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

qq-Kreweras numbers for coincidental Coxeter groups attached to limit symbols

Dongkwan Kim School of Mathematics
University of Minnesota Twin Cities
Minneapolis, MN 55455
U.S.A.
[email protected]
Abstract.

For a coincidental Coxeter group, i.e. of type An1,BCn,H3,A_{n-1},BC_{n},H_{3}, or I2(m)I_{2}(m), we define the corresponding qq-Kreweras numbers attached to limit symbols in the sense of Shoji. The construction of these numbers resembles the argument of Reiner and Sommers and these two share similar properties, but our version is slightly different from and combinatorially simpler than theirs. We also study the combinatorial properties of our qq-Kreweras numbers, i.e. positivity, relation with qq-Narayana numbers, and cyclic sieving phenomenon.

0. Introduction

Catalan numbers are one of the most prominent sequences in algebraic combinatorics and have a lot of applications. For example, this sequence counts binary trees with fixed vertices, Dyck paths of fixed length, non-crossing partitions of points lying on the circumference of a circle, triangulation of a regular polygon, to name a few. We refer readers to Stanley’s expository monograph [Sta15] for more details and its history.

This series has a generalization to any complex reflection group WW so that the original one corresponds to the symmetric group. Moreover, similar to many other sequences in algebraic combinatorics, these Catalan numbers admit a qq-deformation. (cf. [GG12]) Namely, suppose that d1,,dnd_{1},\ldots,d_{n} are the degrees of fundamental invariants of WW. Then the corresponding qq-Catalan number with parameter tt is given by

Cat(W,t;q)=i=1r[t1+di][di]\textnormal{Cat}(W,t;q)=\prod_{i=1}^{r}\frac{[t-1+d_{i}]}{[d_{i}]}

where [k]y=i=0k1yi[k]_{y}=\sum_{i=0}^{k-1}y^{i}. If WW is the symmetric group permuting nn elements, t=n+1t=n+1, and q=1q=1 then it becomes the usual Catalan number Cn=1n+1(2nn)C_{n}=\frac{1}{n+1}\binom{2n}{n}.

When the complex reflection group WW is a coincidental type, i.e. if {d1,,dn}\{d_{1},\ldots,d_{n}\} forms an arithmetic sequence, the corresponding qq-Catalan numbers are refined by another qq-sequence called qq-Narayana numbers. Following [RSS20], for given kk and tt we define

Nar(W,k,t;q)=q(tak1)(nk)[nk]qai=0k1(1qt1ai)i=0k1(1qe+1+ai).\operatorname{\textnormal{Nar}}(W,k,t;q)=q^{(t-ak-1)(n-k)}\genfrac{[}{]}{0.0pt}{}{n}{k}_{q^{a}}\frac{\prod_{i=0}^{k-1}(1-q^{t-1-ai})}{\prod_{i=0}^{k-1}(1-q^{e+1+ai})}.

where a,e>0a,e\in\mathbb{Z}_{>0} satisfies {d1,,dn}={e+ia0in1}\{d_{1},\ldots,d_{n}\}=\{e+ia\mid 0\leq i\leq n-1\} and [nk]qa=i=nk+1n[i]qa/i=1k[i]qa\genfrac{[}{]}{0.0pt}{}{n}{k}_{q^{a}}=\prod_{i=n-k+1}^{n}[i]_{q^{a}}/\prod_{i=1}^{k}[i]_{q^{a}}. This sequence then satisfies that Cat(W,t;q)=k=1nNar(W,k,t;q)\textnormal{Cat}(W,t;q)=\sum_{k=1}^{n}\operatorname{\textnormal{Nar}}(W,k,t;q). This decomposition also has an interpretation in terms of non-crossing partitions; if WW is the symmetric group permuting nn elements, t=n+1t=n+1, and q=1q=1 then Nar(W,k,n+1;1)\operatorname{\textnormal{Nar}}(W,k,n+1;1) enumerates the number of non-crossing partitions of nn with exactly kk pieces.

The main object in this paper is qq-Kreweras numbers, which refines qq-Narayana numbers when WW is of coincidental type. Let Irr(W)\operatorname{\textup{Irr}}(W) be the set of complex irreducible representations of WW up to isomorphism. For each χIrr(W)\chi\in\operatorname{\textup{Irr}}(W), we wish to find a systematic way to define Krew(W,χ,t;q)\operatorname{\textnormal{Krew}}(W,\chi,t;q), called the qq-Kreweras number attached to χ\chi, such that Nar(W,k,t;q)=f(χ)=kKrew(W,χ,t;q)\operatorname{\textnormal{Nar}}(W,k,t;q)=\sum_{f(\chi)=k}\operatorname{\textnormal{Krew}}(W,\chi,t;q) where f:Irr(W)={0,1,2,}f:\operatorname{\textup{Irr}}(W)\rightarrow\mathbb{N}=\{0,1,2,\ldots\} is a certain level function. Moreover, if WW is the symmetric group permuting nn elements, t=n+1t=n+1, and q=1q=1, so that χ\chi corresponds to some partition λn\lambda\vdash n, then we hope that Krew(W,χ,n+1;1)\operatorname{\textnormal{Krew}}(W,\chi,n+1;1) enumerates the number of non-crossing partitions where the sizes of parts are given by λ\lambda.

In [RS18], Reiner and Sommers defined a version of qq-Kreweras numbers for Weyl groups which satisfies nice combinatorial properties. More precisely, they defined qq-Kreweras numbers indexed by the nilpotent orbits in the corresponding Lie algebra instead of Irr(W)\operatorname{\textup{Irr}}(W) and show that they are refinements of qq-Narayana numbers when WW is of type A,B,A,B, and CC. Moreover, they proved that their qq-Kreweras numbers enjoy cyclic sieving phenomena with respect to the generalization of non-crossing partitions studied by Armstrong [Arm09].

Their argument is based on Springer theory. Namely, Sommers [Som11] studied the decomposition of a certain virtual graded WW-representation t\mathcal{H}_{t}, which naturally arises in the representation theory of double affine Hecke algebras, into the Green functions coming from the Springer theory of the corresponding reductive group. Then qq-Kreweras numbers are defined to be the “coefficients” in this decomposition. However, since the Springer correspondence (in type B,CB,C, and DD) from the set of nilpotent orbits (with respect to the trivial local system) to Irr(W)\operatorname{\textup{Irr}}(W) is not in general bijective but only injective, their qq-Kreweras numbers are not parametrized by Irr(W)\operatorname{\textup{Irr}}(W) but nilpotent orbits. Indeed, if one tries to expand their definition to all of Irr(W)\operatorname{\textup{Irr}}(W) using the Springer correspondence with respect to various local systems of nilpotent orbits, then some of qq-Kreweras numbers become identically zero.

One of the main goals of this paper is to overcome this limitation. There is a general method to calculate the Green functions, the algorithm originally developed by Shoji [Sho83] and Lusztig [Lus86], which is now commonly called the Lusztig-Shoji algorithm. Later it is generalized by Shoji [Sho01], [Sho02] so that it is applied to complex reflection groups G(e,p,n)G(e,p,n). In particular, he investigated the Green functions attached to so-called limit symbols [Sho04], which are combinatorially simpler than the usual Green functions.

Motivated from his work, when WW is a coincidental Coxeter group, i.e. of type An1A_{n-1}, BCnBC_{n}, H3H_{3}, or I2(m)I_{2}(m), we define the qq-Kreweras numbers similarly to the argument of Reiner and Sommers but using the Green functions attached to limit symbols. This definition has certain advantages compared to their work. First, our method allows to define qq-Kreweras numbers for type H3H_{3} and I2(m)I_{2}(m), which are not crystallographic so the Springer theory in the usual sense is not applicable. Moreover, in type BCnBC_{n} the qq-Kreweras numbers only depend on the type of WW whereas in [RS18] there are two different series coming from the Springer theory of SO2n+1SO_{2n+1} and Sp2nSp_{2n}, respectively. Also, our qq-Kreweras numbers are indexed by Irr(W)\operatorname{\textup{Irr}}(W), or more precisely none of qq-Kreweras numbers is identically zero. We believe that our version is combinatorially simpler and more uniform than their results.

The main results of this paper are summarized as follows. Here WW is a coincidental Coxeter group, i.e. of type An1A_{n-1}, BCnBC_{n}, H3H_{3}, or I2(m)I_{2}(m).

  1. For each χIrr(W)\chi\in\operatorname{\textup{Irr}}(W), we provide the closed formula of the qq-Kreweras number Krew(W,χ,t;q)\operatorname{\textnormal{Krew}}(W,\chi,t;q) (which also shows that it is not identically zero).

  2. Theorem 3.3: we show that Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{Z}[q] when tt is very good (see Theorem 3.2 for the definition of very good tt).

  3. Theorem 3.4: when tt is very good, we show that Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{N}[q] if and only if χ\chi corresponds to a parabolic subgroup of WW (see 3.3 for the correspondence Φ\Phi from parabolic subgroups from WW to Irr(W)\operatorname{\textup{Irr}}(W)).

  4. Theorem 3.5: we show that Krew(W,χ,t;q)\operatorname{\textnormal{Krew}}(W,\chi,t;q) refines the qq-Narayana numbers.

  5. Theorem 3.6: we prove that Krew(W,χ,t;q)\operatorname{\textnormal{Krew}}(W,\chi,t;q) exhibits certain cyclic sieving phenomena with respect to the generalization of non-crossing partitions defined by Armstrong [Arm09].

Our argument is given case-by-case. For type An1A_{n-1}, the Green functions attached to limit symbols are the same as the ones coming from the usual Springer theory, and thus there is nothing new compared to [RS18]. For type BCnBC_{n}, we exploit the Springer theory of LieSp2n(𝔽2¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{2}}) and the exotic nilcone, which are known to have a strong connection with the Green functions attached to limit symbols. For type H3H_{3} and I2(m)I_{2}(m), our proof mainly relies on direct calculation. For these results, we use the computer program packages such as [GHLMP96], [Ach19], and [Sag20].

In this paper we only discuss coincidental Coxeter groups. However, the Green functions attached to limit symbols are well-defined for any complex reflection group of the form G(r,p,n)G(r,p,n). Therefore, it is natural to consider a generalization of our results to such complex reflection groups, or at least for coincidental types. Moreover, even when we restrict our attention to Coxeter groups, we may choose different symbols to perform the Lusztig-Shoji algorithm, which provides different kinds of Green functions. It would be interesting to ask for which choice of symbols such qq-Kreweras numbers are well-defined and enjoy nice combinatorial properties.

This paper is organized as follows. In Section 1 we discuss basic definitions and notations used throughout this paper; in Section 2 we recall the definition and the classification of coincidental Coxeter groups and parametrizations of their complex irreducible representations; in Section 3 we study the definition and the properties of qq-Kreweras numbers and state the main theorems; from Section 4 to the end we prove the main theorems case-by-case.

Acknowledgement.

The author is grateful to Eric Sommers for his helpful comments and suggestions. In particular, Section 6 and 7 would not be written without his suggestion to look into other coincidental ones than type AA and BCBC.

1. Definitions and notations

For a,ba,b\in\mathbb{Z}, we set [a,b]={xaxb}[a,b]=\{x\in\mathbb{Z}\mid a\leq x\leq b\}. For xx\in\mathbb{R}, we set x\lfloor x\rfloor to be the largest integer not greater than xx and x\lceil x\rceil to be the smallest integer not smaller than xx. In other words, we have x,x\lfloor x\rfloor,\lceil x\rceil\in\mathbb{Z}, xx<x\lfloor x\rfloor\leq x<\lfloor x\rfloor, and x1<xx\lceil x\rceil-1<x\leq\lceil x\rceil.

For mm\in\mathbb{N}, we define [m]y=i=0m1yi[m]_{y}=\sum_{i=0}^{m-1}y^{i}. If y1y\neq 1 then we have [m]y=(ym1)/(y1)[m]_{y}=(y^{m}-1)/(y-1). We set [m]y!=i=1m[m]y[m]_{y}!=\prod_{i=1}^{m}[m]_{y} and

[m1+m2++mkm1,m2,,mk]y=[m1+m2++mk]y![m1]y![m2]y![mk]y!.\genfrac{[}{]}{0.0pt}{}{m_{1}+m_{2}+\cdots+m_{k}}{m_{1},m_{2},\ldots,m_{k}}_{y}=\frac{[m_{1}+m_{2}+\cdots+m_{k}]_{y}!}{[m_{1}]_{y}![m_{2}]_{y}!\cdots[m_{k}]_{y}!}.

When k=2k=2, we also write [m1+m2m1]y=[m1+m2m2]y\genfrac{[}{]}{0.0pt}{}{m_{1}+m_{2}}{m_{1}}_{y}=\genfrac{[}{]}{0.0pt}{}{m_{1}+m_{2}}{m_{2}}_{y} instead of [m1+m2m1,m2]y\genfrac{[}{]}{0.0pt}{}{m_{1}+m_{2}}{m_{1},m_{2}}_{y}. When yy equals qq or q2q^{2}, we also write [m]q=[m][m]_{q}=[m], [m]q2=m[m]_{q^{2}}=\llbracket m\rrbracket, [m]q!=[m]![m]_{q}!=[m]!, [m]q2!=m![m]_{q^{2}}!=\llbracket m\rrbracket!, etc. Note that [m]y,[m]y!,[m1+m2++mkm1,m2,,mk]y[m]_{y},[m]_{y}!,\genfrac{[}{]}{0.0pt}{}{m_{1}+m_{2}+\cdots+m_{k}}{m_{1},m_{2},\ldots,m_{k}}_{y} are elements in [y]\mathbb{N}[y], i.e. polynomials in yy with nonnegative integer coefficients.

A partitions is an integer sequence λ=(λ1,,λl)\lambda=(\lambda_{1},\ldots,\lambda_{l}) such that λ1λl>0\lambda_{1}\geq\cdots\geq\lambda_{l}>0. In such a case we define its length to be l(λ)=ll(\lambda)=l and its size to be |λ|=i=1lλi|\lambda|=\sum_{i=1}^{l}\lambda_{i}. If i>li>l then we set λi=0\lambda_{i}=0. We set the weighted size of λ\lambda to be z(λ)=i1(i1)λi=j1λjT(λjT1)/2z(\lambda)=\sum_{i\geq 1}(i-1)\lambda_{i}=\sum_{j\geq 1}\lambda_{j}^{T}(\lambda_{j}^{T}-1)/2 where λT=(λ1T,λ2T,)\lambda^{T}=(\lambda_{1}^{T},\lambda_{2}^{T},\ldots) is the conjugate partition of λ\lambda. We write 𝔓n\mathfrak{P}_{n} to be the set of partitions of size nn. For n>0n\in\mathbb{Z}_{>0} and λ=(λ1,,λl)\lambda=(\lambda_{1},\ldots,\lambda_{l}), we set λn=(λn,,λl)\lambda_{\geq n}=(\lambda_{n},\ldots,\lambda_{l}). For two partitions λ=(λ1,λ2,)\lambda=(\lambda_{1},\lambda_{2},\ldots) and μ=(μ1,μ2,)\mu=(\mu_{1},\mu_{2},\ldots), we define λ+μ=(λ1+μ1,λ2+μ2,)\lambda+\mu=(\lambda_{1}+\mu_{1},\lambda_{2}+\mu_{2},\ldots). Also we define λμ\lambda\cup\mu to be the partition obtained by combining parts of λ\lambda and μ\mu and rearranging them so that the result is in decreasing order. Let mλ(r)m_{\lambda}(r) be the multiplicity of rr in λ\lambda. We often write (1mλ(1)2mλ(2))(1^{m_{\lambda}(1)}2^{m_{\lambda}(2)}\cdots) to indicate λ\lambda. If mλ(r)0m_{\lambda}(r)\neq 0 then we also write rλr\in\lambda.

Let 𝔓n,2\mathfrak{P}_{n,2} to be the set of pairs of partitions (λ,μ)(\lambda,\mu) such that |λ|+|μ|=n|\lambda|+|\mu|=n. Its elements are called bipartitions of nn. Here we list some notations which will be used in Section 5 for the sake of readers’ convenience.

l(μ,ν)\displaystyle l(\mu,\nu) =max{l(μ)1,l(ν)}=l(μ2+ν)\displaystyle=\max\{l(\mu)-1,l(\nu)\}=l(\mu_{\geq 2}+\nu)
mμ,ν(r)\displaystyle m_{\mu,\nu}(r) =m(μ+ν)(μ2+ν)(r)2\displaystyle=\left\lfloor\frac{m_{(\mu+\nu)\cup(\mu_{\geq 2}+\nu)}(r)}{2}\right\rfloor
L(μ,ν)\displaystyle L(\mu,\nu) =l(μ,ν)r1mμ,ν(r)\displaystyle=l(\mu,\nu)-\sum_{r\geq 1}m_{\mu,\nu}(r)
z(μ,ν)\displaystyle z(\mu,\nu) =2z(μ)+2z(ν)+|ν|=2z(μ+ν)+|ν|\displaystyle=2z(\mu)+2z(\nu)+|\nu|=2z(\mu+\nu)+|\nu|
d(μ,ν)\displaystyle d(\mu,\nu) =r1mμ,ν(r)(mμ,ν(r)+1)\displaystyle=\sum_{r\geq 1}m_{\mu,\nu}(r)(m_{\mu,\nu}(r)+1)

For a finite group GG, let Irr(G)\operatorname{\textup{Irr}}(G) be the set of complex irreducible representations of GG (up to isomorphism). Also we let 𝒦(G)\mathcal{K}(G) be its Grothendieck group which is a free \mathbb{Z}-module with basis Irr(G)\operatorname{\textup{Irr}}(G), and 𝒦(G)+\mathcal{K}(G)^{+} be the submonoid of 𝒦(G)\mathcal{K}(G) generated by Irr(G)\operatorname{\textup{Irr}}(G). For any ring RR, we write 𝒦(G)R=𝒦(G)R\mathcal{K}(G)_{R}=\mathcal{K}(G)\otimes R.

2. Coincidental Coxeter groups

2.1. Classification of coincidental Coxeter groups

Let WW be a Coxeter group and e1,,ene_{1},\ldots,e_{n} be its exponents such that e1<<ene_{1}<\cdots<e_{n}. We say that WW is coincidental if e1,,ene_{1},\ldots,e_{n} is an arithmetic sequence, i.e. there exists a>0a\in\mathbb{Z}_{>0} such that ei=e1+(i1)ae_{i}=e_{1}+(i-1)a for i[1,n]i\in[1,n]. Such groups are classified as follows.

  1. (1)

    Type An1A_{n-1}: exponents are 1,2,,n11,2,\ldots,n-1 and the Coxeter number is nn.

  2. (2)

    Type BCnBC_{n}: exponents are 1,3,,2n11,3,\ldots,2n-1 and the Coxeter number is 2n2n.

  3. (3)

    Type H3H_{3}: exponents are 1,5,91,5,9 and the Coxeter number is 1010.

  4. (4)

    Type I2(m)I_{2}(m): exponents are 1,m11,m-1 and the Coxeter number is mm.

Throughout this paper, we assume that WW is one of these groups unless otherwise specified.

2.2. Irreducible representations

Let us discuss the parametrization of irreducible representations of WW (up to isomorphism). Here we adopt Carter’s notations, see [Car93, Chapter 13.2]. First suppose that WW is of type An1A_{n-1}, so WW is the symmetric group permuting nn elements. Then its irreducible representations are parametrized by partitions of nn. Let us write χλ\chi^{\lambda} to denote such a representation parametrized by λ\lambda. For example, χ(n)\chi^{(n)} is the identity representation and χ(1n)\chi^{(1^{n})} is the sign representation.

This time suppose that WW is of type BCnBC_{n}, so WW is the hyperoctahedral group of rank nn. In this case its irreducible representations are parametrized by bipartitions of nn. Let us write χ(μ,ν)\chi^{(\mu,\nu)} to denote such a representation parametrized by (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2}. For example, χ((n),)\chi^{((n),\emptyset)} is the identity representation and χ(,(1n))\chi^{(\emptyset,(1^{n}))} is the sign representation.

Before we proceed let us define the fake degree of an irreducible representation. Let S(V)=iSi(V)qiS^{*}(V)=\bigoplus_{i\in\mathbb{N}}S^{i}(V)q^{i} be the symmetric algebra of the reflection representation of WW which is a graded WW-representation such that each Si(V)S^{i}(V) is of degree ii. (Here, qq is a degree-keeping indeterminate.) Then for each χIrr(W)\chi\in\operatorname{\textup{Irr}}(W), there exists an integer ee\in\mathbb{N} such that χ,Si(V)0ei\langle\chi,S^{i}(V)\rangle\neq 0\Rightarrow e\leq i and χ,Se(V)=1\langle\chi,S^{e}(V)\rangle=1. We say that ee is the fake degree of χ\chi and write b(χ)=eb(\chi)=e.

When WW is of type H3H_{3}, each irreducible representation is completely determined by its dimension and fake degree. We write ϕd,e\phi_{d,e} to be an irreducible representation such that dimϕd,e=d\dim\phi_{d,e}=d and b(ϕd,e)=eb(\phi_{d,e})=e. Then we have:

Irr(W)={ϕ1,0,ϕ3,1,ϕ5,2,ϕ4,3,ϕ3,3,ϕ4,4,ϕ5,5,ϕ3,6,ϕ3,8,ϕ1,15}.\operatorname{\textup{Irr}}(W)=\{\phi_{1,0},\phi_{3,1},\phi_{5,2},\phi_{4,3},\phi_{3,3},\phi_{4,4},\phi_{5,5},\phi_{3,6},\phi_{3,8},\phi_{1,15}\}.

Now suppose that WW is of type I2(m)I_{2}(m). If mm is odd then similarly we have

Irr(W)={ϕ1,0,ϕ1,m}{ϕ2,rr[1,(m1)/2]}.\operatorname{\textup{Irr}}(W)=\{\phi_{1,0},\phi_{1,m}\}\cup\{\phi_{2,r}\mid r\in[1,(m-1)/2]\}.

If mm is even, then there are two irreducible representations both of whose dimension is 11 and fake degree is m/2m/2. Here we write ϕ1,m/2,ϕ1,m/2′′\phi_{1,m/2}^{\prime},\phi_{1,m/2}^{\prime\prime} to distinguish these two. Then we have

Irr(W)={ϕ1,0,ϕ1,m/2,ϕ1,m/2′′,ϕ1,m}{ϕ2,rr[1,m/21]}.\operatorname{\textup{Irr}}(W)=\{\phi_{1,0},\phi_{1,m/2}^{\prime},\phi_{1,m/2}^{\prime\prime},\phi_{1,m}\}\cup\{\phi_{2,r}\mid r\in[1,m/2-1]\}.

3. qq-Kreweras numbers attached to limit symbols

We keep assuming that WW is a coincidental Coxeter group. Here we provide a definition of qq-Kreweras numbers attached to limit symbols. To this end we start with the corresponding Green functions of WW.

3.1. Green functions attached to limit symbols

Recall the Lusztig-Shoji algorithm with respect to limit symbols as follows. We choose a total order on Irr(G)\operatorname{\textup{Irr}}(G), say p\leq_{p}, which satisfies that χpχb(χ)b(χ)\chi\leq_{p}\chi^{\prime}\Rightarrow b(\chi)\geq b(\chi^{\prime}). In each type this order is chosen as follows.

  1. (1)

    Type An1A_{n-1}: Irr(W)\operatorname{\textup{Irr}}(W) is parametrized by 𝔓n\mathfrak{P}_{n}. Here we choose any linearization of dominance order on 𝔓n\mathfrak{P}_{n}.

  2. (2)

    Type BCnBC_{n}: Irr(W)\operatorname{\textup{Irr}}(W) is parametrized by 𝔓n,2\mathfrak{P}_{n,2}. To each bipartition (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2} we attach a sequence (μ1,ν1,μ2,ν2,)(\mu_{1},\nu_{1},\mu_{2},\nu_{2},\ldots) which is not necessarily a partition. Then we choose any linearization of dominance order on these sequences.

  3. (3)

    Type H3H_{3}: we choose the order as follows.

    ϕ1,0>pϕ3,1>pϕ5,2>pϕ4,3>pϕ3,3>pϕ4,4>pϕ5,5>pϕ3,6>pϕ3,8>pϕ1,15\phi_{1,0}>_{p}\phi_{3,1}>_{p}\phi_{5,2}>_{p}\phi_{4,3}>_{p}\phi_{3,3}>_{p}\phi_{4,4}>_{p}\phi_{5,5}>_{p}\phi_{3,6}>_{p}\phi_{3,8}>_{p}\phi_{1,15}
  4. (4)

    Type I2(m)I_{2}(m): if mm is odd, then the order is uniquely determined. If mm is even then there are two possible choices of linear orders and we take either one.

Now we run the Lusztig-Shoji algorithm following [Sho01, Sho02, Sho04] with respect to the total order described above. If WW is of type An1A_{n-1} or BCnBC_{n}, this process is exactly the same as described in [Sho04]. In the case of H3H_{3} and I2(m)I_{2}(m), there does not exist such a “limit symbol” but we may still follow the argument therein to calculate the Green functions with respect to the order described above. By abuse of terminology, we still say that these are the Green functions attached to limit symbols.

Remark.

Since we use a total order to perform the Lusztig-Shoji algorithm, the corresponding “PP-matrix” is not only block triangular but indeed triangular, i.e. each block consists of one irreducible representation. Moreover, the “aa-function” of each irreducible representation is equal to its fake degree.

Let Qχ𝒦(W)(q)Q_{\chi}\in\mathcal{K}(W)_{\mathbb{Q}(q)} be the Green function of χ\chi as a result of the aforementioned algorithm. In particular, as a graded WW-representation with degq=1\deg q=1, QIdQ_{Id} is isomorphic to the coinvariant algebra of WW. In crystallographic cases, it is the total cohomology of the corresponding flag variety with the usual Springer WW-action where the degree of the 2i2i-th cohomology group is ii. Then it is known that {QχχIrr(W)}\{Q_{\chi}\mid\chi\in\operatorname{\textup{Irr}}(W)\} is a basis of 𝒦(W)(q)\mathcal{K}(W)_{\mathbb{Q}(q)}. Moreover, in our case we have the following stronger statement.

Theorem 3.1.

Suppose that WW is a coincidental Coxeter group. Then for χIrr(W)\chi\in\operatorname{\textup{Irr}}(W), QχQ_{\chi} (with respect to the limit symbol) is contained in 𝒦(W)+[q]\mathcal{K}(W)^{+}[q]. In other words, QχQ_{\chi} is a genuine \mathbb{N}-graded WW-representation.

Proof.

For type An1A_{n-1}, QχQ_{\chi} is equal to the Green function coming from the usual Springer theory (or Green polynomials), in which case the result is well-known. For type BCnBC_{n}, it follows from [AH08, Corollary 5.3]. For type H3H_{3}, it follows from direct calculation. For type I2(m)I_{2}(m), it follows from [AA08, Theorem 3]. ∎

3.2. qq-Kreweras numbers attached to limit symbols

Here we define the qq-Kreweras numbers attached to limit symbols similarly to [RS18, Section 1.3]. To this end, first we introduce a virtual graded WW-representation

t=i=0n(qt)iS(V)iV𝒦(W)q\mathcal{H}_{t}=\bigoplus_{i=0}^{n}(-q^{t})^{i}S^{*}(V)\otimes\wedge^{i}V\in\mathcal{K}(W)\llbracket q\rrbracket

where tt is a positive integer, nn is the rank of WW, VV is the reflection representation of WW, and S(V)=iSi(V)qiS^{*}(V)=\bigoplus_{i\in\mathbb{N}}S^{i}(V)q^{i} is its symmetric algebra as a graded representation of WW. By Chevalley’s result, we have S(V)=QId/i=1n(1qdi)S^{*}(V)=Q_{Id}/\prod_{i=1}^{n}(1-q^{d_{i}}) where {d1,,dn}\{d_{1},\ldots,d_{n}\} are the fundamental degrees of WW. Thus we have t𝒦(W)(q)\mathcal{H}_{t}\in\mathcal{K}(W)_{\mathbb{Q}(q)}. Moreover, we have t𝒦(W)+[q]\mathcal{H}_{t}\in\mathcal{K}(W)^{+}[q] for certain tt as the following theorem states.

Theorem 3.2.

For t>0t\in\mathbb{Z}_{>0}, we say that tt is very good if it satisfies the following condition.

  1. (1)

    Type An1A_{n-1}: gcd(t,n)=1\gcd(t,n)=1.

  2. (2)

    Type BCnBC_{n}: tt is odd.

  3. (3)

    Type H3H_{3}: t1,5,9(mod10)t\equiv 1,5,9\pmod{10}.

  4. (4)

    Type I2(m)I_{2}(m): t±1(modm)t\equiv\pm 1\pmod{m}.

Then, we have t𝒦(W)+[q]\mathcal{H}_{t}\in\mathcal{K}(W)^{+}[q] if and only if tt is very good. In particular, tt is very good if t1t-1 is a multiple of the Coxeter number of WW.

Proof.

The proof of [Som11, Proposition 13] still applies here (even when WW is of type H3H_{3} or I2(m)I_{2}(m)). ∎

Since {QχχIrr(W)}\{Q_{\chi}\mid\chi\in\operatorname{\textup{Irr}}(W)\} is a basis of 𝒦(W)(q)\mathcal{K}(W)_{\mathbb{Q}(q)}, t\mathcal{H}_{t} can be uniquely written as a linear combination of QχQ_{\chi}. Let us define the qq-Kreweras numbers, Krew(W,χ,t;q)(q)\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{Q}(q), to be such that the following equation holds:

t=χIrr(W)Krew(W,χ,t;q)Qχ.\mathcal{H}_{t}=\bigoplus_{\chi\in\operatorname{\textup{Irr}}(W)}\operatorname{\textnormal{Krew}}(W,\chi,t;q)Q_{\chi}.

We are especially interested in the case when tt is very good. The following theorem will be shown case-by-case in later sections.

Theorem 3.3.

For tt very good, we have Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{Z}[q] for any χIrr(W)\chi\in\operatorname{\textup{Irr}}(W).

3.3. Positivity of qq-Kreweras numbers

In general, even if Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{Z}[q], it does not always hold that Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{N}[q]. Let us state the sufficient and necessary condition for the positivity of the qq-Kreweras numbers. To this end, we define a map Φ\Phi from the set of (the conjugacy classes of) parabolic subgroups in WW to Irr(W)\operatorname{\textup{Irr}}(W).

If WW is of type An1A_{n-1}, then a parabolic subgroup is of the form 𝔖a1××𝔖ak\operatorname{\mathfrak{S}}_{a_{1}}\times\cdots\times\operatorname{\mathfrak{S}}_{a_{k}} for some a1,,ak>0a_{1},\ldots,a_{k}\in\mathbb{Z}_{>0} such that a1++ak=na_{1}+\cdots+a_{k}=n. (Here, 𝔖a\operatorname{\mathfrak{S}}_{a} is the symmetric group permuting aa elements.) Let λ\lambda be the partition of nn obtained by rearranging a1,,aka_{1},\ldots,a_{k} if necessarly. Then we set Φ(P)=χλ\Phi(P)=\chi^{\lambda}. Note that this is also equivalent to the “principal-in-a-Levi” condition of [RS18]. Indeed, if we let 𝒪λLieGLn()\mathcal{O}_{\lambda}\subset\operatorname{\textup{Lie}}GL_{n}(\mathbb{C}) be the nilpotent orbit which contains a regular nilpotent element in the Levi subalgebra corresponding to PWP\subset W, then the (usual) Springer correspondence sends 𝒪λ\mathcal{O}_{\lambda} to χλ\chi^{\lambda}.

If WW is of type BCnBC_{n}, then any parabolic subgroup PWP\subset W is isomorphic to Hb×𝔖a1××𝔖akH_{b}\times\operatorname{\mathfrak{S}}_{a_{1}}\times\cdots\times\operatorname{\mathfrak{S}}_{a_{k}} where HbH_{b} is a Coxeter group of type BCbBC_{b} and b+a1++ak=nb+a_{1}+\cdots+a_{k}=n. (Here bb can be zero.) In this case, without loss of generality we may assume that a1am>bam+1aka_{1}\geq\cdots\geq a_{m}>b\geq a_{m+1}\geq\cdots\geq a_{k} for some 0mk0\leq m\leq k. We set

μ=(b,b,,b,am+1,,ak) and ν=(a1b,a2b,,amb)\mu=(b,b,\ldots,b,a_{m+1},\ldots,a_{k})\quad\textnormal{ and }\quad\nu=(a_{1}-b,a_{2}-b,\ldots,a_{m}-b)

where there are m+1m+1 bb’s in μ\mu. Now we set Φ(P)=χ(μ,ν)\Phi(P)=\chi^{(\mu,\nu)}.

Remark.

This correspondence is different from the usual Springer correspondence of either SO2n+1()SO_{2n+1}(\mathbb{C}) or Sp2n()Sp_{2n}(\mathbb{C}) that is used in [RS18]. Rather, here we exploit the Springer correspondence of LieSp2n\operatorname{\textup{Lie}}Sp_{2n} over characteristic 2. This is explained in Section 5 in more detail.

If WW is of type H3H_{3}, then we are no longer able to argue using the Levi subalgebra of some Lie algebra. Instead, for each PWP\subset W there exists a unique χIrr(W)\chi\in\operatorname{\textup{Irr}}(W) which appears in IndPWIdP\operatorname{\textup{Ind}}_{P}^{W}\operatorname{\textup{Id}}_{P} with the highest fake degree. (For such a representation we have IndPWIdP,χ=1\langle\operatorname{\textup{Ind}}_{P}^{W}\operatorname{\textup{Id}}_{P},\chi\rangle=1.) Then we set Φ(P)=χ\Phi(P)=\chi. We list the types of parabolic subgroups of WW and their images under Φ\Phi:

{id}ϕ1,15,\displaystyle\{id\}\mapsto\phi_{1,15}, A1ϕ3,8,\displaystyle A_{1}\mapsto\phi_{3,8}, A1×A1ϕ5,5,\displaystyle A_{1}\times A_{1}\mapsto\phi_{5,5},
A2ϕ4,4,\displaystyle A_{2}\mapsto\phi_{4,4}, I2(5)ϕ3,3,\displaystyle I_{2}(5)\mapsto\phi_{3,3}, H3ϕ1,0.\displaystyle H_{3}\mapsto\phi_{1,0}.

Note that there are three different parabolic subgroups of type A1A_{1}; all of them are mapped to the same representation ϕ3,8\phi_{3,8}.

When WW is of type I2(m)I_{2}(m) we define Φ\Phi similarly to above. If mm is odd then we have:

{id}ϕ1,m,A1ϕ2,(m1)/2,I2(m)ϕ1,0.\{id\}\mapsto\phi_{1,m},\quad A_{1}\mapsto\phi_{2,(m-1)/2},\quad I_{2}(m)\mapsto\phi_{1,0}.

If mm is even then we have:

{id}ϕ1,m,A1ϕ1,m/2,A1′′ϕ1,m/2′′,I2(m)ϕ1,0.\{id\}\mapsto\phi_{1,m},\quad A_{1}^{\prime}\mapsto\phi_{1,m/2}^{\prime},\quad A_{1}^{\prime\prime}\mapsto\phi_{1,m/2}^{\prime\prime},\quad I_{2}(m)\mapsto\phi_{1,0}.

Here A1A_{1}^{\prime} and A1′′A_{1}^{\prime\prime} indicate two different parabolic subgroups, respectively, generated by each simple reflection of I2(m)I_{2}(m).

Now we state the positivity theorem of qq-Kreweras numbers. Its proof is given in later sections case-by-case.

Theorem 3.4 (cf. [RS18, Theorem 1.6]).

For χIrr(W)\chi\in\operatorname{\textup{Irr}}(W) and very good tt, the corresponding qq-Kreweras number Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{Z}[q] has nonnegative integer coefficients if and only if χimΦ\chi\in\operatorname{\textup{im}}\Phi.

3.4. Relation with qq-Narayana numbers

Recently Reiner-Shepler-Sommers [RSS20] defined the qq-Narayana numbers for any complex coincidental reflection group. Here we recall their definition. Suppose that the exponents of WW are e,e+a,,e+(n1)ae,e+a,\ldots,e+(n-1)a for some e,ae,a\in\mathbb{N} where nn is the rank of WW. Then the qq-Narayana number with parameters kk and tt is defined to be

Nar(W,k,t;q)=q(tak1)(nk)[nk]qai=0k1(1qt1ai)i=0k1(1qe+1+ai).\operatorname{\textnormal{Nar}}(W,k,t;q)=q^{(t-ak-1)(n-k)}\genfrac{[}{]}{0.0pt}{}{n}{k}_{q^{a}}\frac{\prod_{i=0}^{k-1}(1-q^{t-1-ai})}{\prod_{i=0}^{k-1}(1-q^{e+1+ai})}.

This definition also coincides with the ones given in [RS18] for type An1A_{n-1} and BCnBC_{n}.

In the following sections, we prove the following theorem case-by-case.

Theorem 3.5 (cf. [RS18, Definition 1.9]).

For a coincidental Coxeter group WW we have

Nar(W,k,t;q)=Qχ,Vq=1=kKrew(W,χ,t;q).\operatorname{\textnormal{Nar}}(W,k,t;q)=\sum_{\langle Q_{\chi},V\rangle_{q=1}=k}\operatorname{\textnormal{Krew}}(W,\chi,t;q).

Here, Qχ,Vq=1\langle Q_{\chi},V\rangle_{q=1} is the (ungraded) multiplicity of the reflection representation VV of WW in QχQ_{\chi}.

Remark.

In [RS18], the qq-Narayana numbers are defined to be the sum of qq-Kreweras numbers (with respect to the usual Springer theory) similar to the above theorem and the definition in [RSS20] is a theorem therein.

3.5. Cyclic sieving

In [RS18] it was conjectured, and proved for classical types (with respect to the usual Springer theory), that the qq-Kreweras numbers exhibit certain cyclic sieving phenomena. To this end, first we introduce the notion of (chains of) non-crossing partitions for general Coxeter groups. For a Coxeter system (W,S={s1,,sn})(W,S=\{s_{1},\ldots,s_{n}\}), we fix a (standard) Coxeter element, say c=s1s2snc=s_{1}s_{2}\cdots s_{n}. Let Ref(W)\operatorname{\textnormal{Ref}}(W) be the set of all reflections in WW. (This set is in general strictly larger than SS.) Define the absolute length of wWw\in W, say la(w)l^{a}(w), to be the minimum number of reflections whose product is ww. We define an order on WW to be the closure of the cover relations wawrw\leq_{a}wr where rRef(W)r\in\operatorname{\textnormal{Ref}}(W) and la(wr)=la(w)+1l^{a}(wr)=l^{a}(w)+1. Now we define

NC(s)(W)={(w1,,ws)Ww1aawsac}.\operatorname{\textnormal{NC}}^{(s)}(W)=\{(w_{1},\ldots,w_{s})\in W\mid w_{1}\leq_{a}\cdots\leq_{a}w_{s}\leq_{a}c\}.

This set depends on the choice of cc but they are all equivalent under conjugation.

Remark.

When WW is a symmetric group permuting nn elements, NC(1)(W)\operatorname{\textnormal{NC}}^{(1)}(W) is equivalent to non-crossing partitions of nn elements originally introduced by Kreweras [Kre72]. Later it is vastly generalized and the definition above is adopted from [Arm09], to which we refer readers for more details.

With (w1,,ws)NC(s)(W)(w_{1},\ldots,w_{s})\in\operatorname{\textnormal{NC}}^{(s)}(W) we associate a sequence

(δ1,,δs1,δs)=(w11w2,,ws11ws,ws1c)(\delta_{1},\ldots,\delta_{s-1},\delta_{s})=(w_{1}^{-1}w_{2},\ldots,w_{s-1}^{-1}w_{s},w_{s}^{-1}c)

called a δ\delta-sequence in [Arm09]. Note that such a δ\delta-sequence uniquely determines (w1,,ws)(w_{1},\ldots,w_{s}). Now we define a /sh\mathbb{Z}/sh-action on NC(s)(W)\operatorname{\textnormal{NC}}^{(s)}(W) (where hh is the Coxeter number of WW) so that in terms of a δ\delta-sequence it is described as

(δ1,,δs1,δs)(cδsc1,δ1,,δs1).(\delta_{1},\ldots,\delta_{s-1},\delta_{s})\mapsto(c\delta_{s}c^{-1},\delta_{1},\ldots,\delta_{s-1}).

This action is indeed well-defined and clearly a /sh\mathbb{Z}/sh-action. (See [Arm09, 3.4] for detailed discussion.)

Recall that when tt is very good, Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{N}[q] if and only if χimΦ\chi\in\operatorname{\textup{im}}\Phi. For a parabolic subgroup PWP\subset W, let VPV^{P} be the set of point-wise fixed points in the reflection representation VV by elements in PP, and let WVP={wVPwW}={VwPw1wW}W\cdot V^{P}=\{w\cdot V^{P}\mid w\in W\}=\{V^{wPw^{-1}}\mid w\in W\}. The following theorem is shown case-by-case in later sections.

Theorem 3.6 (cf. [RS18, Conjecure 1.4]).

Let PP be a parabolic subgroup of WW. For s,ds,d\in\mathbb{N} such that dsh+1d\mid sh+1 where hh is the Coxeter number of WW, the specialization Krew(W,Φ(P),sh+1;ωd)\operatorname{\textnormal{Krew}}(W,\Phi(P),sh+1;\omega_{d}) is equal to the number of elements in

{(w1,,ws)NC(s)(W)Vw1WVP}\{(w_{1},\ldots,w_{s})\in\operatorname{\textnormal{NC}}^{(s)}(W)\mid V^{w_{1}}\in W\cdot V^{P}\}

fixed by the order dd element in /sh\mathbb{Z}/sh with respect to the cyclic action described above. Here, ωd×\omega_{d}\in\mathbb{C}^{\times} is a primitive dd-th root of unity.

4. Type An1A_{n-1}

In type AA, the Green functions attached to limit symbols are the same as the usual Green functions (or Green polynomials). Thus, in this case everything is already covered by the results of Reiner and Sommers [RS18]. Here we review their work for the sake of readers’ convenience.

4.1. qq-Kreweras numbers and positivity

Suppose that WW is of type An1A_{n-1}. Then for t>0t\in\mathbb{Z}_{>0} and λn\lambda\vdash n, we have

Krew(W,χλ,t;q)=qt(nl(λ))c(λ)1[t][ttl(λ),mλ(>0)]\operatorname{\textnormal{Krew}}(W,\chi^{\lambda},t;q)=q^{t(n-l(\lambda))-c(\lambda)}\frac{1}{[t]}\genfrac{[}{]}{0.0pt}{}{t}{t-l(\lambda),m_{\lambda}(\mathbb{Z}_{>0})}

where [ttl(λ),mλ(>0)]=[ttl(λ),mλ(1),mλ(2),].\genfrac{[}{]}{0.0pt}{}{t}{t-l(\lambda),m_{\lambda}(\mathbb{Z}_{>0})}=\genfrac{[}{]}{0.0pt}{}{t}{t-l(\lambda),m_{\lambda}(1),m_{\lambda}(2),\ldots}. Here, c(λ)=i1λiTλi+1Tc(\lambda)=\sum_{i\geq 1}\lambda^{T}_{i}\lambda^{T}_{i+1} where λT\lambda^{T} is the conjugate partition of λ\lambda. In particular, if gcd(t,n)=1\gcd(t,n)=1 then we have Krew(W,χλ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi^{\lambda},t;q)\in\mathbb{N}[q]. (Note that in this case imΦ=Irr(W)\operatorname{\textup{im}}\Phi=\operatorname{\textup{Irr}}(W).)

4.2. Relation with qq-Narayana numbers

When WW is of type An1A_{n-1}, for k,tk,t\in\mathbb{N} such that 0kn10\leq k\leq n-1 and gcd(t,n)=1\gcd(t,n)=1 the qq-Narayana number with parameters k,tk,t is given by

Nar(W,k,t;q)=q(n1k)(t1k)1[k+1][n1k][t1k].\operatorname{\textnormal{Nar}}(W,k,t;q)=q^{(n-1-k)(t-1-k)}\frac{1}{[k+1]}\genfrac{[}{]}{0.0pt}{}{n-1}{k}\genfrac{[}{]}{0.0pt}{}{t-1}{k}.

Then by [RS18, Definition 1.9, Theorem 1.10] we have

Nar(W,k,t;q)=l(λ)=k+1Krew(W,χλ,t;q).\operatorname{\textnormal{Nar}}(W,k,t;q)=\sum_{l(\lambda)=k+1}\operatorname{\textnormal{Krew}}(W,\chi^{\lambda},t;q).

Here, l(λ)=k+1l(\lambda)=k+1 if and only if Qχλ,V|q=1=k\langle Q_{\chi^{\lambda}},V\rangle|_{q=1}=k.

4.3. Cyclic sieving

We recall [RS18, Section 6]. The Coxeter number of WW of type An1A_{n-1} is nn, and here we assume that t=ns+1t=ns+1 for some ss\in\mathbb{N}. Suppose that PWP\subset W is a parabolic subgroup such that Φ(P)=χλ\Phi(P)=\chi^{\lambda}. In this case, for dsh+1d\mid sh+1 we have Krew(W,χλ,sh+1;ωd)0\operatorname{\textnormal{Krew}}(W,\chi^{\lambda},sh+1;\omega_{d})\neq 0 if and only if [dmλ(r)d\mid m_{\lambda}(r) for all r>0r\in\mathbb{Z}_{>0}] or [there exists a unique rr^{\prime} such that dmλ(r)d\nmid m_{\lambda}(r^{\prime}) and it also satisfies dmλ(r)1d\mid m_{\lambda}(r^{\prime})-1]. In this case we have

Krew(W,χλ,sh+1;ωd)={1m(ssl(λ),mλ(>0)) if d=1,(ns/dns/dl(λ)/d,mλ(>0)/d) otherwise.\operatorname{\textnormal{Krew}}(W,\chi^{\lambda},sh+1;\omega_{d})=\left\{\begin{aligned} &\frac{1}{m}\binom{s}{s-l(\lambda),m_{\lambda}(\mathbb{Z}_{>0})}&\textnormal{ if }d=1,\\ &\binom{ns/d}{ns/d-\lfloor l(\lambda)/d\rfloor,\lfloor m_{\lambda}(\mathbb{Z}_{>0})/d\rfloor}&\textnormal{ otherwise.}\end{aligned}\right.

This is indeed the number of fixed points in {(w1,,ws)NC(s)(W)Vw1WVP}\{(w_{1},\ldots,w_{s})\in\operatorname{\textnormal{NC}}^{(s)}(W)\mid V^{w_{1}}\in W\cdot V^{P}\} by the order dd element in /ns\mathbb{Z}/ns, as expected.

5. Type BCnBC_{n}

Let WW be the Weyl group of type BCnBC_{n}. Here we start with the exotic Springer theory first introduced by Kato [Kat09]. Later it was revealed by Achar-Henderson [AH08] that it has a strong connection with the Green functions attached to limit symbols.

5.1. Exotic Springer representations

Let 𝐤\mathbf{k} be an algebraically closed field whose characteristic is not equal to 22. For instance we may set 𝐤=\mathbf{k}=\mathbb{C} or 𝐤=𝔽q¯\mathbf{k}=\overline{\mathbb{F}_{q}} where 2q2\nmid q. Set G=GL2n(𝐤)G=GL_{2n}(\mathbf{k}) and 𝔤=LieG\mathfrak{g}=\operatorname{\textup{Lie}}G. We regard GG as a group of 𝐤\mathbf{k}-linear automorphism of 𝐤2n\mathbf{k}^{2n} and 𝔤\mathfrak{g} as the endomorphism algebra of 𝐤2n\mathbf{k}^{2n}.

We fix a symplectic form ,\langle\ ,\ \rangle on 𝐤2n\mathbf{k}^{2n}. Then there exists an involution θ:GG\theta:G\rightarrow G such that for any gGg\in G and v,w𝐤2nv,w\in\mathbf{k}^{2n} we have g1v,w=v,θ(g)w\langle g^{-1}v,w\rangle=\langle v,\theta(g)w\rangle. This also induces an involution on 𝔤\mathfrak{g}, which we again denote by θ:𝔤𝔤\theta:\mathfrak{g}\rightarrow\mathfrak{g}. We have an eigenspace decomposition 𝔤=𝔤+𝔤\mathfrak{g}=\mathfrak{g}^{+}\oplus\mathfrak{g}^{-} where 𝔤±={X𝔤Xv,w±v,Xw=0 for any v,w𝐤2n}\mathfrak{g}^{\pm}=\{X\in\mathfrak{g}\mid\langle Xv,w\rangle\pm\langle v,Xw\rangle=0\textup{ for any }v,w\in\mathbf{k}^{2n}\}. Note that GθG^{\theta} is isomorphic to the symplectic group Sp2n(𝐤)Sp_{2n}(\mathbf{k}), and thus its Weyl group is identified with WW.

Let 𝒩\mathcal{N}^{-} be the set of nilpotent elements in 𝔤\mathfrak{g}^{-}. The variety 𝒩×𝐤2n\mathcal{N}^{-}\times\mathbf{k}^{2n} is called the exotic nilpotent cone [Kat09, 1.1]. There is a diagonal GθG^{\theta}-action on 𝒩×𝐤2n\mathcal{N}^{-}\times\mathbf{k}^{2n} defined by g(X,v)=(Adg(X),g(v))g\cdot(X,v)=(\operatorname{\textup{Ad}}_{g}(X),g(v)). Then the GθG^{\theta}-orbits in 𝒩×𝐤2n\mathcal{N}^{-}\times\mathbf{k}^{2n} are parametrized by 𝔓n,2\mathfrak{P}_{n,2} as follows. (ref. [AH08], [NRS18]) For a GθG^{\theta}-orbit 𝒪𝒩×𝐤2n\mathcal{O}\subset\mathcal{N}^{-}\times\mathbf{k}^{2n}, choose any (N,v)𝒪(N,v)\in\mathcal{O} and let λ\lambda be the Jordan type of NN as an endomorphism on 𝐤2n\mathbf{k}^{2n}. Also let λ^\hat{\lambda} be the Jordan types of NN on 𝐤2n/𝐤[N]v=𝐤2n/span(Nmvm)\mathbf{k}^{2n}/\mathbf{k}[N]v=\mathbf{k}^{2n}/\textup{span}\left(N^{m}v\mid m\in\mathbb{N}\right). Note that λ\lambda and λ^\hat{\lambda} does not depend on the choice of (N,v)𝒪(N,v)\in\mathcal{O}. Then there exists a unique pair (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2} so that λ=(μ+ν)(μ+ν)\lambda=(\mu+\nu)\cup(\mu+\nu) and λ^=(μ+ν)(μ2+ν)\hat{\lambda}=(\mu+\nu)\cup(\mu_{\geq 2}+\nu). Again, (μ,ν)(\mu,\nu) is independent of the choice of (N,v)𝒪(N,v)\in\mathcal{O} and this gives a bijective correspondence from the set of GθG^{\theta}-orbits in 𝒩×𝐤2n\mathcal{N}^{-}\times\mathbf{k}^{2n} to 𝔓n,2\mathfrak{P}_{n,2}. From now on we write 𝒪μ,ν\mathcal{O}_{\mu,\nu} to be the orbit parametrized by (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2}.

Let \mathcal{B} be the flag variety of GθG^{\theta} defined to be:

={F=[F0F1F2n1F2n=𝐤2n]dimFi=i,Fi,F2ni=0 for all i[0,2n]}.\mathcal{B}=\{F_{\bullet}=[F_{0}\subset F_{1}\subset\cdots\subset F_{2n-1}\subset F_{2n}=\mathbf{k}^{2n}]\mid\dim F_{i}=i,\langle F_{i},F_{2n-i}\rangle=0\textup{ for all }i\in[0,2n]\}.

For (N,v)𝒩×𝐤2n(N,v)\in\mathcal{N}^{-}\times\mathbf{k}^{2n}, we define N,v={FNFiFi for all i[0,2n] and vFn}\mathcal{B}_{N,v}=\{F_{\bullet}\in\mathcal{B}\mid NF_{i}\subset F_{i}\textup{ for all }i\in[0,2n]\textup{ and }v\in F_{n}\}, called the exotic Springer fiber of (N,v)(N,v). Then by [Kat09] (see also [SS13]), there exists an action of WW on Hi(N,v)H^{i}(\mathcal{B}_{N,v}) for each ii\in\mathbb{Z}, called the exotic Springer representation.

Note that Hi(N,v)H^{i}(\mathcal{B}_{N,v}) does not depend on the choice of (N,v)(N,v) as long as it is contained in a fixed GθG^{\theta}-orbit of 𝒩×𝐤2n\mathcal{N}^{-}\times\mathbf{k}^{2n}. Now we define Qμ,ν=iH2i(N,v)qiQ_{\mu,\nu}=\sum_{i\in\mathbb{N}}H^{2i}(\mathcal{B}_{N,v})q^{i} as a graded WW-representation where (N,v)(N,v) is any element in 𝒪μ,ν\mathcal{O}_{\mu,\nu}. Then the leading term of Qμ,νQ_{\mu,\nu} is given by χμ,ν\chi^{\mu,\nu}, i.e. the exotic Springer correspondence is given by 𝒪μ,νχμ,ν\mathcal{O}_{\mu,\nu}\mapsto\chi^{\mu,\nu}. Moreover, Qμ,νQ_{\mu,\nu} coincides with the Green function with respect to χμ,νIrr(W)\chi^{\mu,\nu}\in\operatorname{\textup{Irr}}(W) attached to the limit symbols, which means that Qμ,ν=Qχμ,νQ_{\mu,\nu}=Q_{\chi^{\mu,\nu}}

5.2. Springer representation of LieSp2n(𝔽2¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{2}})

There is another Springer theory, i.e. of LieSp2n\operatorname{\textup{Lie}}Sp_{2n} in characteristic 2, which shares the same Green functions as exotic one. Here we briefly describe its properties.

We start with the parametrization of the nilpotent orbits in LieSp2n(𝔽2¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{2}}). (ref. [Hes79], [Xue12b, Section 2.6]) Let Ω\Omega be the set of pairs (λ,κ)(\lambda,\kappa) where λ𝔓2n\lambda\in\mathfrak{P}_{2n} and κ\kappa is a function from {rλ}\{r\in\lambda\} to \mathbb{N} such that the following conditions hold:

  1. mλ(r)m_{\lambda}(r) is even if rλr\in\lambda is odd,

  2. for any rλr\in\lambda we have 0κ(r)r/20\leq\kappa(r)\leq r/2,

  3. κ(r)=r/2\kappa(r)=r/2 if mλ(r)m_{\lambda}(r) is odd, and

  4. for r,rλr,r^{\prime}\in\lambda such that rrr^{\prime}\leq r, we have κ(r)κ(r)\kappa(r^{\prime})\leq\kappa(r) and rκ(r)rκ(r)r^{\prime}-\kappa(r^{\prime})\leq r-\kappa(r).

For NLieSp2n(𝔽2¯)N\in\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{2}}), we attach (λ,κ)Ω(\lambda,\kappa)\in\Omega to its orbit as follows. Let λ2n\lambda\vdash 2n be the Jordan type of NN. For rλr\in\lambda we define κ\kappa to be κ(r)=min{iN2i+1v,v=0 for any vkerNr}.\kappa(r)=\min\{i\in\mathbb{N}\mid\langle N^{2i+1}v,v\rangle=0\textup{ for any }v\in\ker N^{r}\}. Then (λ,κ)Ω(\lambda,\kappa)\in\Omega and it is independent of the choice of NN in its orbit. This gives a desired parametrization.

We describe the Springer correspondence of LieSp2n(𝔽2¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{2}}). It is summarized by the bijection ι:Ω𝔓n,2\iota:\Omega\rightarrow\mathfrak{P}_{n,2}. (This bijection is deduced from, but not exactly the same as, the one defined in [Xue12a, 8.1]. Also see [Kim19, Section 13.1].) For (λ,κ)Ω(\lambda,\kappa)\in\Omega, choose ss\in\mathbb{N} such that 2sl(λ)2s\geq l(\lambda). We partition [1,2s+1][1,2s+1] into blocks of size 1 or 2 so that:

  1. -

    {i}\{i\} is a single block if and only if κ(λi)=λi/2\kappa(\lambda_{i})=\lambda_{i}/2 and

  2. -

    other blocks consist of two consecutive integers.

(Here we adopt the convention that κ(0)=0\kappa(0)=0.) Note that if {i,i+1}\{i,i+1\} is a block then λi=λi+1\lambda_{i}=\lambda_{i+1} (and thus κ(λi)=κ(λi+1)\kappa(\lambda_{i})=\kappa(\lambda_{i+1})). Now we set cic_{i}\in\mathbb{N} for i[1,2s+1]i\in[1,2s+1] to be:

  1. -

    if {i}\{i\} is a single block then ci=λi/2c_{i}=\lambda_{i}/2, and

  2. -

    otherwise if {i,i+1}\{i,i+1\} is a block then ci=κ(λi)c_{i}=\kappa(\lambda_{i}) and ci+1=λiκ(λi)c_{i+1}=\lambda_{i}-\kappa(\lambda_{i}).

Now we set μ=(c1,c3,,c2s+1)\mu=(c_{1},c_{3},\ldots,c_{2s+1}) and ν=(c2,c4,,c2s)\nu=(c_{2},c_{4},\ldots,c_{2s}) (and remove zeroes at the end if necessary so that μ\mu and ν\nu do not depend on the choice of ss). Then, from the definition of Ω\Omega it follows that (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2}. Now we define ι(λ,κ)=(μ,ν)\iota(\lambda,\kappa)=(\mu,\nu). This is indeed a bijection and describes the Springer correspondence. In particular, if we let Qλ,κQ_{\lambda,\kappa} be the Green function for the nilpotent orbit parametrized by (λ,κ)Ω(\lambda,\kappa)\in\Omega then we have Qλ,κ=Qι(λ,κ)Q_{\lambda,\kappa}=Q_{\iota(\lambda,\kappa)}.

5.3. Analogue of Sommers’ theorem

Here we prove an analogue of [Som11, Theorem 2] in our setting. Let V=χ(n1),(1)Irr(W)V=\chi^{(n-1),(1)}\in\operatorname{\textup{Irr}}(W) be the reflection representation of WW.

Theorem 5.1 (cf. [Som11, Theorem 2]).

For (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2}, we have

i=0nQμ,ν,iVyi=yl(μ,ν)j=1l(μ,ν)(1+yq2j1),\displaystyle\sum_{i=0}^{n}\langle Q_{\mu,\nu},\wedge^{i}V\rangle y^{i}=y^{l(\mu,\nu)}\prod_{j=1}^{l(\mu,\nu)}(1+yq^{2j-1}),
or equivalently i=0nQμ,ν,niVyi=ynl(μ,ν)j=1l(μ,ν)(y+q2j1),\displaystyle\textnormal{or equivalently }\sum_{i=0}^{n}\langle Q_{\mu,\nu},\wedge^{n-i}V\rangle y^{i}=y^{n-l(\mu,\nu)}\prod_{j=1}^{l(\mu,\nu)}(y+q^{2j-1}),

where l(μ,ν)=max{l(μ)1,l(ν)}=l(μ2+ν)l(\mu,\nu)=\max\{l(\mu)-1,l(\nu)\}=l(\mu_{\geq 2}+\nu).

The rest of this section is devoted to the proof of this theorem. Here we mainly follow the argument of [Som11]. Most of the statements therein remain valid with minor modification. We start with the following lemma.

Lemma 5.2.

We have Qμ,ν,V=i=1l(μ,ν)q2i1\langle Q_{\mu,\nu},V\rangle=\sum_{i=1}^{l(\mu,\nu)}q^{2i-1}.

Proof.

Let us define Qλ,κ𝒦(W)[q]Q_{\lambda,\kappa}\in\mathcal{K}(W)[q] as in 5.2. Then it follows from Spaltenstein’s result [Spa91, Proposition 1.7(b)] that Qλ,κ,V=i=1l(λ)/2q2i1\langle Q_{\lambda,\kappa},V\rangle=\sum_{i=1}^{\lfloor l(\lambda)/2\rfloor}q^{2i-1}. (Note that his result is independent of the characteristic of the base field.) On the other hand, it is easy to see that l(λ)/2=l(ι(λ,κ))\lfloor l(\lambda)/2\rfloor=l(\iota(\lambda,\kappa)). Thus the claim follows from the fact that ι\iota is a bijection. ∎

Let us define

gμ,ν=i=0nQμ,ν,niVyi and hμ,ν=|𝒪μ,ν(𝔽q)|gμ,ν,g_{\mu,\nu}=\sum_{i=0}^{n}\langle Q_{\mu,\nu},\wedge^{n-i}V\rangle y^{i}\quad\textnormal{ and }\quad h_{\mu,\nu}=|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|g_{\mu,\nu},

where |𝒪μ,ν(𝔽q)|[q]|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\in\mathbb{Q}[q] is the number of 𝔽q\mathbb{F}_{q}-points in 𝒪μ,ν\mathcal{O}_{\mu,\nu} when the base field is 𝔽q¯\overline{\mathbb{F}_{q}} (and 𝒪μ,ν\mathcal{O}_{\mu,\nu} is split over 𝔽q\mathbb{F}_{q}). Also for χIrr(W)\chi\in\operatorname{\textup{Irr}}(W), we define

τ~(χ)=i,jSi(V)jV,χqiyj=jS(V)jV,χyj\tilde{\tau}(\chi)=\sum_{i,j\in\mathbb{N}}\langle S^{i}(V)\otimes\wedge^{j}V,\chi\rangle q^{i}y^{j}=\sum_{j\in\mathbb{N}}\langle S^{*}(V)\otimes\wedge^{j}V,\chi\rangle y^{j}

where we set S(V)=iSi(V)qiS^{*}(V)=\sum_{i\in\mathbb{N}}S^{i}(V)q^{i}. Now we prove the following lemma.

Lemma 5.3 (cf. [Som11, (9)]).

For χIrr(W)\chi\in\operatorname{\textup{Irr}}(W), we have

(1)nqn2τ~(χ)i=1n(q2i1)=(μ,ν)𝔓n,2hμ,νQμ,ν,χ.(-1)^{n}q^{n^{2}}\tilde{\tau}(\chi)\prod_{i=1}^{n}(q^{2i}-1)=\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2}}h_{\mu,\nu}\langle Q_{\mu,\nu},\chi\rangle.
Proof.

We start with the orthogonality formula of Green functions presented in [Sho82], which is still valid in our situation:

qn2QIdφχ,(1n)=(μ,ν)𝔓n,2|𝒪μ,ν(𝔽q)|Qμ,ν,φQμ,ν.q^{n^{2}}Q_{Id}\otimes\varphi\otimes\chi^{\emptyset,(1^{n})}=\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2}}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\langle Q_{\mu,\nu},\varphi\rangle Q_{\mu,\nu}.

Here we use the fact that the number of positive roots in the root system of WW is n2n^{2} and the centralizer of any (N,v)𝒩×𝐤2n(N,v)\in\mathcal{N}^{-}\times\mathbf{k}^{2n} in GθG^{\theta} is connected. If we set φ=njV\varphi=\wedge^{n-j}V so that φχ,(1n)njVnVjV\varphi\otimes\chi^{\emptyset,(1^{n})}\simeq\wedge^{n-j}V\otimes\wedge^{n}V\simeq\wedge^{j}V, then it follows that

qn2QIdjV=(μ,ν)𝔓n,2|𝒪μ,ν(𝔽q)|Qμ,ν,njVQμ,ν.q^{n^{2}}Q_{Id}\otimes\wedge^{j}V=\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2}}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\langle Q_{\mu,\nu},\wedge^{n-j}V\rangle Q_{\mu,\nu}.

Note that QId=(1)nS(V)i=1n(q2i1)Q_{Id}=(-1)^{n}S^{*}(V)\cdot\prod_{i=1}^{n}(q^{2i}-1) by the formula of Chevalley. Thus we have

((1)nqn2i=1n(q2i1))S(V)jV=(μ,ν)𝔓n,2|𝒪μ,ν(𝔽q)|Qμ,ν,njVQμ,ν.\left((-1)^{n}q^{n^{2}}\prod_{i=1}^{n}(q^{2i}-1)\right)S^{*}(V)\otimes\wedge^{j}V=\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2}}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\langle Q_{\mu,\nu},\wedge^{n-j}V\rangle Q_{\mu,\nu}.

Now the result follows from taking ,χ\langle-,\chi\rangle, multiplying yjy^{j}, and summing up over jj\in\mathbb{N}. ∎

From [GNS99, Proposition 3.3], we have

τ~(χμ,ν)=q2z(μ)+2z(ν)+|ν|xμ1+yq2c(x)+11q2h(x)x′′ν1+yq2c(x′′)11q2h(x′′)\tilde{\tau}(\chi^{\mu,\nu})=q^{2z(\mu)+2z(\nu)+|\nu|}\prod_{x^{\prime}\in\mu}\frac{1+yq^{2c(x^{\prime})+1}}{1-q^{2h(x^{\prime})}}\prod_{x^{\prime\prime}\in\nu}\frac{1+yq^{2c(x^{\prime\prime})-1}}{1-q^{2h(x^{\prime\prime})}}

where z(λ)=i>0(i1)λiz(\lambda)=\sum_{i\in\mathbb{Z}_{>0}}(i-1)\lambda_{i} and c(b)c(b) (resp. h(b)h(b)) is the content(==(column index)-(row index)) (resp. the hook length) of bb in the Young diagram. In particular, by considering boxes at the first column of the Young diagrams of μ\mu and ν\nu, we see that (1+yq2i+1)τ~(χμ,ν)(1+yq^{-2i+1})\mid\tilde{\tau}(\chi^{\mu,\nu}) when i[1,l(μ,ν)]i\in[1,l(\mu,\nu)]. (This statement is vacuous when l(μ,ν)=0l(\mu,\nu)=0, i.e. μ=(n)\mu=(n) and ν=\nu=\emptyset.)

Before the proof of Theorem 5.1 let us observe the triangularity of Springer representations as follows. For (ρ,σ),(μ,ν)𝔓n,2(\rho,\sigma),(\mu,\nu)\in\mathfrak{P}_{n,2}, we say that (ρ,σ)(μ,ν)(\rho,\sigma)\leq(\mu,\nu) if

ρ1+σ1++ρk+σkμ1+ν1++μk+νk, and\displaystyle\rho_{1}+\sigma_{1}+\cdots+\rho_{k}+\sigma_{k}\leq\mu_{1}+\nu_{1}+\cdots+\mu_{k}+\nu_{k},\textnormal{ and}
ρ1+σ1++ρk+σk+ρk+1μ1+ν1++μk+νk+μk+1\displaystyle\rho_{1}+\sigma_{1}+\cdots+\rho_{k}+\sigma_{k}+\rho_{k+1}\leq\mu_{1}+\nu_{1}+\cdots+\mu_{k}+\nu_{k}+\mu_{k+1}

for any kk\in\mathbb{N}. Then the result of [AH08, Theorem 6.3] states that 𝒪ρ,σ𝒪μ,ν¯\mathcal{O}_{\rho,\sigma}\subset\overline{\mathcal{O}_{\mu,\nu}} if and only if (ρ,σ)(μ,ν)(\rho,\sigma)\leq(\mu,\nu). Moreover, by [SS14] and [Kat17], it follows that Qμ,ν,χρ,σ0\langle Q_{\mu,\nu},\chi^{\rho,\sigma}\rangle\neq 0 if and only if (ρ,σ)(μ,ν)(\rho,\sigma)\leq(\mu,\nu). Note that (ρ,σ)(μ,ν)(\rho,\sigma)\leq(\mu,\nu) then l(ρ,σ)l(μ,ν)l(\rho,\sigma)\leq l(\mu,\nu). In particular, if we set χρ,σ=jV\chi^{\rho,\sigma}=\wedge^{j}V, i.e. (ρ,σ)=((nj),(1,1,,1))(\rho,\sigma)=((n-j),(1,1,\ldots,1)), then the above inequalities are translated to njμ1n-j\leq\mu_{1} and nj+kμ1+ν1++μk+νkn-j+k\leq\mu_{1}+\nu_{1}+\cdots+\mu_{k}+\nu_{k} for any 1kj1\leq k\leq j. Thus when k=jk=j we have that l(μ,ν)l(μ),l(ν)jl(\mu,\nu)\leq l(\mu),l(\nu)\leq j.

We are ready to prove Theorem 5.1, i.e. gμ,ν=ynl(μ,ν)j=1l(μ,ν)(y+q2j1)g_{\mu,\nu}=y^{n-l(\mu,\nu)}\prod_{j=1}^{l(\mu,\nu)}(y+q^{2j-1}). To this end, first we prove that the RHS divides hμ,νh_{\mu,\nu} by induction on dim𝒪μ,ν\dim\mathcal{O}_{\mu,\nu}. First, if dim𝒪μ,ν=0\dim\mathcal{O}_{\mu,\nu}=0, i.e. 𝒪μ,ν={0}\mathcal{O}_{\mu,\nu}=\{0\} then it follows from the result of Solomon [Sol63]. In general, by Lemma 5.3 and the triangularity of Springer representations we have

(1)nqn2τ~(χμ,ν)i=1n(q2i1)=hμ,νq2z(μ)+2(ν)+|ν|+(ρ,σ)>(μ,ν)hρ,σQρ,σ,χμ,ν.(-1)^{n}q^{n^{2}}\tilde{\tau}(\chi^{\mu,\nu})\prod_{i=1}^{n}(q^{2i}-1)=h_{\mu,\nu}q^{2z(\mu)+2(\nu)+|\nu|}+\sum_{(\rho,\sigma)>(\mu,\nu)}h_{\rho,\sigma}\langle Q_{\rho,\sigma},\chi^{\mu,\nu}\rangle.

Here we use the fact that degqQμ,ν=2z(μ)+2z(ν)+|ν|=n2dim𝒪μ,ν/2\deg_{q}Q_{\mu,\nu}=2z(\mu)+2z(\nu)+|\nu|=n^{2}-\dim\mathcal{O}_{\mu,\nu}/2 (see [SS13, Remark 2.6]). Now by induction hypothesis and the argument above, it follows that j=1l(μ,ν)(y+q2j1)hμ,ν\prod_{j=1}^{l(\mu,\nu)}(y+q^{2j-1})\mid h_{\mu,\nu}.

Since |𝒪μ,ν(𝔽q)||\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})| does not have any common factor with j=1l(μ,ν)(y+q2j1)\prod_{j=1}^{l(\mu,\nu)}(y+q^{2j-1}), it follows that j=1l(μ,ν)(y+q2j1)|gμ,ν\prod_{j=1}^{l(\mu,\nu)}(y+q^{2j-1})|g_{\mu,\nu}. On the other hand, we previously observed that if nj>l(μ,ν)n-j>l(\mu,\nu) then Qμ,ν,njV=0\langle Q_{\mu,\nu},\wedge^{n-j}V\rangle=0. Thus it follows that ynl(μ,ν)gμ,νy^{n-l(\mu,\nu)}\mid g_{\mu,\nu} as well. Now by considering the degree with respect to yy, it follows that gμ,ν=cynl(μ,ν)j=1l(μ,ν)(y+q2j1)g_{\mu,\nu}=cy^{n-l(\mu,\nu)}\prod_{j=1}^{l(\mu,\nu)}(y+q^{2j-1}) for some c𝐤c\in\mathbf{k}. However, c=1c=1 since the coefficient of yny^{n} in gμ,νg_{\mu,\nu} equals 1 (as the trivial representation only occurs at the zeroth cohomology group of the corresponding Springer fiber). This suffices for the proof of Theorem 5.1.

5.4. qq-Kreweras numbers and positivity

In this section we find a closed formula of Krew(W,χμ,ν,t;q)\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q). To this end, first we let mμ,ν(r)=m(μ+ν)(μ2+ν)(r)/2m_{\mu,\nu}(r)=\lfloor m_{(\mu+\nu)\cup(\mu_{\geq 2}+\nu)}(r)/2\rfloor for (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2} and r>0r\in\mathbb{Z}_{>0}. This can also be interpreted as follows. We set Λ=(μ1,ν1,μ2,ν2,)=(a1,a2,a3,a4,)\Lambda=(\mu_{1},\nu_{1},\mu_{2},\nu_{2},\ldots)=(a_{1},a_{2},a_{3},a_{4},\ldots). Then Λ\Lambda is not necessarily a partition, but it is a quasi-partition in the sense of [AHS11]; we have aiai+2a_{i}\geq a_{i+2} for i>0i\in\mathbb{Z}_{>0}. Also we have μ+ν=(a2i1+a2i)i1\mu+\nu=(a_{2i-1}+a_{2i})_{i\geq 1} and μ2+ν=(a2i+a2i+1)i1\mu_{\geq 2}+\nu=(a_{2i}+a_{2i+1})_{i\geq 1}.

Lemma 5.4.

Set Λ=(a1,a2,)\Lambda=(a_{1},a_{2},\ldots) as above. For r(μ+ν)(μ2+ν)r\in(\mu+\nu)\cup(\mu_{\geq 2}+\nu), suppose that ii (resp. jj) is the smallest (resp. largest) index so that ai+ai+1=ra_{i}+a_{i+1}=r (resp. aj+aj+1=ra_{j}+a_{j+1}=r). Then, there are three possible cases:

  1. (1)

    If ii and jj are both even then mμ+ν(r)=mμ2+ν(r)1m_{\mu+\nu}(r)=m_{\mu_{\geq 2}+\nu}(r)-1. Thus we have mμ,ν(r)=mμ+ν(r)=mμ2+ν(r)1m_{\mu,\nu}(r)=m_{\mu+\nu}(r)=m_{\mu_{\geq 2}+\nu}(r)-1.

  2. (2)

    If ij(mod2)i\not\equiv j\pmod{2} then mμ+ν(r)=mμ2+ν(r)m_{\mu+\nu}(r)=m_{\mu_{\geq 2}+\nu}(r). Thus we have mμ,ν(r)=mμ+ν(r)=mμ2+ν(r)m_{\mu,\nu}(r)=m_{\mu+\nu}(r)=m_{\mu_{\geq 2}+\nu}(r).

  3. (3)

    If ii and jj are both odd then mμ+ν(r)=mμ2+ν(r)+1m_{\mu+\nu}(r)=m_{\mu_{\geq 2}+\nu}(r)+1. Thus we have mμ,ν(r)=mμ+ν(r)1=mμ2+ν(r)m_{\mu,\nu}(r)=m_{\mu+\nu}(r)-1=m_{\mu_{\geq 2}+\nu}(r).

Proof.

It can be shown case-by-case. ∎

Let L(μ,ν)L(\mu,\nu) be the number of r(μ+ν)(μ2+ν)r\in(\mu+\nu)\cup(\mu_{\geq 2}+\nu) such that the first condition in Lemma 5.4 is satisfied. (In this case we necessarily have rμ2+νr\in\mu_{\geq 2}+\nu.) In other words, we have L(μ,ν)+r>0mμ,ν(r)=l(μ,ν)=r>0mμ2+ν(r)L(\mu,\nu)+\sum_{r>0}m_{\mu,\nu}(r)=l(\mu,\nu)=\sum_{r>0}m_{\mu_{\geq 2}+\nu}(r). Also we define

z(μ,ν)=2z(μ)+2z(ν)+|ν|=2z(μ+ν)+|ν|z(\mu,\nu)=2z(\mu)+2z(\nu)+|\nu|=2z(\mu+\nu)+|\nu|

where z(λ)=i>0(i1)λiz(\lambda)=\sum_{i>0}(i-1)\lambda_{i}. We set d(μ,ν)=r>0mμ,ν(r)(mμ,ν(r)+1)d(\mu,\nu)=\sum_{r>0}m_{\mu,\nu}(r)(m_{\mu,\nu}(r)+1).

Theorem 5.5.

For (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2}, the qq-Kreweras number Krew(W,χμ,ν,t;q)\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q) is given by

qt(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)nj=1L(μ,ν)(qt2j+11)(t1)/2L(μ,ν)(t1)/2l(μ,ν),mμ,ν(>0).\displaystyle q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n}\prod_{j=1}^{L(\mu,\nu)}(q^{t-2j+1}-1)\genfrac{\llbracket}{\rrbracket}{0.0pt}{}{(t-1)/2-L(\mu,\nu)}{(t-1)/2-l(\mu,\nu),m_{\mu,\nu}(\mathbb{Z}_{>0})}.

In particular, when tt is odd, Krew(W,χμ,ν,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q)\in\mathbb{N}[q] if and only if L(μ,ν)=0L(\mu,\nu)=0.

The rest of this section is devoted to its proof. Using S(V)=QIdi=1n(1q2i)1S^{*}(V)=Q_{Id}\cdot\prod_{i=1}^{n}(1-q^{2i})^{-1}, we have

t\displaystyle\mathcal{H}_{t} =j(qt)jS(V)jV=i=1n(1q2i)1j(qt)jQIdjV.\displaystyle=\sum_{j\in\mathbb{N}}(-q^{t})^{j}S^{*}(V)\otimes\wedge^{j}V=\prod_{i=1}^{n}(1-q^{2i})^{-1}\sum_{j\in\mathbb{N}}(-q^{t})^{j}Q_{Id}\otimes\wedge^{j}V.

Recall that qn2QIdjV=(μ,ν)𝔓n,2|𝒪μ,ν(𝔽q)|Qμ,ν,njVQμ,νq^{n^{2}}Q_{Id}\otimes\wedge^{j}V=\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2}}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\langle Q_{\mu,\nu},\wedge^{n-j}V\rangle Q_{\mu,\nu}. Thus we have

t\displaystyle\mathcal{H}_{t} =qn2i=1n(1q2i)1j(qt)j(μ,ν)𝔓n,2|𝒪μ,ν(𝔽q)|Qμ,ν,njVQμ,ν.\displaystyle=q^{-n^{2}}\prod_{i=1}^{n}(1-q^{2i})^{-1}\sum_{j\in\mathbb{N}}(-q^{t})^{j}\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2}}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\langle Q_{\mu,\nu},\wedge^{n-j}V\rangle Q_{\mu,\nu}.

It follows that

Krew(W,χμ,ν,t;q)=qn2i=1n(1q2i)1j(qt)j|𝒪μ,ν(𝔽q)|Qμ,ν,njV.\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q)=q^{-n^{2}}\prod_{i=1}^{n}(1-q^{2i})^{-1}\sum_{j\in\mathbb{N}}(-q^{t})^{j}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\langle Q_{\mu,\nu},\wedge^{n-j}V\rangle.

On the other hand, if we substitute yy with qt-q^{t} in the equation of Theorem 5.1 then we have

i=0nQμ,ν,njV(qt)j\displaystyle\sum_{i=0}^{n}\langle Q_{\mu,\nu},\wedge^{n-j}V\rangle(-q^{t})^{j} =(qt)nl(μ,ν)j=1l(μ,ν)(qt+q2j1)\displaystyle=(-q^{t})^{n-l(\mu,\nu)}\prod_{j=1}^{l(\mu,\nu)}(-q^{t}+q^{2j-1})
=(qt)nl(μ,ν)j=1l(μ,ν)(q2j1)j=1l(μ,ν)(qt2j+11)\displaystyle=(-q^{t})^{n-l(\mu,\nu)}\prod_{j=1}^{l(\mu,\nu)}(-q^{2j-1})\prod_{j=1}^{l(\mu,\nu)}(q^{t-2j+1}-1)
=(1)nqt(nl(μ,ν))+l(μ,ν)2j=1l(μ,ν)(qt2j+11).\displaystyle=(-1)^{n}q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}}\prod_{j=1}^{l(\mu,\nu)}(q^{t-2j+1}-1).

Thus we have

Krew(W,χμ,ν,t;q)\displaystyle\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q) =qn2i=1n(1q2i)1|𝒪μ,ν(𝔽q)|(1)nqt(nl(μ,ν))+l(μ,ν)2j=1l(μ,ν)(qt2j+11)\displaystyle=q^{-n^{2}}\prod_{i=1}^{n}(1-q^{2i})^{-1}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|\cdot(-1)^{n}q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}}\prod_{j=1}^{l(\mu,\nu)}(q^{t-2j+1}-1)
=qt(nl(μ,ν))+l(μ,ν)2n2i=1n(q2i1)1j=1l(μ,ν)(qt2j+11)|𝒪μ,ν(𝔽q)|.\displaystyle=q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}-n^{2}}\prod_{i=1}^{n}(q^{2i}-1)^{-1}\prod_{j=1}^{l(\mu,\nu)}(q^{t-2j+1}-1)|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|.

We recall the result of Sun [Sun11, Corollary 3.13] that gives the closed formula of |𝒪μ,ν(𝔽q)||\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|:

|𝒪μ,ν(𝔽q)|=q2n22z(μ,ν)i=1n(1q2i)rJi=1mμ+ν(r)1(1q2i)rJi=1mμ+ν(r)(1q2i)|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|=q^{2n^{2}-2z(\mu,\nu)}\frac{\prod_{i=1}^{n}(1-q^{-2i})}{\prod_{r\in J}\prod_{i=1}^{m_{\mu+\nu}(r)-1}(1-q^{-2i})\prod_{r\not\in J}\prod_{i=1}^{m_{\mu+\nu}(r)}(1-q^{-2i})}

where rJr\in J if and only if rr is in the third case of Lemma 5.4. (See [Sun11, Notation 2.5].) Therefore, by Lemma 5.4 we see that

|𝒪μ,ν(𝔽q)|\displaystyle|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})| =q2n22z(μ,ν)i=1n(1q2i)r>0i=1mμ,ν(r)(1q2i)\displaystyle=q^{2n^{2}-2z(\mu,\nu)}\frac{\prod_{i=1}^{n}(1-q^{-2i})}{\prod_{r>0}\prod_{i=1}^{m_{\mu,\nu}(r)}(1-q^{-2i})}
=qn22z(μ,ν)+d(μ,ν)ni=1n(q2i1)r>0i=1mμ,ν(r)(q2i1).\displaystyle=q^{n^{2}-2z(\mu,\nu)+d(\mu,\nu)-n}\frac{\prod_{i=1}^{n}(q^{2i}-1)}{\prod_{r>0}\prod_{i=1}^{m_{\mu,\nu}(r)}(q^{2i}-1)}.

We substitute |𝒪μ,ν(𝔽q)||\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})| with this formula in the expansion of Krew(W,χμ,ν,t;q)\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q) to see that

Krew(μ,ν;t)(q)=qt(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)nj=1l(μ,ν)(qt2j+11)r>0i=1mμ,ν(r)(q2i1)\displaystyle\operatorname{\textnormal{Krew}}(\mu,\nu;t)(q)=q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n}\frac{\prod_{j=1}^{l(\mu,\nu)}(q^{t-2j+1}-1)}{\prod_{r>0}\prod_{i=1}^{m_{\mu,\nu}(r)}(q^{2i}-1)}
=qt(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)nj=1L(μ,ν)(qt2j+11)j=1l(μ,ν)L(μ,ν)(qt2j2L(μ,ν)+11)r>0i=1mμ,ν(r)(q2i1)\displaystyle=q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n}\prod_{j=1}^{L(\mu,\nu)}(q^{t-2j+1}-1)\frac{\prod_{j=1}^{l(\mu,\nu)-L(\mu,\nu)}(q^{t-2j-2L(\mu,\nu)+1}-1)}{\prod_{r>0}\prod_{i=1}^{m_{\mu,\nu}(r)}(q^{2i}-1)}
=qt(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)nj=1L(μ,ν)(qt2j+11)(t1)/2L(μ,ν)(t1)/2l(μ,ν),mμ,ν(>0),\displaystyle=q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n}\prod_{j=1}^{L(\mu,\nu)}(q^{t-2j+1}-1)\genfrac{\llbracket}{\rrbracket}{0.0pt}{}{(t-1)/2-L(\mu,\nu)}{(t-1)/2-l(\mu,\nu),m_{\mu,\nu}(\mathbb{Z}_{>0})},

which is what we want to prove.

5.5. Relation with qq-Narayana numbers

When WW is of type BCnBC_{n}, due to Lemma 5.2, Theorem 3.5 reads

Nar(W,k,t;q)=l(μ,ν)=kKrew(W,χμ,ν,t;q).\operatorname{\textnormal{Nar}}(W,k,t;q)=\sum_{l(\mu,\nu)=k}\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q).

First, note that we have

Nar(W,k,t;q)=q(nk)(t12k)nk(t1)/2k.\operatorname{\textnormal{Nar}}(W,k,t;q)=q^{(n-k)(t-1-2k)}\genfrac{\llbracket}{\rrbracket}{0.0pt}{}{n}{k}\genfrac{\llbracket}{\rrbracket}{0.0pt}{}{(t-1)/2}{k}.

On the other hand, from 5.4 we have

Krew(W,χμ,ν,t;q)\displaystyle\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q) =qt(nl(μ,ν))+l(μ,ν)2n2i=1n(q2i1)1j=1l(μ,ν)(qt2j+11)|𝒪μ,ν(𝔽q)|.\displaystyle=q^{t(n-l(\mu,\nu))+l(\mu,\nu)^{2}-n^{2}}\prod_{i=1}^{n}(q^{2i}-1)^{-1}\prod_{j=1}^{l(\mu,\nu)}(q^{t-2j+1}-1)|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|.

Comparing the formulas above, it suffices to show that

l(μ,ν)=k|𝒪μ,ν(𝔽q)|=q(nk)(nk1)(q21)nkn!2nk!k!2.\displaystyle\sum_{l(\mu,\nu)=k}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|=q^{(n-k)(n-k-1)}(q^{2}-1)^{n-k}\frac{\llbracket n\rrbracket!^{2}}{\llbracket n-k\rrbracket!\llbracket k\rrbracket!^{2}}.

Note that the RHS is the same as the one given in [RS18, Lemma 5.1]. Indeed, let 𝔓2nC𝔓2n\mathfrak{P}_{2n}^{C}\subset\mathfrak{P}_{2n} be the set of Jordan types of nilpotent elements in LieSp2n(𝔽q¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{q}}) where 2q2\nmid q. For λ𝔓2nC\lambda\in\mathfrak{P}_{2n}^{C}, let |𝒪λSp(𝔽q)||\mathcal{O}^{Sp}_{\lambda}(\mathbb{F}_{q})| be the number of 𝔽q\mathbb{F}_{q}-points in the nilpotent orbit of LieSp2n(𝔽q¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{q}}) whose elements are of Jordan type λ\lambda (when such an orbit is split over 𝔽q\mathbb{F}_{q}). Then [RS18, Lemma 5.1] states that we have

λ𝔓2nC,l(λ)/2=k|𝒪λSp(𝔽q)|=q(nk)(nk1)(q21)nkn!2nk!k!2.\sum_{\lambda\in\mathfrak{P}_{2n}^{C},\lfloor l(\lambda)/2\rfloor=k}|\mathcal{O}^{Sp}_{\lambda}(\mathbb{F}_{q})|=q^{(n-k)(n-k-1)}(q^{2}-1)^{n-k}\frac{\llbracket n\rrbracket!^{2}}{\llbracket n-k\rrbracket!\llbracket k\rrbracket!^{2}}.

(See also [Lus76].) Therefore, it suffices to show that

(μ,ν)𝔓n,2,l(μ,ν)=k|𝒪μ,ν(𝔽q)|=λ𝔓2nC,l(λ)/2=k|𝒪λSp(𝔽q)|.\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2},l(\mu,\nu)=k}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|=\sum_{\lambda\in\mathfrak{P}_{2n}^{C},\lfloor l(\lambda)/2\rfloor=k}|\mathcal{O}^{Sp}_{\lambda}(\mathbb{F}_{q})|.

To this end, following [AHS11], first we define ΦC:𝔓n,2𝔓2n\Phi^{C}:\mathfrak{P}_{n,2}\rightarrow\mathfrak{P}_{2n} as follows. For (μ,ν)𝔓n,2(\mu,\nu)\in\mathfrak{P}_{n,2} such that μ=(μ1,μ2,)\mu=(\mu_{1},\mu_{2},\ldots) and ν=(ν1,ν2,)\nu=(\nu_{1},\nu_{2},\ldots), we consider the sequence (2μ1,2ν1,2μ2,2ν2,)(2\mu_{1},2\nu_{1},2\mu_{2},2\nu_{2},\ldots) and substitute any two consecutive integers s,ts,t such that s<ts<t with (s+t)/2,(s+t)/2(s+t)/2,(s+t)/2, respectively. (These substitutions do not overlap with one another.) Then the result is a partition of 2n2n which is set to be the image of (μ,ν)(\mu,\nu) under ΦC\Phi^{C}. Note that we have l(μ,ν)=l(ΦC(μ,ν))/2l(\mu,\nu)=\lfloor l(\Phi^{C}(\mu,\nu))/2\rfloor.

We define another map C:𝔓n,2𝔓n,2-^{C}:\mathfrak{P}_{n,2}\rightarrow\mathfrak{P}_{n,2} as follows. For (μ,ν)(\mu,\nu) as above, whenever we have μi<νi1\mu_{i}<\nu_{i}-1 we replace μi,νi\mu_{i},\nu_{i} with (μi+νi)/2,(μi+νi)/2\lfloor(\mu_{i}+\nu_{i})/2\rfloor,\lceil(\mu_{i}+\nu_{i})/2\rceil, respectively, and whenever we have νi<μi+1+1\nu_{i}<\mu_{i+1}+1 we replace νi,μi+1\nu_{i},\mu_{i+1} with (νi+μi+1)/2,(νi+μi+1)/2\lfloor(\nu_{i}+\mu_{i+1})/2\rfloor,\lceil(\nu_{i}+\mu_{i+1})/2\rceil, respectively. (These substitutions do not overlap with one another.) Then the result is again an element of 𝔓n,2\mathfrak{P}_{n,2} denoted by (μ,ν)C(\mu,\nu)^{C}. Note that we have l(μ,ν)=l((μ,ν)C)l(\mu,\nu)=l((\mu,\nu)^{C}).

By [AHS11, Theorem 2.23(2)], when ΦC(ρ,σ)𝔓2nC𝔓2n\Phi^{C}(\rho,\sigma)\in\mathfrak{P}_{2n}^{C}\subset\mathfrak{P}_{2n} we have

(μ,ν)𝔓n,2,(μ,ν)C=(ρ,σ)|𝒪μ,ν(𝔽q)|=|𝒪ΦC(ρ,σ)Sp(𝔽q)|.\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2},(\mu,\nu)^{C}=(\rho,\sigma)}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|=|\mathcal{O}^{Sp}_{\Phi^{C}(\rho,\sigma)}(\mathbb{F}_{q})|.

On the other hand, by [AHS11, Proposition 2.4], ΦC\Phi^{C} is a bijection between 𝔓2nC\mathfrak{P}_{2n}^{C} and the image of C-^{C}. Since l(μ,ν)=l(ρ,σ)=l(ΦC(ρ,σ))/2l(\mu,\nu)=l(\rho,\sigma)=\lfloor l(\Phi^{C}(\rho,\sigma))/2\rfloor, it follows that

(μ,ν)𝔓n,2,l(μ,ν)=k|𝒪μ,ν(𝔽q)|=λ𝔓2nC,l(λ)/2=k|𝒪λSp(𝔽q)|\sum_{(\mu,\nu)\in\mathfrak{P}_{n,2},l(\mu,\nu)=k}|\mathcal{O}_{\mu,\nu}(\mathbb{F}_{q})|=\sum_{\lambda\in\mathfrak{P}_{2n}^{C},\lfloor l(\lambda)/2\rfloor=k}|\mathcal{O}^{Sp}_{\lambda}(\mathbb{F}_{q})|

as desired.

5.6. More on the positivity of qq-Kreweras numbers

In Theorem 5.5, for odd tt we have Krew(W,χμ,ν,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q)\in\mathbb{N}[q] if and only if the first condition of Lemma 5.4 is invalid for any r(μ+ν)(μ2ν)r\in(\mu+\nu)\cup(\mu_{\geq 2}\cup\nu). Here we give other equivalent interpretations of this condition and prove Theorem 3.4 for type BCnBC_{n}.

Recall the Springer correspondence of LieSp2n\operatorname{\textup{Lie}}Sp_{2n} in characteristic 2 discussed in 5.2. Following [Kim19, Section 4.2], we define the notion of critical values. We say that (λ,κ)Ω(\lambda,\kappa)\in\Omega is critical at rλr\in\lambda, or rr is a critical value of (λ,κ)(\lambda,\kappa), if

  1. κ(r)0\kappa(r)\neq 0,

  2. for rλ¯r^{\prime}\in\underline{\lambda}, if r<rr^{\prime}<r then κ(r)<κ(r)\kappa(r^{\prime})<\kappa(r), and

  3. for rλ¯r^{\prime}\in\underline{\lambda}, if r>rr^{\prime}>r then rκ(r)<rκ(r)r^{\prime}-\kappa(r^{\prime})<r-\kappa(r).

Using this notion, we may decompose a nilpotent element into a “distinguished part” and an “induced part”. Indeed, we define m~(r)\tilde{m}(r) to be

  1. 0 if rr is critical,

  2. 1 if (rr is critical and) mλ(r)m_{\lambda}(r) is odd, and

  3. 2 if rr is critical and mλ(r)m_{\lambda}(r) is even.

(Indeed, it is easy to show that rr is critical if mλ(r)m_{\lambda}(r) is odd.) Now we set (λ~,κ~)(\tilde{\lambda},\tilde{\kappa}) to be λ~=(1m~(1)2m~(2))\tilde{\lambda}=(1^{\tilde{m}(1)}2^{\tilde{m}(2)}\cdots) and κ~=κ|λ~\tilde{\kappa}=\kappa|_{\tilde{\lambda}}. Also we set (λ,κ)(\lambda^{\prime},\kappa^{\prime}) to be (λ=1mλ(1)m~(1)2mλ(2)m~(2))(\lambda^{\prime}=1^{m_{\lambda}(1)-\tilde{m}(1)}2^{m_{\lambda}(2)-\tilde{m}(2)}\cdots) (so that λ~+λ=λ\tilde{\lambda}+\lambda^{\prime}=\lambda) and κ=0\kappa^{\prime}=0. If we choose N~LieSp|λ~|(𝔽2¯)\tilde{N}\in\operatorname{\textup{Lie}}Sp_{|\tilde{\lambda}|}(\overline{\mathbb{F}_{2}}) and NLieSp|λ|(𝔽2¯)N^{\prime}\in\operatorname{\textup{Lie}}Sp_{|\lambda^{\prime}|}(\overline{\mathbb{F}_{2}}) such that their orbits are parametrized by (λ~,κ~)(\tilde{\lambda},\tilde{\kappa}) and (λ,κ)(\lambda^{\prime},\kappa^{\prime}), respectively, then it is not hard to show that NN~NN\simeq\tilde{N}\oplus N^{\prime}, i.e. there exists a direct sum decomposition 𝔽2¯2n=V~V\overline{\mathbb{F}_{2}}^{2n}=\tilde{V}\oplus V^{\prime} such that N|V~N~N|_{\tilde{V}}\simeq\tilde{N} and N|VNN|_{V^{\prime}}\simeq N^{\prime}. Moreover, one can show that N~\tilde{N} is a distinguished nilpotent element and NN^{\prime} is a regular nilpotent element in a certain Levi subalgebra LieSp2n(𝔽2¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{2}}) all of whose simple factors are of type AA.

Theorem 5.6.

Suppose that ι(λ,κ)=(μ,ν)\iota(\lambda,\kappa)=(\mu,\nu). Then the following conditions are equivalent:

  1. (1)

    Krew(W,χμ,ν,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi^{\mu,\nu},t;q)\in\mathbb{N}[q], i.e. L(μ,ν)=0L(\mu,\nu)=0.

  2. (2)

    There exists at most one critical value of (λ,κ)(\lambda,\kappa), and if rr is such a value then mλ(r)m_{\lambda}(r) is odd.

  3. (3)

    The nilpotent orbit in LieSp2n(𝔽2¯)\operatorname{\textup{Lie}}Sp_{2n}(\overline{\mathbb{F}_{2}}) parametrized by (λ,κ)(\lambda,\kappa) contains a regular nilpotent element in some Levi subalgebra.

  4. (4)

    μ1=μ2==μl(ν)=μl(ν)+1.\mu_{1}=\mu_{2}=\cdots=\mu_{l(\nu)}=\mu_{l(\nu)+1}. (This condition is still satisfied when μ=\mu=\emptyset or ν=\nu=\emptyset.)

Proof.

(2) \Leftrightarrow (3) is easily deduced from the argument right above the theorem. (This is equivalent to saying that N~\tilde{N} is a regular nilpotent element of LieSp|λ~|(𝔽2¯)\operatorname{\textup{Lie}}Sp_{|\tilde{\lambda}|}(\overline{\mathbb{F}_{2}}).) The equivalence of (2) and (4) is easily shown using the definition of ι\iota.

We prove (1) \Rightarrow (4). Let us set Λ=(μ1,ν1,μ2,ν2,)=(a1,a2,a3,a4,)\Lambda=(\mu_{1},\nu_{1},\mu_{2},\nu_{2},\ldots)=(a_{1},a_{2},a_{3},a_{4},\ldots) and for each r(μ+ν)(μ2+ν)r\in(\mu+\nu)\cup(\mu_{\geq 2}+\nu) we attach the interval I(r)=[s(r),e(r)]I(r)=[s(r),e(r)] where s(r)s(r) (resp. e(r)e(r)) is the smallest (resp. largest) index so that as(r)+as(r)+1=ra_{s(r)}+a_{s(r)+1}=r (resp. ae(r)+ae(r)+1=ra_{e(r)}+a_{e(r)+1}=r). Then (1) is equivalent to that there is no rr such that s(r)s(r) and e(r)e(r) are both even. Since these intervals are pairwise disjoint and their union is [1,l(Λ)][1,l(\Lambda)], we have two possible cases:

  1. (1)

    If l(λ)l(\lambda) is even, i.e. l(μ)l(ν)l(\mu)\leq l(\nu), then all the intervals are of the form [[odd, even]].

  2. (2)

    If l(λ)l(\lambda) is odd, i.e. l(μ)>l(ν)l(\mu)>l(\nu), then there exists a unique r(μ+ν)(μ2+ν)r^{\prime}\in(\mu+\nu)\cup(\mu_{\geq 2}+\nu) so that if r>rr>r^{\prime} then I(r)I(r) is [[odd, even]], I(r)I(r^{\prime}) is [[odd, odd]], and if r<rr<r^{\prime} then I(r)I(r) is [[even, odd]].

In the first case, direct calculation shows that a1=a3==al(Λ)+1=0a_{1}=a_{3}=\cdots=a_{l(\Lambda)+1}=0, i.e. μ1=μ2==μl(ν)=μl(ν)+1=0\mu_{1}=\mu_{2}=\cdots=\mu_{l(\nu)}=\mu_{l(\nu)+1}=0. In the second case, similarly we have a1=a3==ae(r)a_{1}=a_{3}=\cdots=a_{e(r^{\prime})} and ae(r)+1==al(Λ)+1=0a_{e(r^{\prime})+1}=\cdots=a_{l(\Lambda)+1}=0, i.e. l(ν)(e(r)1)/2l(\nu)\leq(e(r^{\prime})-1)/2 and μ1=μ2==μ(e(r)+1)/2\mu_{1}=\mu_{2}=\cdots=\mu_{(e(r^{\prime})+1)/2}. This proves (1) \Rightarrow (4). Conversely, if μ1=μ2==μl(ν)=μl(ν)+1\mu_{1}=\mu_{2}=\cdots=\mu_{l(\nu)}=\mu_{l(\nu)+1} then it is easy to deduce (1) using the definition of ι\iota. It suffices for the proof. ∎

Suppose that (λ,κ)Ω(\lambda,\kappa)\in\Omega satisfies the second condition of Theorem 5.6. Let 2b2b be the unique critical value of (λ,κ)(\lambda,\kappa). (If there is no critical value, then we set 2b=02b=0.) From the definition of a critical value, for any rλr\in\lambda we have κ(r)=b\kappa(r)=b if r2br\geq 2b, κ(r)=rb\kappa(r)=r-b if br<2bb\leq r<2b, and κ(r)=0\kappa(r)=0 if r<br<b. Thus, if we let λ=(a1,a1,a2,a2,,am,am,b,am+1,am+1,,ak,ak)\lambda=(a_{1},a_{1},a_{2},a_{2},\ldots,a_{m},a_{m},b,a_{m+1},a_{m+1},\ldots,a_{k},a_{k}) where a1am>bam+1aka_{1}\geq\cdots\geq a_{m}>b\geq a_{m+1}\geq\cdots\geq a_{k} then by direct calculation we have

ι(λ,κ)=((b,b,,b,am+1,,ak),(a1b,a2b,,amb))\iota(\lambda,\kappa)=((b,b,\ldots,b,a_{m+1},\ldots,a_{k}),(a_{1}-b,a_{2}-b,\ldots,a_{m}-b))

where bb is repeated m+1m+1 times in the first factor. Note that this is equivalent to the description of Φ\Phi in Section 3.3 for type BCnBC_{n}, thus it proves Theorem 3.4 in type BCnBC_{n}.

5.7. Cyclic sieving

We prove Theorem 3.6 in our setting. Here we set t=2ns+1t=2ns+1 for some ss\in\mathbb{N} (note that the Coxeter number of WW is 2n2n). As in [RS18, 6.1], for dd\in\mathbb{N} we set d=d/gcd(d,2)d^{-}=d/\gcd(d,2) and d+=2dd^{+}=2d^{-}. Then for any NN\in\mathbb{N} we have d2Nd+2NdN.d\mid 2N\Leftrightarrow d^{+}\mid 2N\Leftrightarrow d^{-}\mid N.

Recall that each parabolic subgroup PWP\subset W may be identified with Hb×𝔖a1××𝔖akH_{b}\times\operatorname{\mathfrak{S}}_{a_{1}}\times\cdots\times\operatorname{\mathfrak{S}}_{a_{k}} where HbH_{b} is a Coxeter group of type BCbBC_{b} and b+a1++ak=nb+a_{1}+\cdots+a_{k}=n. (Here bb can be zero.) Without loss of generality we may assume that a1aka_{1}\geq\cdots\geq a_{k}. Here we set (λ,κ)Ω(\lambda,\kappa)\in\Omega as described at the end of 5.6 and let (μ,ν)=ι(λ,κ)(\mu,\nu)=\iota(\lambda,\kappa) so that Φ(P)=χ(μ,ν)\Phi(P)=\chi^{(\mu,\nu)}.

If we regard WW as the Weyl group of Sp2n()Sp_{2n}(\mathbb{C}) then PP corresponds to the nilpotent orbit of LieSp2n()\operatorname{\textup{Lie}}Sp_{2n}(\mathbb{C}) whose elements are of Jordan type λ\lambda. Therefore, using the result of [RS18, Section 6], here we only need to show that:

  1. (1)

    If d2nsd\mid 2ns but d2mλ(r)/2d\nmid 2\lfloor m_{\lambda}(r)/2\rfloor for some rr, then Krew(W,Φ(P),2ns+1;ωd)=0\operatorname{\textnormal{Krew}}(W,\Phi(P),2ns+1;\omega_{d})=0.

  2. (2)

    If d2nsd\mid 2ns and d2mλ(r)/2d\mid 2\lfloor m_{\lambda}(r)/2\rfloor for every rr, then

    Krew(W,Φ(P),2ns+1;ωd)=(ns/dns/dl(λ)/2/d,mλ(>0)/2/d).\operatorname{\textnormal{Krew}}(W,\Phi(P),2ns+1;\omega_{d})=\binom{ns/d^{-}}{ns/d^{-}-\lfloor l(\lambda)/2\rfloor/d^{-},\lfloor m_{\lambda}(\mathbb{Z}_{>0})/2\rfloor/d^{-}}.

    (Here we use the fact that mλ(r)m_{\lambda}(r) is odd for at most one rλr\in\lambda, so r>0mλ(r)/2=l(λ)/2\sum_{r\in\mathbb{Z}_{>0}}\lfloor m_{\lambda}(r)/2\rfloor=\lfloor l(\lambda)/2\rfloor.)

Let us embark on the proof. From Theorem 5.5, Krew(W,Φ(P),2ns+1;ωd)\operatorname{\textnormal{Krew}}(W,\Phi(P),2ns+1;\omega_{d}) is equal to

ωd(2ns+1)(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)n[nsnsl(μ,ν),mμ,ν(>0)]ωd2.\omega_{d}^{(2ns+1)(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n}\genfrac{[}{]}{0.0pt}{}{ns}{ns-l(\mu,\nu),m_{\mu,\nu}(\mathbb{Z}_{>0})}_{\omega_{d}^{2}}.

Here, Φ(P)=(μ,ν)=((b,b,,b,am+1,,ak),(a1b,a2b,,amb))\Phi(P)=(\mu,\nu)=((b,b,\ldots,b,a_{m+1},\ldots,a_{k}),(a_{1}-b,a_{2}-b,\ldots,a_{m}-b)) where m[0,k]m\in[0,k] is the largest index so that am>ba_{m}>b and bb is repeated m+1m+1 times in μ\mu. Also, note that in this case L(μ,ν)=0L(\mu,\nu)=0. Now we observe the following lemma.

Lemma 5.7.

Let λ,μ,ν\lambda,\mu,\nu be as above. Then we have mμ,ν(b)=(mλ(b)1)/2m_{\mu,\nu}(b)=(m_{\lambda}(b)-1)/2 and mμ,ν(r)=mλ(r)/2m_{\mu,\nu}(r)=m_{\lambda}(r)/2 if rbr\neq b. In any case, we have mμ,ν(r)=mλ(r)/2m_{\mu,\nu}(r)=\lfloor m_{\lambda}(r)/2\rfloor. Also, it follows that l(μ,ν)=r1mμ,ν(r)=l(λ)/2l(\mu,\nu)=\sum_{r\geq 1}m_{\mu,\nu}(r)=\lfloor l(\lambda)/2\rfloor.

Proof.

It is straightforward from the description of (μ,ν)(\mu,\nu). ∎

Remark.

In general if ι(λ,χ)=(μ,ν)\iota(\lambda,\chi)=(\mu,\nu) (see 5.2 for the definition of ι\iota) then we have:

  1. if rr is not a critical value, then mμ,ν(r)=mλ(r)/2m_{\mu,\nu}(r)=m_{\lambda}(r)/2.

  2. if mλ(r)m_{\lambda}(r) is odd, then (rr is critical and) mμ,ν(r)=(mλ(r)1)/2m_{\mu,\nu}(r)=(m_{\lambda}(r)-1)/2.

  3. if mλ(r)m_{\lambda}(r) is even and rr is a critical value, then mμ,ν(r)=(mλ(r)2)/2m_{\mu,\nu}(r)=(m_{\lambda}(r)-2)/2.

We also recall a useful lemma when calculating the specialization of qq-multinomials.

Lemma 5.8 ([RS18, Lemma 6.3]).

Suppose that d2Nd\mid 2N. Then for a sequence of nonnegative integers a1,,ala_{1},\ldots,a_{l} so that N=a1++alN=a_{1}+\cdots+a_{l}, we have

[Na1,,al]ωd2={(N/da1/d,,al/d) if dai for each i,0 otherwise.\genfrac{[}{]}{0.0pt}{}{N}{a_{1},\ldots,a_{l}}_{\omega_{d}^{2}}=\left\{\begin{aligned} &\binom{N/d^{-}}{a_{1}/d^{-},\ldots,a_{l}/d^{-}}&\textnormal{ if }d^{-}\mid a_{i}\textnormal{ for each }i,\\ &0&\textnormal{ otherwise}.\end{aligned}\right.

Now suppose that d2nsd\mid 2ns but d2mλ(r)/2=2mμ,ν(r)d\nmid 2\lfloor m_{\lambda}(r)/2\rfloor=2m_{\mu,\nu}(r) for some rr. Then it is clear that

[nsnsl(μ,ν),mμ,ν(>0)]ωd2=0,\genfrac{[}{]}{0.0pt}{}{ns}{ns-l(\mu,\nu),m_{\mu,\nu}(\mathbb{Z}_{>0})}_{\omega_{d}^{2}}=0,

thus Krew(W,Φ(P),2ns+1;ωd)=0\operatorname{\textnormal{Krew}}(W,\Phi(P),2ns+1;\omega_{d})=0 as required. From now on we assume that d2nsd\mid 2ns and d2mλ(r)/2=2mμ,ν(r)d\mid 2\lfloor m_{\lambda}(r)/2\rfloor=2m_{\mu,\nu}(r) for every r>0r\in\mathbb{Z}_{>0}. (It also follows that d2l(λ)/2=2l(μ,ν)d\mid 2\lfloor l(\lambda)/2\rfloor=2l(\mu,\nu).) Then by Lemma 5.8, Krew(W,Φ(P),2ns+1;ωd)\operatorname{\textnormal{Krew}}(W,\Phi(P),2ns+1;\omega_{d}) equals

ωd(2ns+1)(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)n(ns/dns/dl(λ)/2/d,mλ(>0)/2/d).\omega_{d}^{(2ns+1)(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n}\binom{ns/d^{-}}{ns/d^{-}-\lfloor l(\lambda)/2\rfloor/d^{-},\lfloor m_{\lambda}(\mathbb{Z}_{>0})/2\rfloor/d^{-}}.

Thus it remains to show that the power of ωd\omega_{d} in the above expression is 1.

In our case, it is easy to see that each of μiT1\mu^{T}_{i}-1 and νiT\nu^{T}_{i} are the sum of mμ,ν(r)m_{\mu,\nu}(r) for some rr, and thus z(μ,ν)=2z(μ)+2z(ν)+|ν|z(\mu,\nu)=2z(\mu)+2z(\nu)+|\nu| is easily seen to be a multiple of dd^{-}. (Recall that 2z(μ)=i1μiT(μiT1)2z(\mu)=\sum_{i\geq 1}\mu_{i}^{T}(\mu_{i}^{T}-1) and 2z(ν)=i1νiT(νiT1)2z(\nu)=\sum_{i\geq 1}\nu_{i}^{T}(\nu_{i}^{T}-1).) Thus we have

(2ns+1)(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)n0(modd),\displaystyle(2ns+1)(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n\equiv 0\pmod{d^{-}},

since dd(μ,ν)=r>0mμ,ν(r)(mμ,ν(r)+1)d^{-}\mid d(\mu,\nu)=\sum_{r>0}m_{\mu,\nu}(r)(m_{\mu,\nu}(r)+1).

Therefore, if dd is odd then d=dd=d^{-} and it suffices for the proof. It remains to assume that dd is even and d=2dd=2d^{-}. Here we have

(2ns+1)(nl(μ,ν))+l(μ,ν)22z(μ,ν)+d(μ,ν)n\displaystyle(2ns+1)(n-l(\mu,\nu))+l(\mu,\nu)^{2}-2z(\mu,\nu)+d(\mu,\nu)-n
l(μ,ν)2l(μ,ν)+d(μ,ν)\displaystyle\equiv l(\mu,\nu)^{2}-l(\mu,\nu)+d(\mu,\nu)
(r1mμ,ν(r))2+r1mμ,ν(r)2\displaystyle\equiv\left(\sum_{r\geq 1}m_{\mu,\nu}(r)\right)^{2}+\sum_{r\geq 1}m_{\mu,\nu}(r)^{2}
2r1mμ,ν(r)2+2r>s1mμ,ν(r)mμ,ν(s)0(modd).\displaystyle\equiv 2\sum_{r\geq 1}m_{\mu,\nu}(r)^{2}+2\sum_{r>s\geq 1}m_{\mu,\nu}(r)m_{\mu,\nu}(s)\equiv 0\pmod{d}.

Thus the claim also holds when dd is even.

6. Type H3H_{3}

In this section we assume that WW is of type H3H_{3}.

6.1. qq-Kreweras numbers and positivity.

Here we list Krew(W,χ,t;q)\operatorname{\textnormal{Krew}}(W,\chi,t;q) for χIrr(W)\chi\in\operatorname{\textup{Irr}}(W). (We acknowledge that these numbers were also calculated by Eric Sommers.)

Irr(W)\operatorname{\textup{Irr}}(W) qq-Kreweras Qχ,V\langle Q_{\chi},V\rangle parabolicparabolic
ϕ1,15\phi_{1,15} [t1][t5][t9]/([2][6][10])[t-1][t-5][t-9]/([2][6][10]) q9+q5+qq^{9}+q^{5}+q trivtriv
ϕ1,0\phi_{1,0} q3t3q^{3t-3} 0 H3H_{3}
ϕ5,5\phi_{5,5} q2t10[t1]/[2]q^{2t-10}[t-1]/[2] qq A1×A1A_{1}\times A_{1}
ϕ5,2\phi_{5,2} q2t6(q1)[t1]q^{2t-6}(q-1)[t-1] qq
ϕ3,6\phi_{3,6} qt7(q1)[t1][t5]/[2]q^{t-7}(q-1)[t-1][t-5]/[2] q5+qq^{5}+q
ϕ3,8\phi_{3,8} qt9[t1][t5]/[2]2q^{t-9}[t-1][t-5]/[2]^{2} q5+qq^{5}+q A1A_{1}
ϕ3,1\phi_{3,1} q2t4(q1)[t1]q^{2t-4}(q-1)[t-1] qq
ϕ3,3\phi_{3,3} q2t6[t1]/[2]q^{2t-6}[t-1]/[2] qq I2(5)I_{2}(5)
ϕ4,3\phi_{4,3} q2t8(q1)[t1]q^{2t-8}(q-1)[t-1] qq
ϕ4,4\phi_{4,4} q2t8[t1]/[2]q^{2t-8}[t-1]/[2] qq A2A_{2}

The first column represents the irreducible representations of WW. Here we use Carter’s notation [Car93, 13.2]. The second column represents corresponding qq-Kreweras numbers. The third column represents Qχ,V\langle Q_{\chi},V\rangle where VV is the reflection representation of WW. The fourth column represents the type of each parabolic subgroup PWP\subset W where Φ(P)=χ\Phi(P)=\chi. Here we do not distinguish different parabolic subgroups of type A1A_{1} as they all correspond to the same representation ϕ3,8\phi_{3,8}.

It is clear from the formula that if t1,5,9(mod10)t\equiv 1,5,9\pmod{10} we have Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{N}[q] if and only if χimΦ\chi\in\operatorname{\textup{im}}\Phi.

6.2. Relation with qq-Narayana numbers

We have

Nar(W,k,t;q)=(qt4k1)3k[3k]q4(qt1;q4)k(q2;q4)k,\operatorname{\textnormal{Nar}}(W,k,t;q)=(q^{t-4k-1})^{3-k}\genfrac{[}{]}{0.0pt}{}{3}{k}_{q^{4}}\frac{(q^{t-1};q^{-4})_{k}}{(q^{2};q^{4})_{k}},

or more precisely

Nar(W,0,t;q)\displaystyle\operatorname{\textnormal{Nar}}(W,0,t;q) =q3(t1),\displaystyle=q^{3(t-1)},
Nar(W,1,t;q)\displaystyle\operatorname{\textnormal{Nar}}(W,1,t;q) =q2(t5)(q8+q4+1)[t1][2],\displaystyle=q^{2(t-5)}(q^{8}+q^{4}+1)\frac{[t-1]}{[2]},
Nar(W,2,t;q)\displaystyle\operatorname{\textnormal{Nar}}(W,2,t;q) =qt9(q8+q4+1)[t1][t5][2][6],\displaystyle=q^{t-9}(q^{8}+q^{4}+1)\frac{[t-1][t-5]}{[2][6]},
Nar(W,3,t;q)\displaystyle\operatorname{\textnormal{Nar}}(W,3,t;q) =[t1][t5][t9][2][6][10].\displaystyle=\frac{[t-1][t-5][t-9]}{[2][6][10]}.

Thus Theorem 3.5 holds in this case by direct calculation.

6.3. Cyclic sieving

We identify WW with s1,s2,s3s12,s22,s33,(s1s2)5,(s2s3)3,(s1s3)2\langle s_{1},s_{2},s_{3}\mid s_{1}^{2},s_{2}^{2},s_{3}^{3},(s_{1}s_{2})^{5},(s_{2}s_{3})^{3},(s_{1}s_{3})^{2}\rangle and fix a (standard) Coxeter element c=s1s2s3c=s_{1}s_{2}s_{3}. Let us label all the reflections in WW as

r1=s1,\displaystyle r_{1}=s_{1}, r2=s2,\displaystyle r_{2}=s_{2}, r3=s1232321,\displaystyle r_{3}=s_{1232321}, r4=s232,\displaystyle r_{4}=s_{232}, r5=s31231,\displaystyle r_{5}=s_{31231},
r6=s121,\displaystyle r_{6}=s_{121}, r7=s123123121,\displaystyle r_{7}=s_{123123121}, r8=s2312312312312,\displaystyle r_{8}=s_{2312312312312}, r9=s323123123,\displaystyle r_{9}=s_{323123123}, r10=s323,\displaystyle r_{10}=s_{323},
r11=s3,\displaystyle r_{11}=s_{3}, r12=s12321,\displaystyle r_{12}=s_{12321}, r13=s2312312,\displaystyle r_{13}=s_{2312312}, r14=s31231231231,\displaystyle r_{14}=s_{31231231231}, r15=s23232,\displaystyle r_{15}=s_{23232},

and all the products of two reflections in WW as

o1=s12,\displaystyle o_{1}=s_{12}, o2=s12312321,\displaystyle o_{2}=s_{12312321}, o3=s1232312312,\displaystyle o_{3}=s_{1232312312}, o4=s23231231,\displaystyle o_{4}=s_{23231231}, o5=s3123,\displaystyle o_{5}=s_{3123},
o6=s23,\displaystyle o_{6}=s_{23}, o7=s1231,\displaystyle o_{7}=s_{1231}, o8=s1232,\displaystyle o_{8}=s_{1232}, o9=s231231,\displaystyle o_{9}=s_{231231}, o10=s123232,\displaystyle o_{10}=s_{123232},
o11=s31,\displaystyle o_{11}=s_{31}, o12=s123121,\displaystyle o_{12}=s_{123121}, o13=s123123123121,\displaystyle o_{13}=s_{123123123121}, o14=s323123123123,\displaystyle o_{14}=s_{323123123123}, o15=s323123.\displaystyle o_{15}=s_{323123}.

(Here, sabc=sasbscs_{abc\cdots}=s_{a}s_{b}s_{c}\cdots.) Then we have NC(1)(W)={id,c}{rir[1,15]}{oii[1,15]}\operatorname{\textnormal{NC}}^{(1)}(W)=\{id,c\}\cup\{r_{i}\mid r\in[1,15]\}\cup\{o_{i}\mid i\in[1,15]\}. Moreover, direct calculation shows that cric1=ri+1cr_{i}c^{-1}=r_{i+1} and coic1=oi+1co_{i}c^{-1}=o_{i+1} for i{5,10,15}i\not\in\{5,10,15\}, and cric1=ri4cr_{i}c^{-1}=r_{i-4} and coic1=oi4co_{i}c^{-1}=o_{i-4} for i{5,10,15}i\in\{5,10,15\}.

From now on we calculate the fixed points of NC(s)(W)\operatorname{\textnormal{NC}}^{(s)}(W) under the /10s\mathbb{Z}/10s-action defined in 3.5 case-by-case. To this end, recall the definition of δ\delta-sequence

(δ1,,δs1,δs)=(w11w2,,ws11ws,ws1c)(\delta_{1},\ldots,\delta_{s-1},\delta_{s})=(w_{1}^{-1}w_{2},\ldots,w_{s-1}^{-1}w_{s},w_{s}^{-1}c)

for each (w1,w2,,ws)NC(s)(w_{1},w_{2},\ldots,w_{s})\in\operatorname{\textnormal{NC}}^{(s)}. It is clear that δiNC(1)(W)\delta_{i}\in\operatorname{\textnormal{NC}}^{(1)}(W) for each ii. In terms of δ\delta-sequence the cyclic action is given by

(δ1,,δs1,δs)(cδsc1,δ1,,δs1).(\delta_{1},\ldots,\delta_{s-1},\delta_{s})\mapsto(c\delta_{s}c^{-1},\delta_{1},\ldots,\delta_{s-1}).

From now on let us write φ(w1,s,d)\varphi(w_{1},s,d) to denote the number of fixed points in NC(s)(W)\operatorname{\textnormal{NC}}^{(s)}(W) starting with w1Ww_{1}\in W by an order dd element (where d10sd\mid 10s). Also let Δ\Delta be the sequence obtained from (δ1,,δs1,δs)(\delta_{1},\ldots,\delta_{s-1},\delta_{s}) by removing the identity elements.

  1. (1)

    w1=cw_{1}=c. In this case Δ=\Delta=\emptyset. Thus φ(c,s,d)=1\varphi(c,s,d)=1.

  2. (2)

    w1=oiw_{1}=o_{i} for some i[1,15]i\in[1,15]. In this case Δ=(oi1c)\Delta=(o_{i}^{-1}c) where oi1co_{i}^{-1}c is a reflection. Thus in this case φ(oi,s,d)=s\varphi(o_{i},s,d)=s if and only if d2d\mid 2 and otherwise φ(oi,s,d)=0\varphi(o_{i},s,d)=0.

  3. (3)

    w1=riw_{1}=r_{i}, i[1,5]i\in[1,5]. By conjugation by cc it suffices to consider the case when ri=s1r_{i}=s_{1}. Then possible Δ\Delta are:

    (r2,r11),(r4,r2),(r15,r4),(r10,r15),(r11,r10),(o6).\displaystyle(r_{2},r_{11}),\quad(r_{4},r_{2}),\quad(r_{15},r_{4}),\quad(r_{10},r_{15}),\quad(r_{11},r_{10}),\quad(o_{6}).

    Thus direct calculation shows that φ(ri,s,d)=5s(s1)/2+s\varphi(r_{i},s,d)=5s(s-1)/2+s if and only if d2d\mid 2 and otherwise φ(ri,s,d)=0\varphi(r_{i},s,d)=0.

  4. (4)

    w1=riw_{1}=r_{i}, i[6,10]i\in[6,10]. By conjugation by cc it suffices to consider the case that ri=s121r_{i}=s_{121}. Then possible Δ\Delta are:

    (r1,r11),(r11,r1),(o11).\displaystyle(r_{1},r_{11}),\quad(r_{11},r_{1}),\quad(o_{11}).

    Thus direct calculation shows that φ(ri,s,d)=s2\varphi(r_{i},s,d)=s^{2} if and only if d2d\mid 2 and otherwise φ(ri,s,d)=0\varphi(r_{i},s,d)=0.

  5. (5)

    w1=riw_{1}=r_{i}, i[11,15]i\in[11,15]. By conjugation by cc it suffices to consider the case that ri=s3r_{i}=s_{3}. Then possible Δ\Delta are:

    (r1,r10),(r5,r1),(r10,r5),(o5).\displaystyle(r_{1},r_{10}),\quad(r_{5},r_{1}),\quad(r_{10},r_{5}),\quad(o_{5}).

    Thus direct calculation shows that φ(ri,s,d)=3s(s1)/2+s\varphi(r_{i},s,d)=3s(s-1)/2+s if and only if d2d\mid 2 and otherwise φ(ri,s,d)=0\varphi(r_{i},s,d)=0.

  6. (6)

    w1=idw_{1}=id. In this case possible Δ\Delta can be read from above calculations. Let us denote by φΔ(id,s,d)\varphi_{\Delta}(id,s,d) the number of fixed elements with fixed Δ\Delta.

    1. (a)

      Δ=(c)\Delta=(c). Direct calculation shows that φΔ(id,s,d)=s\varphi_{\Delta}(id,s,d)=s if and only if d10d\mid 10 and otherwise φΔ(id,s,d)=0\varphi_{\Delta}(id,s,d)=0.

    2. (b)

      Δ=(oi,oi1s)\Delta=(o_{i},o_{i}^{-1}s) or Δ=(oi1s,oi)\Delta=(o_{i}^{-1}s,o_{i}) for some i[1,15]i\in[1,15]. It is clear that φΔ(id,s,d)=s(s1)/2\varphi_{\Delta}(id,s,d)=s(s-1)/2 if and only if d2d\mid 2 and otherwise φΔ(id,s,d)=0\varphi_{\Delta}(id,s,d)=0.

    3. (c)

      Δ=(ri,rj,rk)\Delta=(r_{i},r_{j},r_{k}) for some i,j,k[1,15]i,j,k\in[1,15]. Direct calculation shows that there are 50 possible Δ\Delta, and we have φΔ(id,s,d)=s(s1)(s2)/6\varphi_{\Delta}(id,s,d)=s(s-1)(s-2)/6 if d2d\mid 2 and φΔ(id,s,d)=0\varphi_{\Delta}(id,s,d)=0 if d1,2,3,6d\neq 1,2,3,6. Moreover, if 3s3\mid s and Δ\Delta is one of

      (r1,r4,r2),(r2,r5,r3),(r3,r1,r4),(r4,r2,r5),(r5,r3,r1)(r_{1},r_{4},r_{2}),(r_{2},r_{5},r_{3}),(r_{3},r_{1},r_{4}),(r_{4},r_{2},r_{5}),(r_{5},r_{3},r_{1})

      then the δ\delta-sequence (δ1,,δs)(\delta_{1},\ldots,\delta_{s}) where Δ=(δi,δs/3+i,δ2s/3+i)\Delta=(\delta_{i},\delta_{s/3+i},\delta_{2s/3+i}) for some i[1,s/3]i\in[1,s/3] is fixed by an element of order 3 and 6. Thus we have φΔ(id,s,6)=φΔ(id,s,3)=s/3\varphi_{\Delta}(id,s,6)=\varphi_{\Delta}(id,s,3)=s/3. Otherwise we have φΔ(id,s,3)=0\varphi_{\Delta}(id,s,3)=0.

    In sum, we have:

    φ(id,s,d)={s(5s2)(5s4)/3 if d{1,2},5s/3 if 3s and d{3,6},s if d{5,10},0 otherwise.\varphi(id,s,d)=\left\{\begin{aligned} &s(5s-2)(5s-4)/3&\textnormal{ if }d\in\{1,2\},\\ &5s/3&\textnormal{ if }3\mid s\textnormal{ and }d\in\{3,6\},\\ &s&\textnormal{ if }d\in\{5,10\},\\ &0&\textnormal{ otherwise.}\end{aligned}\right.

Now we consider the qq-Kreweras numbers Krew(W,Φ(P),10s+1;ωd)\operatorname{\textnormal{Krew}}(W,\Phi(P),10s+1;\omega_{d}) where d10sd\mid 10s. Direct calculation shows that:

  1. PP is of type H3H_{3}. Then Φ(P)=ϕ1,0\Phi(P)=\phi_{1,0}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=cw=c. We have Krew(W,ϕ1,0,10s+1;q)=q30s\operatorname{\textnormal{Krew}}(W,\phi_{1,0},10s+1;q)=q^{30s}, thus Krew(W,ϕ1,0,10s+1;ωd)=1\operatorname{\textnormal{Krew}}(W,\phi_{1,0},10s+1;\omega_{d})=1 whenever d10sd\mid 10s. This coincides with φ(c,s,d)=1\varphi(c,s,d)=1.

  2. PP is of type I2(5)I_{2}(5). Then Φ(P)=ϕ3,3\Phi(P)=\phi_{3,3}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=oiw=o_{i} for i[1,5]i\in[1,5]. We have Krew(W,ϕ3,3,10s+1;q)=q20s4[10s]/[2]\operatorname{\textnormal{Krew}}(W,\phi_{3,3},10s+1;q)=q^{20s-4}[10s]/[2], thus Krew(W,ϕ3,3,10s+1;ωd)\operatorname{\textnormal{Krew}}(W,\phi_{3,3},10s+1;\omega_{d}) equals 5s5s if d2d\mid 2 and 0 otherwise. Thus coincides with i=15φ(oi,s,d)\sum_{i=1}^{5}\varphi(o_{i},s,d).

  3. PP is of type A2A_{2}. Then Φ(P)=ϕ4,4\Phi(P)=\phi_{4,4}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=oiw=o_{i} for i[6,10]i\in[6,10]. We have Krew(W,ϕ4,4,10s+1;q)=q20s6[10s]/[2]\operatorname{\textnormal{Krew}}(W,\phi_{4,4},10s+1;q)=q^{20s-6}[10s]/[2], thus Krew(W,ϕ4,4,10s+1;ωd)\operatorname{\textnormal{Krew}}(W,\phi_{4,4},10s+1;\omega_{d}) equals 5s5s if d2d\mid 2 and 0 otherwise. Thus coincides with i=610φ(oi,s,d)\sum_{i=6}^{10}\varphi(o_{i},s,d).

  4. PP is of type A1×A1A_{1}\times A_{1}. Then Φ(P)=ϕ5,5\Phi(P)=\phi_{5,5}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=oiw=o_{i} for i[11,15]i\in[11,15]. We have Krew(W,ϕ5,5,10s+1;q)=q20s8[10s]/[2]\operatorname{\textnormal{Krew}}(W,\phi_{5,5},10s+1;q)=q^{20s-8}[10s]/[2], thus Krew(W,ϕ5,5,10s+1;ωd)\operatorname{\textnormal{Krew}}(W,\phi_{5,5},10s+1;\omega_{d}) equals 5s5s if d2d\mid 2 and 0 otherwise. Thus coincides with i=1115φ(oi,s,d)\sum_{i=11}^{15}\varphi(o_{i},s,d).

  5. PP is of type A1A_{1}. Then Φ(P)=ϕ3,8\Phi(P)=\phi_{3,8}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=riw=r_{i} for i[1,15]i\in[1,15]. We have Krew(W,ϕ3,8,10s+1;q)=q10s8[10s][10s4]/[2]2\operatorname{\textnormal{Krew}}(W,\phi_{3,8},10s+1;q)=q^{10s-8}[10s][10s-4]/[2]^{2}, thus Krew(W,ϕ3,8,10s+1;ωd)\operatorname{\textnormal{Krew}}(W,\phi_{3,8},10s+1;\omega_{d}) equals 5s(5s2)5s(5s-2) if d2d\mid 2 and 0 otherwise. Thus coincides with i=115φ(ri,s,d)\sum_{i=1}^{15}\varphi(r_{i},s,d).

  6. P={id}P=\{id\}. Then Φ(P)=ϕ1,15\Phi(P)=\phi_{1,15}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=idw=id. We have Krew(W,ϕ1,15,10s+1;q)=[10s][10s4][10s8]/([2][6][10])\operatorname{\textnormal{Krew}}(W,\phi_{1,15},10s+1;q)=[10s][10s-4][10s-8]/([2][6][10]), thus

    Krew(W,ϕ1,15,10s+1;ωd)={s(5s2)(5s4)/3 if d{1,2},5s/3 if 3s and d{3,6},s if d{5,10},0 otherwise.\operatorname{\textnormal{Krew}}(W,\phi_{1,15},10s+1;\omega_{d})=\left\{\begin{aligned} &s(5s-2)(5s-4)/3&\textnormal{ if }d\in\{1,2\},\\ &5s/3&\textnormal{ if }3\mid s\textnormal{ and }d\in\{3,6\},\\ &s&\textnormal{ if }d\in\{5,10\},\\ &0&\textnormal{ otherwise.}\end{aligned}\right.

    This coincides with φ(id,s,d)\varphi(id,s,d).

We exhaust all the possible cases and thus Theorem 3.6 is valid when WW is of type H3H_{3}.

7. Type I2(m)I_{2}(m)

In this section we assume that WW is of type I2(m)I_{2}(m) for m2m\geq 2.

7.1. qq-Kreweras numbers and positivity.

Here we list Krew(W,χ,t;q)\operatorname{\textnormal{Krew}}(W,\chi,t;q) for χIrr(W)\chi\in\operatorname{\textup{Irr}}(W) in the same way as type H3H_{3} case. When mm is even we have:

Irr(W)\operatorname{\textup{Irr}}(W) qq-Kreweras Qχ,V\langle Q_{\chi},V\rangle parabolicparabolic
ϕ1,0\phi_{1,0} q2t2q^{2t-2} 0 I2(m)I_{2}(m)
ϕ1,m/2\phi_{1,m/2}^{\prime} qtm+1[t1]/[2]q^{t-m+1}[t-1]/[2] qq A1A_{1}^{\prime}
ϕ1,m/2′′\phi_{1,m/2}^{\prime\prime} qtm+1[t1]/[2]q^{t-m+1}[t-1]/[2] qq A1′′A_{1}^{\prime\prime}
ϕ1,m\phi_{1,m} [t1][tm+1]/([2][m])[t-1][t-m+1]/([2][m]) qm1+qq^{m-1}+q trivtriv
ϕ2,r,r[1,m/21]\phi_{2,r},r\in[1,m/2-1] qt2r1(qt11)q^{t-2r-1}(q^{t-1}-1) qq

When mm is odd we have:

Irr(W)\operatorname{\textup{Irr}}(W) qq-Kreweras Qχ,V\langle Q_{\chi},V\rangle parabolicparabolic
ϕ1,0\phi_{1,0} q2t2q^{2t-2} 0 I2(m)I_{2}(m)
ϕ1,m\phi_{1,m} [t1][tm+1]/([2][m])[t-1][t-m+1]/([2][m]) qm1+qq^{m-1}+q trivtriv
ϕ2,r,r[1,(m3)/2]\phi_{2,r},r\in[1,(m-3)/2] qt2r1(qt11)q^{t-2r-1}(q^{t-1}-1) qq
ϕ2,(m1)/2\phi_{2,(m-1)/2} qtm+1[t1]q^{t-m+1}[t-1] qq A1A_{1}

It is clear from the formula that if t±1(modm)t\equiv\pm 1\pmod{m} we have Krew(W,χ,t;q)[q]\operatorname{\textnormal{Krew}}(W,\chi,t;q)\in\mathbb{N}[q] if and only if χimΦ\chi\in\operatorname{\textup{im}}\Phi.

7.2. Relation with qq-Narayana numbers

We have

Nar(W,k,t;q)=q(t(m2)k1)(2k)[2k]qm2(qt1;qm+2)k(q2;qm2)k\operatorname{\textnormal{Nar}}(W,k,t;q)=q^{(t-(m-2)k-1)(2-k)}\genfrac{[}{]}{0.0pt}{}{2}{k}_{q^{m-2}}\frac{(q^{t-1};q^{-m+2})_{k}}{(q^{2};q^{m-2})_{k}}

or more precisely

Nar(W,k,0;q)\displaystyle\operatorname{\textnormal{Nar}}(W,k,0;q) =q2(t1),\displaystyle=q^{2(t-1)},
Nar(W,k,1;q)\displaystyle\operatorname{\textnormal{Nar}}(W,k,1;q) =qtm+1(qm2+1)[t1][2],\displaystyle=q^{t-m+1}(q^{m-2}+1)\frac{[t-1]}{[2]},
Nar(W,k,2;q)\displaystyle\operatorname{\textnormal{Nar}}(W,k,2;q) =[t1][tm+1][2][m].\displaystyle=\frac{[t-1][t-m+1]}{[2][m]}.

Thus Theorem 3.5 holds in this case by direct calculation.

7.3. Cyclic sieving

We identify WW with s1,s2s12,s22,(s1s2)m\langle s_{1},s_{2}\mid s_{1}^{2},s_{2}^{2},(s_{1}s_{2})^{m}\rangle and fix a (standard) Coxeter element c=s1s2c=s_{1}s_{2}. We have Ref(W)={ci1s1c1i1i(m+1)/2}{ci1s2c1i1im/2]}\operatorname{\textnormal{Ref}}(W)=\{c^{i-1}s_{1}c^{1-i}\mid 1\leq i\leq(m+1)/2\}\cup\{c^{i-1}s_{2}c^{1-i}\mid 1\leq i\leq m/2]\}. One can show that each element is fixed by conjugation by ckc^{k} if and only if mkm\mid k (resp. m/2km/2\mid k) if mm is odd (resp. even).

From now on we calculate the fixed points of NC(s)(W)\operatorname{\textnormal{NC}}^{(s)}(W) under the /ms\mathbb{Z}/ms-action defined in 3.5 case-by-case in the same manner as type H3H_{3} case. Let us write φ(w1,s,d)\varphi(w_{1},s,d) to denote the number of fixed points in NC(s)(W)\operatorname{\textnormal{NC}}^{(s)}(W) starting with w1Ww_{1}\in W by an order dd element (where dmsd\mid ms). Also let Δ\Delta be the sequence obtained from (δ1,,δs1,δs)(\delta_{1},\ldots,\delta_{s-1},\delta_{s}) by removing the identity elements.

  1. (1)

    w1=cw_{1}=c. In this case Δ=\Delta=\emptyset. Thus ϕ(c,s,d)=1\phi(c,s,d)=1.

  2. (2)

    w1w_{1} is a reflection. In this case Δ=(w1c)\Delta=(w_{1}c). Direct calculation shows that ϕ(w1,s,d)\phi(w_{1},s,d) equals ss if dgcd(2,m)d\mid\gcd(2,m) and 0 otherwise.

  3. (3)

    w1=idw_{1}=id. Let us denote by φΔ(id,s,d)\varphi_{\Delta}(id,s,d) the number of fixed elements with fixed Δ\Delta.

    1. If Δ=(c)\Delta=(c), direct calculation shows that ϕ(c)(w1,s,d)\phi_{(c)}(w_{1},s,d) equals ss if dmd\mid m and 0 otherwise.

    2. If Δ\Delta consists of two reflections, then there are mm possibilities of Δ\Delta.

      1. If mm is even then for any Δ=(r,r)\Delta=(r,r^{\prime}), rr and rr^{\prime} are not conjugate by powers of cc. Thus φΔ(w1,s,d)\varphi_{\Delta}(w_{1},s,d) equals s(s1)/2s(s-1)/2 if d2d\mid 2 and 0 otherwise.

      2. If mm is odd then for any Δ=(r,r)\Delta=(r,r^{\prime}) we have r=c(m1)/2rc(m1)/2r^{\prime}=c^{(m-1)/2}rc^{-(m-1)/2}. Thus we have φΔ(w1,s,d)\varphi_{\Delta}(w_{1},s,d) equals s(s1)/2s(s-1)/2 if d=1d=1 and 0 if d2d\nmid 2. Moreover, if 2s2\mid s and Δ=(δi,δs/2+i)\Delta=(\delta_{i},\delta_{s/2+i}) for some i[1,s/2]i\in[1,s/2] then the corresponding δ\delta-sequence is fixed by an order 2 element. Therefore φΔ(w1,s,2)\varphi_{\Delta}(w_{1},s,2) equals s/2s/2 if 2s2\mid s and 0 otherwise.

    In sum, we have

    φ(w1,s,d)={s(msm+2)/2 if d=1 or [2m and d=2],ms/2 if 2m,2s, and d=2,s if dm and d{1,2}, and0 otherwise.\varphi(w_{1},s,d)=\left\{\begin{aligned} &s(ms-m+2)/2&\textnormal{ if }d=1\textnormal{ or }[2\mid m\textnormal{ and }d=2],\\ &ms/2&\textnormal{ if }2\nmid m,2\mid s,\textnormal{ and }d=2,\\ &s&\textnormal{ if }d\mid m\textnormal{ and }d\not\in\{1,2\},\textnormal{ and}\\ &0&\textnormal{ otherwise.}\end{aligned}\right.

Now we consider the qq-Kreweras numbers Krew(W,Φ(P),ms+1;ωd)\operatorname{\textnormal{Krew}}(W,\Phi(P),ms+1;\omega_{d}) where dmsd\mid ms. Direct calculation shows that:

  1. PP is of type I2(m)I_{2}(m). Then Φ(P)=ϕ1,0\Phi(P)=\phi_{1,0}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=cw=c. We have Krew(W,ϕ1,0,ms+1;q)=q2ms\operatorname{\textnormal{Krew}}(W,\phi_{1,0},ms+1;q)=q^{2ms}, thus Krew(W,ϕ1,0,ms+1;ωd)=1\operatorname{\textnormal{Krew}}(W,\phi_{1,0},ms+1;\omega_{d})=1 whenever dmsd\mid ms. This coincides with φ(c,s,d)=1\varphi(c,s,d)=1.

  2. PP is of type A1A_{1}, mm is odd. Then Φ(P)=ϕ2,(m1)/2\Phi(P)=\phi_{2,(m-1)/2}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if ww is a reflection. We have Krew(W,ϕ2,(m1)/2,ms+1;q)=qmsm+2[ms]\operatorname{\textnormal{Krew}}(W,\phi_{2,(m-1)/2},ms+1;q)=q^{ms-m+2}[ms], thus Krew(W,ϕ2,(m1)/2,ms+1;ωd)\operatorname{\textnormal{Krew}}(W,\phi_{2,(m-1)/2},ms+1;\omega_{d}) equals msms if d=1d=1 and 0 otherwise. This coincides with rRef(W)φ(r,s,d)\sum_{r\in\operatorname{\textnormal{Ref}}(W)}\varphi(r,s,d).

  3. PP is of type A1A_{1}, mm is even. By symmetry it suffices to assume that P={id,s1}P=\{id,s_{1}\} and Φ(P)=ϕ1,m/2\Phi(P)=\phi_{1,m/2}^{\prime}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w{cjs1cjj}w\in\{c^{j}s_{1}c^{-j}\mid j\in\mathbb{Z}\} (there are m/2m/2 of them). We have Krew(W,ϕ1,m/2,ms+1;q)=qmsm+2[ms]/[2]\operatorname{\textnormal{Krew}}(W,\phi_{1,m/2}^{\prime},ms+1;q)=q^{ms-m+2}[ms]/[2], thus Krew(W,ϕ1,m/2,ms+1;ωd)\operatorname{\textnormal{Krew}}(W,\phi_{1,m/2}^{\prime},ms+1;\omega_{d}) equals ms/2ms/2 if d2d\mid 2 and 0 otherwise. This coincides with j=1m/2φ(cjs1cj,s,d)\sum_{j=1}^{m/2}\varphi(c^{j}s_{1}c^{-j},s,d).

  4. P={id}P=\{id\}. Then Φ(P)=ϕ1,m\Phi(P)=\phi_{1,m}. For wNC(1)(W)w\in\operatorname{\textnormal{NC}}^{(1)}(W), VwWVPV^{w}\in W\cdot V^{P} if and only if w=idw=id. We have Krew(W,ϕ1,m,ms+1;q)=[ms][msm+2]/([2][m])\operatorname{\textnormal{Krew}}(W,\phi_{1,m},ms+1;q)=[ms][ms-m+2]/([2][m]), thus

    Krew(W,ϕ1,m,ms+1;ωd)={s(msm+2)/2 if d=1 or [2m and d=2],ms/2 if 2m,2s, and d=2,s if dm and d{1,2}, and0 otherwise.\operatorname{\textnormal{Krew}}(W,\phi_{1,m},ms+1;\omega_{d})=\left\{\begin{aligned} &s(ms-m+2)/2&\textnormal{ if }d=1\textnormal{ or }[2\mid m\textnormal{ and }d=2],\\ &ms/2&\textnormal{ if }2\nmid m,2\mid s,\textnormal{ and }d=2,\\ &s&\textnormal{ if }d\mid m\textnormal{ and }d\not\in\{1,2\},\textnormal{ and}\\ &0&\textnormal{ otherwise.}\end{aligned}\right.

    This coincides with φ(id,s,d)\varphi(id,s,d).

We exhaust all the possible cases and thus Theorem 3.6 is valid when WW is of type I2(m)I_{2}(m).

References

  • [AA08] Achar, P. N. and Aubert, A.-M., Springer correspondences for dihedral groups, Transform. Groups 13 (2008), no. 1, 1–24.
  • [Ach19] Achar, P. N., An implementation of the generalized Lusztig-Shoji algorithm, version 0.98f, Available at https://www.math.lsu.edu/~pramod/docs/lsalg.g, 2019.
  • [AH08] Achar, P. N. and Henderson, A., Orbit closures in the enhanced nilpotent cone, Adv. Math. 219 (2008), 27–62.
  • [AHS11] Achar, P. N., Henderson, A., and Sommers, E., Pieces of nilpotent cones for classical groups, Represent. Theory 15 (2011), 584–616.
  • [Arm09] Armstrong, D., Generalized Noncrossing Partitions and Combinatorics of Coxeter Groups, vol. 202, Mem. Amer. Math. Soc., no. 949, American Mathematical Society, Providence, Rhode Island, 2009.
  • [Car93] Carter, R. W., Finite Groups of Lie Type: Conjugacy Classes and Complex Characters, Wiley Classics Library, no. 48, Wiley, 1993.
  • [GG12] Gordon, I. G. and Griffeth, S., Catalan numbers for complex reflection groups, Amer. J. Math. 134 (2012), no. 6, 1491–1502.
  • [GHLMP96] Geck, M., Hiss, G., Lübeck, F., Malle, G., and Pfeiffer, G., Chevie – a system for computing and processing generic character tables, Appl. Algebra Engrg. Comm. Comput. 7 (1996), 175–210.
  • [GNS99] Gyoja, A., Nishiyama, K., and Shimura, H., Invariants for representations of Weyl groups and two-sided cells, J. Math. Soc. Japan 51 (1999), no. 1, 1–34.
  • [Hes79] Hesselink, W. H., Nilpotency in classical groups over a field of characteristic 2, Math. Z. 166 (1979), 165–181.
  • [Kat09] Kato, S., An exotic Deligne-Langlands correspondence for symplectic groups, Duke Math. J. 148 (2009), no. 2, 305–371.
  • [Kat17] by same author, An algebraic study of extension algebras, Amer. J. Math. 139 (2017), no. 3, 567–615.
  • [Kim19] Kim, D., On total Springer representations for the symplectic Lie algebra in characteristic 2 and the exotic case, Available at https://arxiv.org/abs/1909.01901 (2019).
  • [Kre72] Kreweras, G., Sur les partitions non croisees d’un cycle, Discrete Math. 1 (1972), no. 4, 333–350.
  • [Lus76] Lusztig, G., A note on counting nilpotent matrices of fixed rank, Bull. Lond. Math. Soc. 8 (1976), no. 1, 77–80.
  • [Lus86] by same author, Character sheaves, V, Adv. Math. 61 (1986), 103–155.
  • [NRS18] Nandakumar, V., Rosso, D., and Saunders, N., Irreducible components of exotic Springer fibres, J. Lond. Math. Soc. (2) 98 (2018), 609–637.
  • [RS18] Reiner, V. and Sommers, E., Weyl group qq-Kreweras numbers and cyclic sieving, Ann. Comb. 22 (2018), 819–874.
  • [RSS20] Reiner, V., Shepler, A. V., and Sommers, E., Invariant theory for coincidental complex reflection groups, Math. Z. (2020), Available at https://doi.org/10.1007/s00209-020-02592-8.
  • [Sag20] The Sage Developers, Sagemath, the Sage Mathematics Software System (Version 9.2), 2020, http://www.sagemath.org.
  • [Sho82] Shoji, T., On the Green polynomials of a Chevalley group of type F4{F}_{4}, Comm. Algebra 10 (1982), no. 5, 505–543.
  • [Sho83] by same author, On the Green polynomials of classical groups, Invent. Math. 74 (1983), 239–267.
  • [Sho01] by same author, Green functions associated to complex reflection groups, J. Algebra 245 (2001), 650–694.
  • [Sho02] by same author, Green functions associated to complex reflection groups, II, J. Algebra 258 (2002), 563–598.
  • [Sho04] by same author, Green functions attached to limit symbols, Adv. Stud. Pure Math. 40 (2004), 443–467.
  • [Sol63] Solomon, L., Invariants of finite reflection groups, Nagoya Math. J. 22 (1963), 57–64.
  • [Som11] Sommers, E., Exterior powers of the reflection representation in Springer theory, Transform. Groups 16 (2011), no. 3, 889–911.
  • [Spa91] Spaltenstein, N., On the reflection representation in Springer’s theory, Comment. Math. Helv. 66 (1991), 618–636.
  • [SS13] Shoji, T. and Sorlin, K., Exotic symmetric space over a finite field, I, Transform. Groups 18 (2013), no. 3, 877–929.
  • [SS14] by same author, Exotic symmetric space over a finite field, II, Transform. Groups 19 (2014), no. 3, 887–926.
  • [Sta15] Stanley, R. P., Catalan Numbers, Cambridge University Press, New York, NY, 2015.
  • [Sun11] Sun, M., Point stabilisers for the enhanced and exotic nilpotent cones, J. Group Theory 14 (2011), 825–839.
  • [Xue12a] Xue, T., Combinatorics of the Springer correspondence for classical Lie algebras and their duals in characteristic 2, Adv. Math. 230 (2012), 229–262.
  • [Xue12b] by same author, On unipotent and nilpotent pieces for classical groups, J. Algebra 349 (2012), 165–184.