t},\beta)e_{i_{1}}(\sigma_{1})e_{i_{2}}(\sigma_{1},\sigma_{2})\cdots e_{i_{n-1}}(\sigma_{1},\cdots,\sigma_{n-1})\right]
=\displaystyle= i1,i2,,in1Ci1in1π(𝒕,β)ein1(σ1,,σn1)\displaystyle\sum_{i_{1},i_{2},\cdots,i_{n-1}}C^{\pi}_{i_{1}\cdots i_{n-1}}(\bm{t},\beta)e_{i_{n-1}}(\sigma_{1},\cdots,\sigma_{n-1})
[wSn1𝒢w(β)(x1,,xn2;σ1,,σn2)¯wβ(ei1(σ1)ei2(σ1,σ2)ein2(σ1,,σn2))]\displaystyle\cdot\left[\sum_{w\in S_{n-1}}\mathcal{G}^{(\beta)}_{w}(x_{1},\cdots,x_{n-2};\ominus\sigma_{1},\cdots,\ominus\sigma_{n-2})\cdot\bar{\partial}^{\beta}_{w}\left(e_{i_{1}}(\sigma_{1})e_{i_{2}}(\sigma_{1},\sigma_{2})\cdots e_{i_{n-2}}(\sigma_{1},\cdots,\sigma_{n-2})\right)\right]
=\displaystyle= i1,i2,,in1Ci1in1π(𝒕,β)ei1(x1)ei2(x1,x2)ein2(x1,,xn2)ein1(σ1,,σn1)\displaystyle\sum_{i_{1},i_{2},\cdots,i_{n-1}}C^{\pi}_{i_{1}\cdots i_{n-1}}(\bm{t},\beta)e_{i_{1}}(x_{1})e_{i_{2}}(x_{1},x_{2})\cdots e_{i_{n-2}}(x_{1},\cdots,x_{n-2})e_{i_{n-1}}(\sigma_{1},\cdots,\sigma_{n-1})
=\displaystyle= i1,i2,,in1Ci1in1π(𝒕,β)ei1(x1)ei2(x1,x2)ein2(x1,,xn2)ein1(x1,,xn1)\displaystyle\sum_{i_{1},i_{2},\cdots,i_{n-1}}C^{\pi}_{i_{1}\cdots i_{n-1}}(\bm{t},\beta)e_{i_{1}}(x_{1})e_{i_{2}}(x_{1},x_{2})\cdots e_{i_{n-2}}(x_{1},\cdots,x_{n-2})e_{i_{n-1}}(x_{1},\cdots,x_{n-1})
=\displaystyle= 𝒢π(β)(x1,,xn1;t1,,tn1),\displaystyle\mathcal{G}^{(\beta)}_{\pi}(x_{1},\cdots,x_{n-1};\ominus t_{1},\cdots,\ominus t_{n-1}),

where the second equality is due to ¯iβ(fg)=f(¯iβg)\bar{\partial}^{\beta}_{i}(f\cdot g)=f(\bar{\partial}^{\beta}_{i}g) if ff is symmetric in (σ1,,σn1)(\sigma_{1},\cdots,\sigma_{n-1}), the third equality is due to Corollary 4.9, and ein1(σ1,,σn1)e_{i_{n-1}}(\sigma_{1},\cdots,\sigma_{n-1}) is identified with ein1(x1,,xn1)e_{i_{n-1}}(x_{1},\cdots,x_{n-1}) in the fourth equality. ∎

Example 2.

Notice that 𝒢12(β)(x1;σ1)=1,𝒢21(β)(x1;σ1)=x1σ1\mathcal{G}^{(\beta)}_{12}(x_{1};\ominus\sigma_{1})=1,\mathcal{G}^{(\beta)}_{21}(x_{1};\ominus\sigma_{1})=x_{1}\ominus\sigma_{1}. In the case n=3n=3, following Example 1, one can check

𝒢12(β)(x1;σ1)𝒢321(β)(σ1,σ2;t1,t2)+𝒢21(β)(x1;σ1)¯1β𝒢321(β)(σ1,σ2;t1,t2)\displaystyle\mathcal{G}^{(\beta)}_{12}(x_{1};\ominus\sigma_{1})\mathcal{G}^{(\beta)}_{321}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})+\mathcal{G}^{(\beta)}_{21}(x_{1};\ominus\sigma_{1})\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{321}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})
=\displaystyle= (σ1t1)(σ2t1)(x1t2)=(x1t1)(x2t1)(x1t2)\displaystyle(\sigma_{1}\ominus t_{1})(\sigma_{2}\ominus t_{1})(x_{1}\ominus t_{2})=(x_{1}\ominus t_{1})(x_{2}\ominus t_{1})(x_{1}\ominus t_{2})
=\displaystyle= 𝒢321(β)(x1,x2;t1,t2),\displaystyle\mathcal{G}^{(\beta)}_{321}(x_{1},x_{2};\ominus t_{1},\ominus t_{2}),
𝒢12(β)(x1;σ1)𝒢231(β)(σ1,σ2;t1,t2)=(σ1t1)(σ2t1)=(x1t1)(x2t1)\displaystyle\mathcal{G}^{(\beta)}_{12}(x_{1};\ominus\sigma_{1})\mathcal{G}^{(\beta)}_{231}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=(\sigma_{1}\ominus t_{1})(\sigma_{2}\ominus t_{1})=(x_{1}\ominus t_{1})(x_{2}\ominus t_{1})
=\displaystyle= 𝒢231(β)(x1,x2;t1,t2),\displaystyle\mathcal{G}^{(\beta)}_{231}(x_{1},x_{2};\ominus t_{1},\ominus t_{2}),
𝒢12(β)(x1;σ1)𝒢312(β)(σ1,σ2;t1,t2)+𝒢21(β)(x1;σ1)¯1β𝒢312(β)(σ1,σ2;t1,t2)\displaystyle\mathcal{G}^{(\beta)}_{12}(x_{1};\ominus\sigma_{1})\mathcal{G}^{(\beta)}_{312}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})+\mathcal{G}^{(\beta)}_{21}(x_{1};\ominus\sigma_{1})\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{312}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})
=\displaystyle= x1(σ1+σ2t1t2)+(t1t2σ1σ2)(1+βt1)(1+βt2)=(x1t1)(x2t2)\displaystyle\frac{x_{1}(\sigma_{1}+\sigma_{2}-t_{1}-t_{2})+(t_{1}t_{2}-\sigma_{1}\sigma_{2})}{(1+\beta t_{1})(1+\beta t_{2})}=(x_{1}\ominus t_{1})(x_{2}\ominus t_{2})
=\displaystyle= 𝒢312(β)(x1,x2;t1,t2),\displaystyle\mathcal{G}^{(\beta)}_{312}(x_{1},x_{2};\ominus t_{1},\ominus t_{2}),
𝒢12(β)(x1;σ1)𝒢132(β)(σ1,σ2;t1,t2)=x1+x2t1t2+β(x1x2t1t2)(1+βt1)(1+βt2)\displaystyle\mathcal{G}^{(\beta)}_{12}(x_{1};\ominus\sigma_{1})\mathcal{G}^{(\beta)}_{132}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})=\frac{x_{1}+x_{2}-t_{1}-t_{2}+\beta(x_{1}x_{2}-t_{1}t_{2})}{(1+\beta t_{1})(1+\beta t_{2})}
=\displaystyle= 𝒢132(β)(x1,x2;t1,t2),\displaystyle\mathcal{G}^{(\beta)}_{132}(x_{1},x_{2};\ominus t_{1},\ominus t_{2}),
𝒢12(β)(x1;σ1)𝒢213(β)(σ1,σ2;t1,t2)+𝒢21(β)(x1;σ1)¯1β𝒢213(β)(σ1,σ2;t1,t2)\displaystyle\mathcal{G}^{(\beta)}_{12}(x_{1};\ominus\sigma_{1})\mathcal{G}^{(\beta)}_{213}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})+\mathcal{G}^{(\beta)}_{21}(x_{1};\ominus\sigma_{1})\bar{\partial}^{\beta}_{1}\mathcal{G}^{(\beta)}_{213}(\sigma_{1},\sigma_{2};\ominus t_{1},\ominus t_{2})
=\displaystyle= (σ1t1)+(x1σ1)(1+β(σ1t1))=x1t1\displaystyle(\sigma_{1}\ominus t_{1})+(x_{1}\ominus\sigma_{1})(1+\beta(\sigma_{1}\ominus t_{1}))=x_{1}\ominus t_{1}
=\displaystyle= 𝒢213(β)(x1,x2;t1,t2),\displaystyle\mathcal{G}^{(\beta)}_{213}(x_{1},x_{2};\ominus t_{1},\ominus t_{2}),

where the following identifications have been made

e1(σ1,σ2)=e1(x1,x2),e2(σ1,σ2)=e2(x1,x2).e_{1}(\sigma_{1},\sigma_{2})=e_{1}(x_{1},x_{2}),\quad e_{2}(\sigma_{1},\sigma_{2})=e_{2}(x_{1},x_{2}).

Finally, we show that the Bethe state satisfies the following

Theorem 4.11.

When el(σ1(i),,σNi(i))e_{l}(\sigma^{(i)}_{1},\cdots,\sigma^{(i)}_{N-i}) is identified with el(x1,,xNi)e_{l}(x_{1},\cdots,x_{N-i}) for all i=1,,N1i=1,\cdots,N-1 and l=1,,Nil=1,\cdots,N-i, the Bethe ansatz state (3.4) (in the case ki=Nik_{i}=N-i) has the expansion

|ψ(0)=wSN𝒢w(β)(x1,,xN1;t1,,tN1)|ωNw1.\ket{\psi^{(0)}}=\sum_{w\in S_{N}}\mathcal{G}^{(\beta)}_{w}(x_{1},\cdots,x_{N-1};\ominus t_{1},\cdots,\ominus t_{N-1})\ket{\omega_{N}w^{-1}}. (4.38)
Proof.

We prove by induction on the number of sites. For a two-site system, it is easy to check that, when σ1(1)=x1\sigma^{(1)}_{1}=x_{1},

01|B1(σ1(1))|00=σ1(1)t1=x1t1=𝒢21(β)(x1;t1),\displaystyle\bra{01}B_{1}(\sigma^{(1)}_{1})\ket{00}=\sigma^{(1)}_{1}\ominus t_{1}=x_{1}\ominus t_{1}=\mathcal{G}^{(\beta)}_{21}(x_{1};\ominus t_{1}),
10|B1(σ1(1))|00=1=𝒢12(β)(x1;t1),\displaystyle\bra{10}B_{1}(\sigma^{(1)}_{1})\ket{00}=1=\mathcal{G}^{(\beta)}_{12}(x_{1};\ominus t_{1}),

so Eq.(4.38) holds for N=2N=2. Now assume Eq.(4.38) holds for N=2,3,,mN=2,3,\cdots,m with m2m\geq 2, then we have, from Eq.(3.14) and (3.15), for wlSlw_{l}\in S_{l} (notice that |ψ(Ns)\ket{\psi^{(N-s)}} is the Bethe state of a GL(s)(s) system with ss sites),

ω2w21|ψ(N2)\displaystyle\braket{\omega_{2}w_{2}^{-1}}{\psi^{(N-2)}} =𝒢w2(β)(x1;σ1(N2)),\displaystyle=\mathcal{G}^{(\beta)}_{w_{2}}(x_{1};\ominus\sigma_{1}^{(N-2)}),
ω3w31|ψ(N3)\displaystyle\braket{\omega_{3}w_{3}^{-1}}{\psi^{(N-3)}} =𝒢w3(β)(x1,x2;σ1(N3),σ2(N3)),\displaystyle=\mathcal{G}^{(\beta)}_{w_{3}}(x_{1},x_{2};\ominus\sigma_{1}^{(N-3)},\ominus\sigma_{2}^{(N-3)}),
\displaystyle~{}~{}\vdots
ωmwm1|ψ(Nm)\displaystyle\braket{\omega_{m}w_{m}^{-1}}{\psi^{(N-m)}} =𝒢wm(β)(x1,,xm1;σ1(Nm),,σm1(Nm)).\displaystyle=\mathcal{G}^{(\beta)}_{w_{m}}(x_{1},\cdots,x_{m-1};\ominus\sigma_{1}^{(N-m)},\cdots,\ominus\sigma_{m-1}^{(N-m)}).

Eq.(3.14) and (3.15) yield

|ψ(Nm1)=vmSmvm|ψ(Nm)Bvm(1)(Nm1)(σ1(Nm))Bvm(m)(Nm1)(σm(Nm))|Ω(Nm1).\ket{\psi^{(N-m-1)}}=\sum_{v_{m}\in S_{m}}\braket{v_{m}}{\psi^{(N-m)}}B^{(N-m-1)}_{v_{m}(1)}(\sigma^{(N-m)}_{1})\cdots B^{(N-m-1)}_{v_{m}(m)}(\sigma^{(N-m)}_{m})\ket{\Omega^{(N-m-1)}}. (4.39)

It is easy to verify that

ωm+1wm+11|Bvm(1)(Nm1)(σ1(Nm))Bvm(m)(Nm1)(σm(Nm))|Ω(Nm1)\displaystyle\bra{\omega_{m+1}w_{m+1}^{-1}}B^{(N-m-1)}_{v_{m}(1)}(\sigma^{(N-m)}_{1})\cdots B^{(N-m-1)}_{v_{m}(m)}(\sigma^{(N-m)}_{m})\ket{\Omega^{(N-m-1)}}
=\displaystyle= ¯vm1ωmβωm+1wm+11|Bm(Nm1)(σ1(Nm))B1(Nm1)(σm(Nm))|Ω(Nm1)\displaystyle\bar{\partial}^{\beta}_{v^{-1}_{m}\omega_{m}}\bra{\omega_{m+1}w_{m+1}^{-1}}B^{(N-m-1)}_{m}(\sigma^{(N-m)}_{1})\cdots B^{(N-m-1)}_{1}(\sigma^{(N-m)}_{m})\ket{\Omega^{(N-m-1)}}
=\displaystyle= ¯vm1ωmβ𝒢wm+1(β)(σ1(Nm),,σm(Nm);σ1(Nm1),,σm(Nm1)),\displaystyle\bar{\partial}^{\beta}_{v^{-1}_{m}\omega_{m}}\mathcal{G}^{(\beta)}_{w_{m+1}}(\sigma^{(N-m)}_{1},\cdots,\sigma^{(N-m)}_{m};\ominus\sigma_{1}^{(N-m-1)},\cdots,\ominus\sigma_{m}^{(N-m-1)}),

where the first equality is due to Lemma 4.7 and the second equality is due to Theorem 4.6. As vm|ψ(Nm)=𝒢vm1ωm(β)(x1,,xm1;σ1(Nm),,σm1(Nm))\braket{v_{m}}{\psi^{(N-m)}}=\mathcal{G}^{(\beta)}_{v_{m}^{-1}\omega_{m}}(x_{1},\cdots,x_{m-1};\ominus\sigma^{(N-m)}_{1},\cdots,\ominus\sigma^{(N-m)}_{m-1}) from the induction hypothesis, Lemma 4.10 and Eq.(4.39) then yield

ωm+1wm+11|ψ(Nm1)\displaystyle\braket{\omega_{m+1}w_{m+1}^{-1}}{\psi^{(N-m-1)}}
=\displaystyle= vmSmvm|ψ(Nm)ωm+1wm+11|Bvm(1)(Nm1)(σ1(Nm))Bvm(m)(Nm1)(σm(Nm))|Ω(Nm1)\displaystyle\sum_{v_{m}\in S_{m}}\braket{v_{m}}{\psi^{(N-m)}}\bra{\omega_{m+1}w_{m+1}^{-1}}B^{(N-m-1)}_{v_{m}(1)}(\sigma^{(N-m)}_{1})\cdots B^{(N-m-1)}_{v_{m}(m)}(\sigma^{(N-m)}_{m})\ket{\Omega^{(N-m-1)}}
=\displaystyle= vmSm𝒢vm1ωm(β)(x1,,xm1;σ1(Nm),,σm1(Nm))\displaystyle\sum_{v_{m}\in S_{m}}\mathcal{G}^{(\beta)}_{v^{-1}_{m}\omega_{m}}(x_{1},\cdots,x_{m-1};\ominus\sigma^{(N-m)}_{1},\cdots,\ominus\sigma^{(N-m)}_{m-1})
¯vm1ωmβ𝒢wm+1(β)(σ1(Nm),,σm(Nm);σ1(Nm1),,σm(Nm1))\displaystyle\quad\quad\quad\cdot\bar{\partial}^{\beta}_{v_{m}^{-1}\omega_{m}}\mathcal{G}^{(\beta)}_{w_{m+1}}(\sigma^{(N-m)}_{1},\cdots,\sigma^{(N-m)}_{m};\ominus\sigma^{(N-m-1)}_{1},\cdots,\ominus\sigma^{(N-m-1)}_{m})
=\displaystyle= 𝒢wm+1(β)(x1,,xm;σ1(Nm1),,σm(Nm1)).\displaystyle\mathcal{G}^{(\beta)}_{w_{m+1}}(x_{1},\cdots,x_{m};\ominus\sigma^{(N-m-1)}_{1},\cdots,\ominus\sigma^{(N-m-1)}_{m}).

Notice that σi(0)=ti\sigma^{(0)}_{i}=t_{i}, the proof is thus complete by induction. ∎

Remark 4.

As discussed in Section 3, when the Bethe ansatz equations Eq.(3.5)-(3.7) are taken into account, we should interpret σa(Ni)\sigma^{(N-i)}_{a} (β=0\beta=0) or 1σa(Ni)1-\sigma^{(N-i)}_{a} (β=1\beta=-1), a=1,,ia=1,\cdots,i, as the Chern roots of 𝒮i\mathcal{S}_{i}. Therefore, xix_{i} should be interpreted as the first Chern class of 𝒮i/𝒮i1\mathcal{S}_{i}/\mathcal{S}_{i-1} when β=0\beta=0 in the cohomology ring, and 1xi1-x_{i} should be interpreted as the K-theory class of 𝒮i/𝒮i1\mathcal{S}_{i}/\mathcal{S}_{i-1} when β=1\beta=-1 in the K-theory ring.

Remark 5.

More generally, for β0\beta\neq 0, we expect 1+βxi1+\beta x_{i} to represent 𝒮i/𝒮i1\mathcal{S}_{i}/\mathcal{S}_{i-1} in the connective K-theory ring of the flag variety [21].

Remark 6.

In the case of full flag variety, iaibi_{a}\neq i_{b} for aba\neq b in Eq.(3.14). In the case of general partial flag variety Fl(kn1,kn2,,k1;N)\mathrm{Fl}(k_{n-1},k_{n-2},\cdots,k_{1};N), it is possible that ia=ibi_{a}=i_{b} for some aba\neq b. Because of [Bi(x),Bi(y)]=0[B_{i}(x),B_{i}(y)]=0 (Eq.(4.27)), we expect the expansion coefficients of the Bethe ansatz state in the natural basis to be given by the β\beta-Grothendieck polynomials indexed by SN/WS_{N}/W, where WW is generated by the simple transpositions sis_{i} for i{kn1,kn2,,k1}i\not\in\{k_{n-1},k_{n-2},\cdots,k_{1}\}, i.e. β\beta-Grothendieck polynomials symmetric in (xks+1,,xks1)(x_{k_{s}+1},\cdots,x_{k_{s-1}}), which can be identified with the Chern roots of 𝒮ns+1/𝒮ns\mathcal{S}_{n-s+1}/\mathcal{S}_{n-s}, for all s=1,,ns=1,\cdots,n.

Acknowledgement

The author would like to thank Xiang-Mao Ding, Leonardo Mihalcea and Eric Sharpe for useful comments. This work is partially supported by the National Natural Science Foundation of China (Grant No. 12475005), the Natural Science Foundation of Shanghai (Grant No. 24ZR1468600), and the Fundamental Research Funds for the Central Universities.

Appendix A Quantum cohomology/K-theory of flag varieties and the double β\beta-Grothendieck polynomials

In this appendix, we give a brief review of the quantum cohomology/K-theory ring of flag varieties, and the double β\beta-Grothendieck polynomials. The reader may refer to e.g. [13, 17, 18, 20, 22, 23, 24] and references therein for more information.

Given E=NE=\mathbb{C}^{N} and 𝒂=(a1,a2,,am)\bm{a}=(a_{1},a_{2},\cdots,a_{m}) such that 0<a1<a2<<am<N0<a_{1}<a_{2}<\cdots<a_{m}<N, the flag variety Fl(𝒂;N)=Fl(a1,a2,,am;N)\mathrm{Fl}(\bm{a};N)=\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N) is defined to be the set of the flags

V1V2VmEV_{1}\subset V_{2}\subset\cdots\subset V_{m}\subset E

such that dim(Vi)=ai\dim(V_{i})=a_{i} for all ii.

There is a sequence of vector bundles on Fl(a1,a2,,am;N)\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N) (the tautological sequence):

𝒮1𝒮2𝒮mE\mathcal{S}_{1}\subset\mathcal{S}_{2}\subset\cdots\subset\mathcal{S}_{m}\subset E

with rank(𝒮i)=ai\mathrm{rank}(\mathcal{S}_{i})=a_{i}. Let SNS_{N} be the group of permutations of NN elements, and let l(w)l(w) be the length777The length of a permutation wSNw\in S_{N} is the number of pairs (i,j)(i,j) such that i<ji<j and w(i)>w(j)w(i)>w(j). of a permutation wSNw\in\mathrm{S}_{N}. The Schubert varieties are subvarieties of Fl(a1,a2,,am;N)\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N) indexed by the set SN(𝒂)=SN/W𝒂S_{N}(\bm{a})=S_{N}/W_{\bm{a}}, where W𝒂SNW_{\bm{a}}\subset S_{N} is the subgroup generated by the simple transpositions si=(i,i+1)s_{i}=(i,i+1) for i{a1,,am}i\not\in\{a_{1},\cdots,a_{m}\}. The codimension of the Schubert variety indexed by the representative wSNw\in S_{N} with shortest length is equal to l(w)l(w).

Let Ωw(𝒂)\Omega^{(\bm{a})}_{w} denote the cohomology class Poincaré dual to the Schubert variety indexed by ww, then the Schubert classes Ωw(𝒂)\Omega^{(\bm{a})}_{w} form a basis for the cohomology ring H(Fl(a1,a2,,al;N),)H^{*}(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{l};N),\mathbb{Z}).

Let Fl(N):=Fl(1,2,,N1;N)\mathrm{Fl}(N):=\mathrm{Fl}(1,2,\cdots,N-1;N) denote the full flag variety. The cohomology of Fl(N)\mathrm{Fl}(N) has the following presentation

H(Fl(N))=[x1,,xN]/(e1N,,eNN),H^{*}(\mathrm{Fl}(N))=\mathbb{Z}[x_{1},\cdots,x_{N}]/(e^{N}_{1},\cdots,e^{N}_{N}),

where eiN=ei(x1,,xN)e^{N}_{i}=e_{i}(x_{1},\cdots,x_{N}) is the iith elementary symmetric polynomial in NN variables. In this presentation, xix_{i} is identified with the Chern class c1(𝒮i/𝒮i1)c_{1}(\mathcal{S}_{i}/\mathcal{S}_{i-1}) and the Schubert class Ωw\Omega_{w} is represented by the Schubert polynomial Xw(x1,,xN)X_{w}(x_{1},\cdots,x_{N}), which means that the multiplication rules of the Schubert classes are the same as those of the Schubert polynomials modulo eiN,i=1,,Ne^{N}_{i},i=1,\cdots,N.

The quantum elementary polynomials EiNE^{N}_{i} are defined via the Givental-Kim determinant. Let

ΓN=(x1q1001x2q2001x30000xN),\Gamma_{N}=\left(\begin{array}[]{ccccc}x_{1}&q_{1}&0&\cdots&0\\ -1&x_{2}&q_{2}&\cdots&0\\ 0&-1&x_{3}&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&x_{N}\end{array}\right),

then EiNE^{N}_{i} is defined as the coefficient of λi\lambda^{i} in the characteristic polynomial det(1+λΓN)\det(1+\lambda\Gamma_{N}). The quantum cohomology ring of the full flag variety can be presented as

QH(Fl(N))=[x1,,xN]/(E1N,,ENN).QH^{*}(\mathrm{Fl}(N))=\mathbb{Z}[x_{1},\cdots,x_{N}]/(E^{N}_{1},\cdots,E^{N}_{N}).

Ring relations of the quantum cohomology of partial flag varieties QH(Fl(a1,a2,,am;N))QH^{*}(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N)) can be presented similarly, see [13, 22]. The structure constants of QH(Fl(a1,a2,,am;N))QH^{*}(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N)) encode the Gromov-Witten invariants of the flag variety.

The K-theory ring of the flag variety K(Fl(a1,a2,,am;N))K(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N)) is the Grothendieck group of the coherent sheaves of Fl(a1,a2,,am;N)\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N). A basis of K(Fl(a1,a2,,am;N))K(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N)) can be chosen as the structure sheaves of the Schubert varieties {𝒪w|wSn(𝒂)}\{\mathcal{O}_{w}|w\in S_{n}(\bm{a})\}, which can be represented by the Grothendieck polynomials Gw(x1,,xN)G_{w}(x_{1},\cdots,x_{N}). Ring relations can be found in e.g. [23]. One way to present the ring relations of the (quantum) K-theory ring of Fl(a1,a2,,am;N)\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N) is to encapsulate them in the (quantum) Whitney relations [12]. The structure constants of the quantum K-theory ring QK(Fl(a1,a2,,am;N))QK(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N)) encode the K-theoretic Gromov-Witten invariants of the flag variety.

For the T=()NT=(\mathbb{C}^{*})^{N} action on E=NE=\mathbb{C}^{N} with characters (equivariant parameters) (t1,,tN)(t_{1},\cdots,t_{N}), we can define the equivariant quantum cohomology ring QHT(Fl(a1,a2,,am;N))QH_{T}^{*}(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N)) and equivariant quantum K-theory ring QKT(Fl(a1,a2,,am;N))QK_{T}(\mathrm{Fl}(a_{1},a_{2},\cdots,a_{m};N)). The representatives of the Schubert classes in the equivariant cohomology and K-theory rings are given by the double Schubert polynomials Xw(x1,,xN;t1,,tN)X_{w}(x_{1},\cdots,x_{N};t_{1},\cdots,t_{N}) and the double Grothendieck polynomials Gw(x1,,xN;t1,,tN)G_{w}(x_{1},\cdots,x_{N};t_{1},\cdots,t_{N}) respectively.

The double Schubert and double Grothendieck polynomials are special cases of a one-parameter family of polynomials called the double β\beta-Grothedieck polynomials 𝒢w(β),wSN\mathcal{G}^{(\beta)}_{w},w\in S_{N} [18]. Let 𝕂\mathbb{K} be a field of zero characteristic and β\beta be a formal variable, 𝒙=(x1,,xN),𝒚=(y1,,yN)\bm{x}=(x_{1},\cdots,x_{N}),\bm{y}=(y_{1},\cdots,y_{N}), define the β\beta-divided-difference operator iβ\partial^{\beta}_{i} acting on 𝕂(β)[x1,,xN,y1,,yN]\mathbb{K}(\beta)[x_{1},\cdots,x_{N},y_{1},\cdots,y_{N}] by

iβf(x1,,xN;𝒚):=(1+βxi+1)f(x1,,xN;𝒚)(1+βxi)f(x1,,xi+1,xi,,xN;𝒚)xixi+1.\partial^{\beta}_{i}f(x_{1},\cdots,x_{N};\bm{y}):=\frac{(1+\beta x_{i+1})f(x_{1},\cdots,x_{N};\bm{y})-(1+\beta x_{i})f(x_{1},\cdots,x_{i+1},x_{i},\cdots,x_{N};\bm{y})}{x_{i}-x_{i+1}}. (A.1)

If w=ωNw=\omega_{N} is the permutation in SNS_{N} with maximal length, then we define

𝒢ωN(β)(𝒙;𝒚)=i+jN(xi+yj+βxiyj),\mathcal{G}^{(\beta)}_{\omega_{N}}(\bm{x};\bm{y})=\prod_{i+j\leq N}(x_{i}+y_{j}+\beta x_{i}y_{j}), (A.2)

and for other wSNw\in\mathrm{S}_{N}, 𝒢w(β)\mathcal{G}^{(\beta)}_{w} can be defined recursively by

𝒢wsi(β)(𝒙;𝒚)=iβ𝒢w(β)(𝒙;𝒚)\mathcal{G}^{(\beta)}_{ws_{i}}(\bm{x};\bm{y})=\partial^{\beta}_{i}\mathcal{G}^{(\beta)}_{w}(\bm{x};\bm{y}) (A.3)

if l(wsi)=l(w)1l(ws_{i})=l(w)-1.

The double β\beta-Grothendieck polynomial 𝒢w(β)(𝒙;𝒚)\mathcal{G}^{(\beta)}_{w}(\bm{x};\bm{y}) reduces to the double Schubert polynomial Xw(𝒙;𝒚)X_{w}(\bm{x};\bm{y}) when β=0\beta=0, and reduces to the double Grothendieck polynomial Gw(𝒙;𝒚)G_{w}(\bm{x};\bm{y}) when β=1\beta=-1.

The double β\beta-Grothendieck polynomials satisfy (see Lemma 5.8 of [21])

𝒢w(β)(𝒙;𝒚)=𝒢w1(β)(𝒚;𝒙).\mathcal{G}^{(\beta)}_{w}(\bm{x};\bm{y})=\mathcal{G}^{(\beta)}_{w^{-1}}(\bm{y};\bm{x}). (A.4)

The dual double β\beta-Grothendieck polynomials w(β)(𝒙;𝒚)\mathcal{H}^{(\beta)}_{w}(\bm{x};\bm{y}) can be defined recursively by setting wN(β)(𝒙;𝒚)=𝒢wN(β)(𝒙;𝒚)\mathcal{H}^{(\beta)}_{w_{N}}(\bm{x};\bm{y})=\mathcal{G}^{(\beta)}_{w_{N}}(\bm{x};\bm{y}) and

wsi(β)(𝒙;𝒚)=¯iβw(β)(𝒙;𝒚)\mathcal{H}^{(\beta)}_{ws_{i}}(\bm{x};\bm{y})=\bar{\partial}^{\beta}_{i}\mathcal{H}^{(\beta)}_{w}(\bm{x};\bm{y})

if l(wsi)=l(w)1l(ws_{i})=l(w)-1 [19], where ¯iβ=iβ+β\bar{\partial}^{\beta}_{i}=\partial^{\beta}_{i}+\beta.

More generally, for v,wSNv,w\in S_{N}, one can define the biaxial double β\beta-Grothendieck polynomials 𝒢v,w(β)(𝒙;𝒚)\mathcal{G}^{(\beta)}_{v,w}(\bm{x};\bm{y}) as follows (see Definition 2.6 of [9]):

𝒢1,w(β)(𝒙;𝒚)=𝒢w(β)(𝒙;𝒚).\mathcal{G}^{(\beta)}_{1,w}(\bm{x};\bm{y})=\mathcal{G}^{(\beta)}_{w}(\bm{x};\bm{y}). (A.5)

If l(vsi)=l(v)+1l(vs_{i})=l(v)+1, set

𝒢vsi,w(β)(𝒙;𝒚)=¯i,𝒚β𝒢v,w(β)(𝒙;𝒚),\mathcal{G}^{(\beta)}_{vs_{i},w}(\bm{x};\bm{y})=\bar{\partial}^{\beta}_{i,\bm{y}}\mathcal{G}^{(\beta)}_{v,w}(\bm{x};\bm{y}), (A.6)

where ¯i,𝒚β\bar{\partial}^{\beta}_{i,\bm{y}} is the ¯iβ\bar{\partial}^{\beta}_{i} operator acting on the set of variables 𝒚=(y1,y2,,yN)\bm{y}=(y_{1},y_{2},\cdots,y_{N}). We have the following generalized Cauchy identity:

Theorem A.1 (Theorem 7.1 of [9]).

For any wSNw\in S_{N},

𝒢w(β)(𝒙;𝒚)=vSN𝒢v(β)(𝒚;𝒛)𝒢v1,w(β)(𝒙;𝒛),\mathcal{G}^{(\beta)}_{w}(\bm{x};\bm{y})=\sum_{v\in S_{N}}\mathcal{G}^{(\beta)}_{v}(\bm{y};\bm{z})\mathcal{G}^{(\beta)}_{v^{-1},w}(\bm{x};\ominus\bm{z}), (A.7)

where zi=zi/(1+βzi)\ominus z_{i}=-z_{i}/(1+\beta z_{i}).

References

  • [1]
  • [2] V. Gorbounov and C. Korff, “Quantum integrability and generalised quantum Schubert calculus,” Advances in Mathematics 313, 282-356 (2014), [arXiv:1408.4718 [math.RT]].
  • [3] K. Ueda and Y. Yoshida, “3d N=2N=2 Chern-Simons-matter theory, Bethe ansatz, and quantum K-theory of Grassmannians,” JHEP 08 (2020) 157, [arXiv:1912.03792 [hep-th]].
  • [4] A. Givental and B. Kim, “Quantum cohomology of flag manifolds and Toda lattices,” Commun. Math. Phys. 168, 609-642 (1995), [arXiv:hep-th/9312096].
  • [5] A. Givental and Y. P. Lee, “Quantum K-theory on flag manifolds, finite difference Toda lattices and quantum groups,” Invent. Math. 151, 193-219 (2003), [arXiv:math/0108105].
  • [6] P. Koroteev, P. P. Pushkar, A. V. Smirnov and A. M. Zeitlin, “Quantum K-theory of quiver varieties and many-body systems,” Selecta Math. New Ser. 27 (2021) 87, [arXiv:1705.10419 [math.AG]].
  • [7] V. Buciumas, T. Scrimshaw and K. Weber, “Colored five-vertex models and Lascoux polynomials and atoms,” Journal of the London Mathematical Society, 102 (2019), [arXiv:1908.07364 [math.CO]].
  • [8] V. Buciumas and T. Scrimshaw, “Double Grothendieck Polynomials and Colored Lattice Models,” International Mathematics Research Notices (2020), [arXiv:2007.04533 [math.CO]].
  • [9] B. Brubaker, C. Frechette, A. Hardt, E. Tibor, K. Weber, “Frozen pipes: lattice models for Grothendieck polynomials,” Algebraic Combinatorics, 6 (2023) no. 3, pp. 789-833, [arXiv:2007.04310 [math.CO]].
  • [10] P. P. Kulish and N. Y. Reshetikhin, “Diagonalisation of GL(N)(N) invariant transfer matrices and quantum NN-wave system (Lee model),” J. Phys. A 16 (1983) L591-L596.
  • [11] J. Guo, “Quantum Sheaf Cohomology and Duality of Flag Manifolds,” Commun. Math. Phys. 374, no.2, 661-688 (2019), [arXiv:1808.00716 [hep-th]].
  • [12] W. Gu, L. Mihalcea, E. Sharpe, W. Xu, H. Zhang and H. Zou, “Quantum K theory rings of partial flag manifolds,” J. Geom. Phys. 198, 105127 (2024), [arXiv:2306.11094 [hep-th]].
  • [13] A. Astashkevich and V. Sadov, “Quantum cohomology of partial flag manifolds Fn1nkF_{n_{1}\cdots n_{k}},” Commun. Math. Phys. 170, 503–528 (1995), [arXiv:hep-th/9401103].
  • [14] R. Donagi and E. Sharpe, “GLSM’s for partial flag manifolds,” J. Geom. Phys. 58 (2008) 1662-1692, [arXiv:0704.1761 [hep-th]].
  • [15] N. Nekrasov and S. Shatashvili, “Supersymmetric Vacua and Bethe Ansatz,” Nuclear Physics B Proceedings Supplements 192 (2009), 91-112, [arXiv:0901.4744 [hep-th]].
  • [16] N. Nekrasov and S. Shatashvili, “Quantum Integrability and Supersymmetric Vacua,” Progress of Theoretical Physics Supplement 177 (2009), 105-119, [arXiv:0901.4748 [hep-th]].
  • [17] A. N. Kirillov and T. Maeno, “Quantum double Schubert polynomials, quantum Schubert polynomials and Vafa-Intriligator formula,” Discret. Math. 217 (1996), 191-223, [arXiv:q-alg/9610022].
  • [18] S. Fomin and A. N. Kirillov, “Yang-Baxter equation, symmetric functions and Grothendieck polynomials,” [arXiv:hep-th/9306005].
  • [19] A. Lascoux and M. P. Schützenberger, “Symmetry and flag manifolds,” in Invariant Theory (Montecatini, 1982), vol. 996 of Lecture Notes in Math., Springer, Berlin, 1983, pp. 118-144.
  • [20] C. Lenart and M. Toshiaki, “Quantum Grothendieck Polynomials,” [arXiv:math/0608232].
  • [21] T. Hudson, “A Thom-Porteous formula for connective K -theory using algebraic cobordism,” Journal of K-theory 14 (2013), 343-369, [arXiv:1310.0895 [math.AG]].
  • [22] A. S. Buch, “Quantum Cohomology of Partial Flag Manifolds,” [arXiv:math/0303245 [math.AG]].
  • [23] T. Maeno, S. Naito and D. Sagaki, “A presentation of the torus-equivariant quantum K-theory ring of flag manifolds of type A, Part I: the defining ideal,” [arXiv:2302.09485 [math.QA]].
  • [24] T. Maeno, S. Naito and D. Sagaki, “A presentation of the torus-equivariant quantum K-theory ring of flag manifolds of type A, Part II: quantum double Grothendieck polynomials,” [arXiv:2305.17685 [math.QA]].