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

Supersymmetric Shimura operators and interpolation polynomials

Siddhartha SAHI Department of Mathematics, Rutgers University
Hill Center for the Mathematical Sciences
110 Frelinghuysen Rd.
Piscataway, NJ 08854-8019
[email protected]
 and  Songhao ZHU Department of Mathematics, Rutgers University
Hill Center for the Mathematical Sciences
110 Frelinghuysen Rd.
Piscataway, NJ 08854-8019
[email protected]
Abstract.

The Shimura operators are a certain distinguished basis for invariant differential operators on a Hermitian symmetric space. Answering a question of Shimura, Sahi–Zhang showed that the Harish-Chandra images of these operators are specializations of certain BCBC-symmetric interpolation polynomials that were defined by Okounkov.

We consider the analogs of Shimura operators for the Hermitian symmetric superpair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) where 𝔤=𝔤𝔩(2p|2q)\mathfrak{g}=\mathfrak{gl}(2p|2q) and 𝔨=𝔤𝔩(p|q)𝔤𝔩(p|q)\mathfrak{k}=\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q) and we prove their Harish-Chandra images are specializations of certain BCBC-supersymmetric interpolation polynomials introduced by Sergeev–Veselov.

1. Introduction

The Harish-Chandra homomorphism for a symmetric space gives an explicit isomorphism of the algebra of invariant differential operators with a certain polynomial algebra, and knowing the Harish-Chandra image of an operator allows one to determine its spectrum. In [Shi90] Shimura introduced a basis for the algebra of invariant differential operators on a Hermitian symmetric space, and formulated the problem of determining their eigenvalues. Shimura’s problem was solved in [SZ19] where it was shown that the Harish-Chandra images of Shimura operators are specializations of certain interpolation polynomials of Type BC introduced by Okounkov [Oko98].

Our main result, Theorem A below, solves the analogous problem for the Hermitian symmetric superpair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) with 𝔤=𝔤𝔩(2p|2q)\mathfrak{g}=\mathfrak{gl}(2p|2q) and 𝔨=𝔤𝔩(p|q)𝔤𝔩(p|q)\mathfrak{k}=\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q).

Let 𝔘\mathfrak{U} be the universal enveloping algebra of 𝔤\mathfrak{g}, then the algebra of invariant differential operators is the quotient of 𝔨\mathfrak{k}-invariants 𝔇:=𝔘𝔨/(𝔘𝔨)𝔨\mathfrak{D}:=\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}. In Section 3.3 we describe a basis 𝒟μ\mathscr{D}_{\mu} of 𝔇\mathfrak{D} indexed by the set (p,q)\mathscr{H}(p,q) of (p,q)(p,q)-hook partitions, which are partitions μ\mu that satisfy μp+1q\mu_{p+1}\leq q. These 𝒟μ\mathscr{D}_{\mu} are the super analogs of Shimura operators, and their definition involves the Cartan decomposition 𝔤=𝔨𝔭=𝔨𝔭+𝔭\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{p}=\mathfrak{k}\oplus\mathfrak{p}^{+}\oplus\mathfrak{p}^{-}, and the multiplicity-free 𝔨\mathfrak{k}-decompositions of 𝔖(𝔭+)\mathfrak{S}(\mathfrak{p}^{+}) and 𝔖(𝔭)\mathfrak{S}(\mathfrak{p}^{-}), which have 𝔨\mathfrak{k}-summands WμW_{\mu} and WμW_{\mu}^{*}, naturally indexed by μ(p,q)\mu\in\mathscr{H}(p,q) [CW01].

The Harish-Chandra homomorphism (Section 2) is an algebra map γ𝟶:𝔇𝔓(𝔞)\gamma^{\mathtt{0}}:\mathfrak{D}\to\mathfrak{P}(\mathfrak{a}^{*}), where 𝔞\mathfrak{a} is an even Cartan subspace of 𝔭\mathfrak{p} and 𝔓(𝔞)\mathfrak{P}(\mathfrak{a}^{*}) is the algebra of polynomials on 𝔞\mathfrak{a}^{*}. For our pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) the image of γ𝟶\gamma^{\mathtt{0}} can be identified with the ring 𝒮p,q𝟶\mathscr{S}^{\mathtt{0}}_{p,q} of even supersymmetric polynomials in p+qp+q variables. In Section 6 we introduce a basis {Jμ:μ(p,q)}\{J_{\mu}:\mu\in\mathscr{H}(p,q)\} of 𝒮p,q𝟶\mathscr{S}^{\mathtt{0}}_{p,q}. These JμJ_{\mu} are suitable specializations of the supersymmetric interpolation polynomials of Type BC introduced by Sergeev–Veselov [SV09], which are characterized up to multiple by certain vanishing properties.

Theorem A.

The Harish-Chandra image of 𝒟μ\mathscr{D}_{\mu} is kμJμk_{\mu}J_{\mu} where kμk_{\mu} is explicitly given in Eq. (55).

We will deduce this from two other results. First, let \mathfrak{Z} be the center of 𝔘\mathfrak{U}, then we have a natural map π:𝔘𝔨𝔇\pi:\mathfrak{Z}\hookrightarrow\mathfrak{U}^{\mathfrak{k}}\twoheadrightarrow\mathfrak{D}, and we prove the following result.

Theorem B.

The map π\pi is surjective. In particular, there exist ZμZ_{\mu}\in\mathfrak{Z} such that π(Zμ)=𝒟μ\pi(Z_{\mu})=\mathscr{D}_{\mu}.

For λ(p,q)\lambda\in\mathscr{H}(p,q) let IλI_{\lambda} be the generalized Verma module for 𝔤\mathfrak{g} obtained by parabolic induction from 𝔮=𝔨𝔭+\mathfrak{q}=\mathfrak{k}\oplus\mathfrak{p}^{+} of the 𝔨\mathfrak{k}-summand WλW_{\lambda} of 𝔖(𝔭+)\mathfrak{S}(\mathfrak{p}^{+}) (Section 5).

Theorem C.

The central element ZμZ_{\mu} acts on IλI_{\lambda} by 0 unless λiμi\lambda_{i}\geq\mu_{i} for all ii.

The special case p=q=1p=q=1 of Theorem A was proved previously in [Zhu22], where it was also shown that the general case would follow if one knew that certain irreducible finite dimensional 𝔤\mathfrak{g}-modules are 𝔨\mathfrak{k}-spherical [Zhu22, Conjecture 1, Theorem A]. For ordinary Lie algebras, the Cartan–Helgason Theorem [Hel00] provides necessary and sufficient conditions for 𝔨\mathfrak{k}-sphericity. However its precise analog for Lie superalgebras remains open and seems quite hard. Therefore in this paper we give a different argument for Theorem A, which does not rely on the 𝔨\mathfrak{k}-sphericity conjecture.

The Shimura problem discussed in this paper is closely related to the Capelli problem studied in [KS93, Sah94] for usual Lie algebras, in [SS16, ASS18], and [SSS20] for Lie superalgebra, and more recently [LSS23, LSS22b, LSS22a] for quantum groups. Thus it is possible that the results of this paper can be generalized to many of the Hermitian symmetric superpairs constructed using Jordan superalgebras, e.g. in [SSS20, Theorem 1.4], and also to the quantum setting, and we hope to address this in future work.

We now briefly discuss the proofs of these results.

The proof of Theorem C is a generalization of an analogous argument in [SZ19] and relies on two key properties of IλI_{\lambda}, namely that it is 𝔨\mathfrak{k}-spherical and has a natural grading by the center of 𝔨\mathfrak{k}.

For Theorem B let 𝔥=𝔞𝔱+\mathfrak{h}=\mathfrak{a}\oplus\mathfrak{t}_{+} be a Cartan subalgebra of 𝔤\mathfrak{g} containing 𝔞\mathfrak{a}, let γ:𝔓(𝔥)\gamma:\mathfrak{Z}\to\mathfrak{P}(\mathfrak{h}^{*}) be the usual Harish-Chandra homomorphism [Mus12], and consider the diagram

(1) {\mathfrak{Z}}𝔇{\mathfrak{D}}𝔓(𝔥){\mathfrak{P}(\mathfrak{h}^{*})}𝔓(𝔞){\mathfrak{P}(\mathfrak{a}^{*})}π\scriptstyle{\pi}γ\scriptstyle{\gamma}γ𝟶\scriptstyle{\gamma^{\mathtt{0}}}𝚁𝚎𝚜\scriptstyle{\operatorname{\mathtt{Res}}}

where 𝚁𝚎𝚜\operatorname{\mathtt{Res}} is the restriction map induced by the decomposition 𝔥=𝔞𝔱+\mathfrak{h}=\mathfrak{a}\oplus\mathfrak{t}_{+}. We first show that the diagram commutes and then we prove that the image of γ\gamma surjects onto the image of γ𝟶\gamma^{\mathtt{0}} using an explicit description of generators of 𝒮p,q𝟶\mathscr{S}^{\mathtt{0}}_{p,q} from [Ste85]. This implies Theorem B.

For Theorem A the main point is to show that γ𝟶(𝒟μ)\gamma^{\mathtt{0}}(\mathscr{D}_{\mu}) satisfies the vanishing properties which characterize JμJ_{\mu}. The action of ZμZ_{\mu} on IλI_{\lambda} can be expressed in terms of its Harish-Chandra image and thus Theorem C, implies certain vanishing properties for γ(Zμ)\gamma(Z_{\mu}). Now we use Diagram 1 to deduce the necessary vanishing properties for γ𝟶(𝒟μ)\gamma^{\mathtt{0}}(\mathscr{D}_{\mu}). This last step requires an additional result that we deduce from [AS15]: consider the set of λ\lambda for which the irreducible quotient VλV_{\lambda} of IλI_{\lambda} is 𝔨\mathfrak{k}-spherical, and let SS be the set of highest 𝔞\mathfrak{a}-weights of these VλV_{\lambda}, then SS is a Zariski dense subset of 𝔞\mathfrak{a}^{*}.

The structure of the paper is as follows. In Section 2, we review some known results related to the pair (𝔤𝔩(2p|2q),𝔤𝔩(p|q)𝔤𝔩(p|q))(\mathfrak{gl}(2p|2q),\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q)). In particular, we discuss the Iwasawa decomposition, the two Harish-Chandra homomorphisms, the Cheng–Wang decomposition of 𝔖(𝔭±)\mathfrak{S}(\mathfrak{p}^{\pm}), and Alldridge’s necessary conditions for 𝔨\mathfrak{k}-sphericity of finite dimensional irreducible 𝔤\mathfrak{g}-modules. We also present the Sergeev–Veselov supersymmetric interpolation polynomials of Type BC. In Section 3, we define the supersymmetric Shimura operators. In Section 4, we describe the image of γ𝟶\gamma^{\mathtt{0}} as the ring of even supersymmetric polynomials. We prove the surjectivity of 𝚁𝚎𝚜\operatorname{\mathtt{Res}} in Subsection 4.1 (Proposition 4.1) and Theorem B in Subsection 4.2. In Section 5 we discuss the generalized Verma modules and prove Theorem C, which provides the necessary representation theoretic machinery for the vanishing properties. Finally in Section 6, we reformulate the Type BC interpolation polynomials in our context. Then we prove Theorem A using the results obtained in previous sections.

2. Preliminaries

Throughout the paper, we assume that \mathbb{N}, the set of natural numbers, includes 0. For a finite dimensional vector space VV, we identify the polynomial algebra on VV, 𝔓(V)\mathfrak{P}(V), with the symmetric algebra 𝔖(V)\mathfrak{S}(V^{*}) on the dual of VV. We denote the \mathbb{C}-vector superspace of superdimension m|nm|n by m|n\mathbb{C}^{m|n} where the even subspace is m\mathbb{C}^{m} and the odd subspace is n\mathbb{C}^{n}. A 2\mathbb{Z}_{2}-graded algebra is called a superalgebra. For any superalgebra AA, we denote its even subspace as A0¯A_{\overline{0}} and the odd subspace as A1¯A_{\overline{1}}. For xAix\in A_{i}, we write |x|=i|x|=i for its parity ii. The space of linear endomorphisms End(m|n)\operatorname{End}(\mathbb{C}^{m|n}) is naturally 2\mathbb{Z}_{2}-graded and identified with the space of (m+n)×(m+n)(m+n)\times(m+n) matrices. We denote the Lie superalgebra structure defined on End(m|n)\operatorname{End}(\mathbb{C}^{m|n}) as 𝔤𝔩(m|n)\mathfrak{gl}(m|n). In this subsection, we denote the standard diagonal Cartan subalgebra of 𝔤𝔩(m|n)\mathfrak{gl}(m|n) as 𝔱\mathfrak{t}, that is, 𝔱=Span{Ei,i:1im+n}\mathfrak{t}=\operatorname{Span}\{E_{i,i}:1\leq i\leq m+n\} where Ei,jE_{i,j} denotes the matrix with 1 in the (i,j)(i,j)-th entry and 0 elsewhere. We also let ϵi\epsilon_{i} and δj\delta_{j} be the coordinate functions of Ei,iE_{i,i} for 1im1\leq i\leq m and of Em+j,m+jE_{m+j,m+j} for 1jn1\leq j\leq n respectively.

2.1. Symmetric superpairs

In this subsection, we let (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) be a symmetric superpair. Specifically, the Cartan decomposition 𝔤=𝔨𝔭\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{p} is given by an involution θ\theta where 𝔨\mathfrak{k} is the fixed point subalgebra and 𝔭\mathfrak{p} is the (1)(-1)-eigenspace of θ\theta. We assume the chosen nondegenerate invariant form bb on 𝔤\mathfrak{g} is θ\theta-invariant. Let 𝔞𝔭0¯\mathfrak{a}\subseteq\mathfrak{p}_{\overline{0}} be a maximal toral subalgebra. Let Σ:=Σ(𝔤,𝔞)\Sigma:=\Sigma(\mathfrak{g},\mathfrak{a}) be the restricted root system of 𝔤\mathfrak{g} with respect to 𝔞\mathfrak{a}, and Σ+\Sigma^{+} be a positive system. We denote the form on 𝔞\mathfrak{a}^{*} by (,)(\cdot,\cdot) induced from bb. As in Subsection 2.2, we say a restricted root α\alpha is anisotropic if (α,α)0(\alpha,\alpha)\neq 0, and isotropic otherwise. We write 𝔤α\mathfrak{g}_{\alpha} for the root space of α\alpha, and let

Σ0¯:={αΣ:𝔤α𝔤0¯0},Σ1¯:={αΣ:𝔤α𝔤1¯0}\Sigma_{\overline{0}}:=\{\alpha\in\Sigma:\mathfrak{g}_{\alpha}\cap\mathfrak{g}_{\overline{0}}\neq 0\},\quad\Sigma_{\overline{1}}:=\{\alpha\in\Sigma:\mathfrak{g}_{\alpha}\cap\mathfrak{g}_{\overline{1}}\neq 0\}

be the sets of even and odd restricted roots respectively. Set Σ0¯+=Σ+Σ0¯\Sigma^{+}_{\overline{0}}=\Sigma^{+}\cap\Sigma_{\overline{0}}. If αΣ\alpha\in\Sigma, but α/2Σ\alpha/2\notin\Sigma, we say α\alpha is indivisible. We define the multiplicity of α\alpha as mα:=dim(𝔤α)0¯dim(𝔤α)1¯m_{\alpha}:=\dim(\mathfrak{g}_{\alpha})_{\overline{0}}-\dim(\mathfrak{g}_{\alpha})_{\overline{1}}. For a positive system Σ+\Sigma^{+}, we define the nilpotent subalgebra for Σ+\Sigma^{+} as 𝔫:=αΣ+𝔤α\mathfrak{n}:=\bigoplus_{\alpha\in\Sigma^{+}}\mathfrak{g}_{\alpha}. We denote the restricted Weyl vector 12αΣ+mαα\frac{1}{2}\sum_{\alpha\in\Sigma^{+}}m_{\alpha}\alpha for Σ+\Sigma^{+} as ρ\rho.

We assume

(2) 𝔤=𝔫𝔞𝔨,\mathfrak{g}=\mathfrak{n}\oplus\mathfrak{a}\oplus\mathfrak{k},

which is called an Iwasawa decomposition for the pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}). For an in-depth discussion and recent developments on Iwasawa decomposition, see [She22].

The Poincaré–Birkhoff–Witt theorem applied to Eq. (2) yields the following identity:

𝔘=(𝔘𝔨+𝔫𝔘)𝔖(𝔞).\mathfrak{U}=\left(\mathfrak{U}\mathfrak{k}+\mathfrak{n}\mathfrak{U}\right)\oplus\mathfrak{S}(\mathfrak{a}).

Let πρ\pi_{\rho} be the respective projection onto 𝔖(𝔞)𝔓(𝔞)\mathfrak{S}(\mathfrak{a})\cong\mathfrak{P}(\mathfrak{a}^{*}), and define Γρ(D)(λ):=πρ(D)(λ+ρ){\Gamma}_{\rho}(D)(\lambda):=\pi_{\rho}(D)(\lambda+\rho) on 𝔘𝔨\mathfrak{U}^{\mathfrak{k}}. The map Γρ{\Gamma}_{\rho} is called the Harish-Chandra homomorphism associated with (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) (with respect to the choice Σ+Σ\Sigma^{+}\subseteq\Sigma). The quotient isomorphism

γρ𝟶:𝔘𝔨/kerΓρImΓρ\gamma^{\mathtt{0}}_{\rho}:\mathfrak{U}^{\mathfrak{k}}/\ker{\Gamma}_{\rho}\rightarrow\operatorname{Im}{\Gamma}_{\rho}

is called the Harish-Chandra isomorphism associated with (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}). Note Imγρ𝟶=ImΓρ\operatorname{Im}\gamma^{\mathtt{0}}_{\rho}=\operatorname{Im}{\Gamma}_{\rho}.

Next, we introduce Alldridge’s result on spherical representations. Suppose 𝔥𝔞\mathfrak{h}\supseteq\mathfrak{a} is a θ\theta-invariant Cartan subalgebra extended from 𝔞\mathfrak{a}. If 𝔟𝔤\mathfrak{b}\subseteq\mathfrak{g} is a Borel subalgebra containing 𝔥\mathfrak{h}, then we denote the irreducible representation of highest weight λ𝔥\lambda\in\mathfrak{h}^{*} with respect to 𝔟\mathfrak{b} as V(λ,𝔟)V(\lambda,\mathfrak{b}). A 𝔤\mathfrak{g}-module VV is said to be 𝔨\mathfrak{k}-spherical if V𝔨:={vV:X.v=0 for all X𝔨}V^{\mathfrak{k}}:=\{v\in V:X.v=0\textup{ for all }X\in\mathfrak{k}\} is non-zero. A non-zero vector in V𝔨V^{\mathfrak{k}} is called a 𝔨\mathfrak{k}-spherical vector. When the context is clear, we simply say spherical instead of 𝔨\mathfrak{k}-spherical. Define λα:=(λ|𝔞,α)(α,α)\lambda_{\alpha}:=\dfrac{(\lambda|_{\mathfrak{a}},\alpha)}{(\alpha,\alpha)} for anisotropic αΣ\alpha\in\Sigma. We say λ\lambda is high enough if

  1. (1)

    (λ,β)>0(\lambda,\beta)>0 for any isotropic root βΣ+\beta\in\Sigma^{+},

  2. (2)

    λα+mα+2m2α>0\lambda_{\alpha}+m_{\alpha}+2m_{2\alpha}>0 and λα+mα+m2α+1>0\lambda_{\alpha}+m_{\alpha}+m_{2\alpha}+1>0 for any odd anisotropic indivisible root α\alpha.

Then by [AS15, Theorem 2.3, Corollary 2.7], we have the following assertion.

Theorem 2.1.

If λα\lambda_{\alpha}\in\mathbb{N} where αΣ0¯+\alpha\in\Sigma^{+}_{\overline{0}}, λ|𝔥𝔨=0\lambda|_{\mathfrak{h}\cap\mathfrak{k}}=0, and λ\lambda is high enough, then V(λ,𝔟)V(\lambda,\mathfrak{b}) is spherical.

Let 𝔫\mathfrak{n}^{-} be the nilpotent subalgebra for Σ:=Σ+\Sigma^{-}:=-\Sigma^{+} with the corresponding Weyl vector ρ=ρ\rho^{-}=-\rho. We may then write down the opposite Iwasawa decomposition

(3) 𝔤=𝔫𝔞𝔨.\mathfrak{g}=\mathfrak{n}^{-}\oplus\mathfrak{a}\oplus\mathfrak{k}.

From now on, we set

(4) Γ:=Γρ,γ𝟶:=γρ𝟶.{\Gamma}:={\Gamma}_{\rho^{-}},\quad\gamma^{\mathtt{0}}:=\gamma^{\mathtt{0}}_{\rho^{-}}.

The following Lemma is [Zhu22, Lemma 5.1].

Lemma 2.1.

Let VV be a 𝔤\mathfrak{g}-module. If dimV𝔨=1\dim V^{\mathfrak{k}}=1, then D𝔘𝔨D\in\mathfrak{U}^{\mathfrak{k}} acts on V𝔨V^{\mathfrak{k}} by a scalar.

A result of the second author [Zhu22, Proposition 4.1] says that any finite dimensional irreducible 𝔨\mathfrak{k}-spherical 𝔤\mathfrak{g}-module has a spherical vector unique up to constant. Another result in the same work [Zhu22, Theorem 5.2] relates this scalar with the highest weight restricted on 𝔞\mathfrak{a}.

Theorem 2.2.

If V=V(λ,𝔟)V=V(\lambda,\mathfrak{b}) is spherical then dimV𝔨=1\dim V^{\mathfrak{k}}=1, and D𝔘𝔨D\in\mathfrak{U}^{\mathfrak{k}} acts on V𝔨V^{\mathfrak{k}} by the scalar given by Γ(D)(λ|𝔞+ρ){\Gamma}(D)\left(\lambda|_{\mathfrak{a}}+\rho\right).

2.2. The Harish-Chandra homomorphism for (𝔤𝔩(m|n))\mathfrak{Z}(\mathfrak{gl}(m|n))

Let 𝔤=𝔤𝔩(m|n)\mathfrak{g}=\mathfrak{gl}(m|n), and 𝔱\mathfrak{t} be the standard Cartan subalgebra. We denote the root system of 𝔤\mathfrak{g} with respect to 𝔱\mathfrak{t} as Σ(𝔤,𝔱)\Sigma(\mathfrak{g},\mathfrak{t}), and fix a choice of positive system Σ+(𝔤,𝔱)\Sigma^{+}(\mathfrak{g},\mathfrak{t}) in Σ(𝔤,𝔱)\Sigma(\mathfrak{g},\mathfrak{t}). Let 𝔑+\mathfrak{N}^{+} and 𝔑\mathfrak{N}^{-} be the sums of positive root spaces and negative root spaces respectively. Then 𝔤=𝔑𝔱𝔑+\mathfrak{g}=\mathfrak{N}^{-}\oplus\mathfrak{t}\oplus\mathfrak{N}^{+} is the triangular decomposition of 𝔤\mathfrak{g}. We denote the Weyl vector 12αΣ+(𝔤,𝔱)mαα\frac{1}{2}\sum_{\alpha\in\Sigma^{+}(\mathfrak{g},\mathfrak{t})}m_{\alpha}\alpha for Σ+(𝔤,𝔱)\Sigma^{+}(\mathfrak{g},\mathfrak{t}), as ρ𝔱\rho_{\mathfrak{t}}, where mαm_{\alpha} is the multiplicity of αΣ+(𝔤,𝔱)\alpha\in\Sigma^{+}(\mathfrak{g},\mathfrak{t}).

We denote the universal enveloping algebra 𝔘(𝔤)\mathfrak{U}(\mathfrak{g}) as 𝔘\mathfrak{U}, and the center of 𝔘\mathfrak{U} as \mathfrak{Z}. Following [Hum78, Mus12], by the Poincaré–Birkhoff–Witt theorem, we have

(5) 𝔘=𝔖(𝔱)(𝔑𝔘+𝔘𝔑+)\mathfrak{U}=\mathfrak{S}(\mathfrak{t})\oplus(\mathfrak{N}^{-}\mathfrak{U}+\mathfrak{U}\mathfrak{N}^{+})

Let ξ\xi be the respective projection onto 𝔖(𝔱)\mathfrak{S}(\mathfrak{t}). On 𝔖(𝔱)\mathfrak{S}(\mathfrak{t}), we define an automorphism η\eta so that (λ+ρ𝔱)(η(H))=λ(H)(\lambda+\rho_{\mathfrak{t}})(\eta(H))=\lambda(H) for all H𝔖(𝔱)H\in\mathfrak{S}(\mathfrak{t}). This automorphism can be equivalently extended from η(h):=hρ𝔱(h)\eta(h):=h-\rho_{\mathfrak{t}}(h) for h𝔱h\in\mathfrak{t}. The Harish-Chandra isomorphism is then defined as

γ𝔱:=ηξ|.\gamma_{\mathfrak{t}}:=\eta\circ\xi|_{\mathfrak{Z}}.

Denote by χλ\chi_{\lambda} the central character afforded by a 𝔤\mathfrak{g}-module of highest weight λ𝔱\lambda\in\mathfrak{t}^{*}. Then

(6) χλ(z)=ξ(z)=(λ+ρ𝔱)γ𝔱(z)=γ𝔱(z)(λ+ρ𝔱).\chi_{\lambda}(z)=\xi(z)=(\lambda+\rho_{\mathfrak{t}})\gamma_{\mathfrak{t}}(z)=\gamma_{\mathfrak{t}}(z)(\lambda+\rho_{\mathfrak{t}}).

The last equality is due to identification between 𝔓(𝔱)\mathfrak{P}(\mathfrak{t}^{*}) and 𝔖(𝔱)\mathfrak{S}(\mathfrak{t}).

To describe the image of γ𝔱\gamma_{\mathfrak{t}}, we first introduce the supersymmetric polynomials. Let {xi}:={xi}i=1m\{x_{i}\}:=\{x_{i}\}_{i=1}^{m} and {yj}:={yj}j=1n\{y_{j}\}:=\{y_{j}\}_{j=1}^{n} be two sets of independent variables, and set {xi,yj}:={xi}{yj}\{x_{i},y_{j}\}:=\{x_{i}\}\cup\{y_{j}\}. We write f(x1=t,y1=t)f(x_{1}=t,y_{1}=-t) for the polynomial obtained by the substitutions x1=tx_{1}=t and y1=ty_{1}=-t. A polynomial ff in {xi,yj}\{x_{i},y_{j}\} is said to be supersymmetric if

  1. (1)

    ff is invariant under permutations of {xi}\{x_{i}\} and of {yj}\{y_{j}\} separately.

  2. (2)

    f(x1=t,y1=t)f(x_{1}=t,y_{1}=-t) is independent of tt.

Let I(xi,yj)I_{\mathbb{C}}(x_{i},y_{j}) denote the \mathbb{C}-algebra of supersymmetric polynomials in {xi,yj}\{x_{i},y_{j}\}. By [Ste85, Theorem 1] (c.f. [Mus12, Theorem 12.4.1]), I(xi,yj)I_{\mathbb{C}}(x_{i},y_{j}) is generated by the power sums:

(7) pr(m,n)(xi,yj):=i=1mxir(1)rj=1nyjr,r>0.p_{r}^{(m,n)}(x_{i},y_{j}):=\sum_{i=1}^{m}x_{i}^{r}-(-1)^{r}\sum_{j=1}^{n}y_{j}^{r},\;r\in\mathbb{Z}_{>0}.

We record the following standard result about γ\gamma where we identify {xi,yj}\{x_{i},y_{j}\} as standard basis for 𝔱\mathfrak{t}. Thus xi=Ei,ix_{i}=E_{i,i} for 1im1\leq i\leq m, and yj=Em+j,m+jy_{j}=E_{m+j,m+j} for 1jn1\leq j\leq n. We refer to [Mus12, Theorem 13.1.1, Theorem 13.4.1], which are based on the original works by Kac, Serganova, and Gorelik [Kac84, Ser99, Gor04].

Theorem 2.3.

The homomorphism γ𝔱\gamma_{\mathfrak{t}} is an isomorphism and Imγ𝔱=I(xi,yj)\operatorname{Im}\gamma_{\mathfrak{t}}=I_{\mathbb{C}}(x_{i},y_{j}).

We also use the notation Λ(𝔱)\Lambda(\mathfrak{t}^{*}) for the algebra of supersymmetric polynomials on 𝔱\mathfrak{t}^{*} when we suppress the choice of coordinates. Then we have:

(8) Imγ𝔱=Λ(𝔱),𝛾Λ(𝔱).\operatorname{Im}\gamma_{\mathfrak{t}}=\Lambda(\mathfrak{t}^{*}),\quad\mathfrak{Z}\xrightarrow[\sim]{\gamma}\Lambda(\mathfrak{t}^{*}).
Remark 1.

The above result is true in a more general term. In particular, Eq. (8) is true when [Mus12, Hypothesis 8.3.4] is assumed, that is, when 𝔤\mathfrak{g} is 𝔤𝔩\mathfrak{gl}, or basic simple, excluding type A(m,n)A(m,n).

2.3. Cheng–Wang decomposition

Let 𝔨=𝔤𝔩(m|n)𝔤𝔩(m|n)\mathfrak{k}=\mathfrak{gl}(m|n)\oplus\mathfrak{gl}(m|n). We recall a multiplicity-free 𝔨\mathfrak{k}-module decomposition, known as Howe duality in [CW01], which generalizes Schmid’s decomposition [Sch70, FK90]. It will allow us to define the supersymmetric Shimura operators in Section 3.

We first introduce some notation. A partition λ\lambda is a sequence of non-negative integers (λ1,λ2,)(\lambda_{1},\lambda_{2},\dots) with only finitely many non-zero terms and λiλi+1\lambda_{i}\geq\lambda_{i+1} (c.f. [Mac95]). Let |λ|:=iλi|\lambda|:=\sum_{i}\lambda_{i} denote the size of λ\lambda, (λ):=max{i:λi>0}\ell(\lambda):=\max\{i:\lambda_{i}>0\} the length of λ\lambda, and λ\lambda^{\prime} for which λi:=|{j:λji}|\lambda^{\prime}_{i}:=|\{j:\lambda_{j}\geq i\}| the transpose of λ\lambda. When viewed as the corresponding Young diagram, λ\lambda is the collection of “boxes” (i,j)(i,j)

{(i,j):1i(λ),1jλi}.\{(i,j):1\leq i\leq\ell(\lambda),1\leq j\leq\lambda_{i}\}.

A (m,n)(m,n)-hook partition is a partition λ\lambda such that λm+1n\lambda_{m+1}\leq n. We define

(m,n):={λ:λm+1n},d(m,n):={λ(m,n):|λ|=d}.\mathscr{H}(m,n):=\left\{\lambda:\lambda_{m+1}\leq n\right\},\quad\mathscr{H}^{d}(m,n):=\left\{\lambda\in\mathscr{H}(m,n):|\lambda|=d\right\}.

For λ(m,n)\lambda\in\mathscr{H}(m,n), we define a (m+n)(m+n)-tuple

(9) λ:=(λ1,,λm,λ1m,,λnm)\lambda^{\natural}:=\left(\lambda_{1},\dots,\lambda_{m},\left\langle\lambda_{1}^{\prime}-m\right\rangle,\dots,\left\langle\lambda_{n}^{\prime}-m\right\rangle\right)

where x:=max{x,0}\langle x\rangle:=\max\{x,0\} for xx\in\mathbb{Z}. The last nn coordinates can be viewed as the lengths of the remaining columns after discarding the first mm rows of λ\lambda.

Let 𝔅\mathfrak{B} be any Borel subalgebra of 𝔤𝔩(m|n)\mathfrak{gl}(m|n) containing a Cartan subalgebra 𝔤𝔩(m|n)0¯\mathfrak{H}\subseteq\mathfrak{gl}(m|n)_{\overline{0}}. Then 𝔅\mathfrak{B} can be described by an ϵδ\epsilon\delta-chain, [X1Xm+n][X_{1}\cdots X_{m+n}], a sequence consisting of characters XiX_{i}\in\mathfrak{H}^{*} where XiXi+1X_{i}-X_{i+1} exhaust all the simple roots defining 𝔅\mathfrak{B}. Therefore,

[ϵ1ϵmδ1δn][\epsilon_{1}\cdots\epsilon_{m}\delta_{1}\cdots\delta_{n}]

gives the standard Borel subalgebra 𝔟st\mathfrak{b}^{\operatorname{st}}, while [δnδ1ϵmϵ1][\delta_{n}\cdots\delta_{1}\epsilon_{m}\cdots\epsilon_{1}] gives the opposite one, denoted as 𝔟op\mathfrak{b}^{\operatorname{op}}. For an (m+n)(m+n)-tuple (a1,,am,b1,,bn)(a_{1},\dots,a_{m},b_{1},\dots,b_{n}), we associate an irreducible 𝔤𝔩(m|n)\mathfrak{gl}(m|n)-module of highest weight i=1maiϵi+j=1nbjδj\sum_{i=1}^{m}a_{i}\epsilon_{i}+\sum_{j=1}^{n}b_{j}\delta_{j} with respect to 𝔟st\mathfrak{b}^{\operatorname{st}}, denoted as L(a1,,am,b1,,bn)L(a_{1},\dots,a_{m},b_{1},\dots,b_{n}).

Let VV be a vector superspace and 𝔖(V)\mathfrak{S}(V) be the supersymmetric algebra on VV, so 𝔖(V)\mathfrak{S}(V) has a natural \mathbb{N}-grading k𝔖n(V)\bigoplus_{k\in\mathbb{N}}\mathfrak{S}^{n}(V), and explicitly, 𝔖d(V)=i+j=d𝔖i(V0¯)j(V1¯)\mathfrak{S}^{d}(V)=\bigoplus_{i+j=d}\mathfrak{S}^{i}(V_{\overline{0}})\otimes\bigwedge^{j}(V_{\overline{1}}) as vector spaces. The natural action of 𝔤𝔩(m|n)\mathfrak{gl}(m|n) on m|n\mathbb{C}^{m|n} gives an action of 𝔨=𝔤𝔩(m|n)𝔤𝔩(m|n)\mathfrak{k}=\mathfrak{gl}(m|n)\oplus\mathfrak{gl}(m|n) on m|nm|n\mathbb{C}^{m|n}\otimes\mathbb{C}^{m|n}, which extends to an action on 𝔖(m|nm|n)\mathfrak{S}(\mathbb{C}^{m|n}\otimes\mathbb{C}^{m|n}). We record the following result regarding the 𝔨\mathfrak{k}-module structure on 𝔖(m|nm|n)\mathfrak{S}(\mathbb{C}^{m|n}\otimes\mathbb{C}^{m|n}). See [CW01, Theorem 3.2].

Theorem 2.4.

The supersymmetric algebra 𝔖(m|nm|n)\mathfrak{S}(\mathbb{C}^{m|n}\otimes\mathbb{C}^{m|n}) as a 𝔨\mathfrak{k}-module is completely reducible and multiplicity-free. In particular,

𝔖d(m|nm|n)=λd(m,n)L(λ)L(λ).\mathfrak{S}^{d}(\mathbb{C}^{m|n}\otimes\mathbb{C}^{m|n})=\bigoplus_{\lambda\in\mathscr{H}^{d}(m,n)}L(\lambda^{\natural})\otimes L(\lambda^{\natural}).

2.4. Supersymmetric Polynomials

We devote this subsection to an overview of known results of supersymmetric polynomials, including the Type BC interpolation polynomials introduced and studied by Sergeev and Veselov in [SV09]. These are super analog for the now-classic Okounkov polynomials [Oko98, OO06]. We specialize them to JμJ_{\mu} (as in Theorem A) in Section 6.

Let {αjF}{αiB}\{\alpha^{\textsc{F}}_{j}\}\cup\{\alpha^{\textsc{B}}_{i}\} be the standard basis for m+n\mathbb{C}^{m+n}, and wjw_{j} and ziz_{i} be the coordinate functions of αjF\alpha^{\textsc{F}}_{j} and αiB\alpha^{\textsc{B}}_{i}, for j=1,,m,i=1,nj=1,\dots,m,i=1\dots,n. Let 𝗄\mathsf{k} and 𝗁\mathsf{h} be two parameters. Following [SV09], we assume that 𝗄>0\mathsf{k}\notin\mathbb{Q}_{>0} (generic), and set 𝗁^\hat{\mathsf{h}} by 2𝗁^1=𝗄1(2𝗁1)2\hat{\mathsf{h}}-1=\mathsf{k}^{-1}(2\mathsf{h}-1). We also set

Cλ+(x;𝗄)\displaystyle C_{\lambda}^{+}(x;\mathsf{k}) :=(i,j)λ(λi+j+𝗄(λj+i)+x)\displaystyle:=\prod_{(i,j)\in\lambda}\left(\lambda_{i}+j+\mathsf{k}(\lambda^{\prime}_{j}+i)+x\right)
Cλ(x;𝗄)\displaystyle C_{\lambda}^{-}(x;\mathsf{k}) :=(i,j)λ(λij𝗄(λji)+x)\displaystyle:=\prod_{(i,j)\in\lambda}\left(\lambda_{i}-j-\mathsf{k}(\lambda^{\prime}_{j}-i)+x\right)

As in [SV09, Section 6], we let ρ:=ρjFαjF+ρiBαiB\rho:=\sum\rho_{j}^{\textsc{F}}\alpha^{\textsc{F}}_{j}+\sum\rho_{i}^{\textsc{B}}\alpha^{\textsc{B}}_{i} where

(10) ρjF:=(𝗁+𝗄j+n),ρiB:=𝗄1(𝗁+12𝗄12+i).\rho_{j}^{\textsc{F}}:=-(\mathsf{h}+\mathsf{k}j+n),\;\rho_{i}^{\textsc{B}}:=-\mathsf{k}^{-1}\left(\mathsf{h}+\frac{1}{2}\mathsf{k}-\frac{1}{2}+i\right).

This is in fact the deformed Weyl vector calculated with the deformed root multiplicities in [SV09]. We identify Pm,n:=[wj,zi]P_{m,n}:=\mathbb{C}[w_{j},z_{i}] with 𝔓(m+n)\mathfrak{P}(\mathbb{C}^{m+n}), and define Λm,n(𝗄,𝗁)Pm,n\Lambda^{(\mathsf{k},\mathsf{h})}_{m,n}\subseteq P_{m,n} as the subalgebra of polynomials ff which: (1) are symmetric separately in shifted variables (wjρjF)(w_{j}-\rho_{j}^{\textsc{F}}) and (ziρiB)(z_{i}-\rho_{i}^{\textsc{B}}), and invariant under their sign changes; (2) satisfy the condition

f(XαiB+αjF)=f(X)f(X-\alpha^{\textsc{B}}_{i}+\alpha^{\textsc{F}}_{j})=f(X)

on the hyperplane wj+𝗄(j1)=𝗄zi+inw_{j}+\mathsf{k}(j-1)=\mathsf{k}z_{i}+i-n. If we equip m+n\mathbb{C}^{m+n} with an inner product defined by

(αmF,αnF)=δm,n,(αmB,αnB)=𝗄δm,n,(αjF,αiB)=0(\alpha^{\textsc{F}}_{m},\alpha^{\textsc{F}}_{n})=\delta_{m,n},\;(\alpha^{\textsc{B}}_{m},\alpha^{\textsc{B}}_{n})=\mathsf{k}\delta_{m,n},\;(\alpha^{\textsc{F}}_{j},\alpha^{\textsc{B}}_{i})=0

Then Condition (2) is equivalent to: (2’) f(X+α)=f(X)f(X+\alpha)=f(X) for any α=αiB±αjF\alpha=\alpha^{\textsc{B}}_{i}\pm\alpha^{\textsc{F}}_{j} on the hyperplane

(Xρ,α)+12(α,α)=0.(X-\rho,\alpha)+\frac{1}{2}(\alpha,\alpha)=0.

For λ(m,n)\lambda\in\mathscr{H}(m,n), we set w(λ)=(λ1,,λn)w(\lambda)=(\lambda^{\prime}_{1},\dots,\lambda^{\prime}_{n}) and z(λ)=(λ1n,,λmn)z(\lambda)=(\left\langle\lambda_{1}-n\right\rangle,\dots,\left\langle\lambda_{m}-n\right\rangle). Equivalently, (z(λ),w(λ))=(λ)(z(\lambda),w(\lambda))=(\lambda^{\prime})^{\natural} (see Eq. (9)). Then [SV09, Proposition 6.3] says the following.

Theorem 2.5.

For each μ(m,n)\mu\in\mathscr{H}(m,n), there exists a unique polynomial IμSV(X;𝗄,𝗁)Λn,m(𝗄,𝗁)I^{\textup{SV}}_{\mu}(X;\mathsf{k,h})\in\Lambda^{(\mathsf{k},\mathsf{h})}_{n,m} of degree 2|μ|2|\mu| such that

IμSV(z(λ),w(λ);𝗄,𝗁)=0,for all λμ,λ(m,n).I^{\textup{SV}}_{\mu}(z(\lambda),w(\lambda);\mathsf{k,h})=0,\quad\text{for all }\lambda\nsupseteq\mu,\,\lambda\in\mathscr{H}(m,n).

and satisfies the normalization condition IμSV(w(μ),z(μ);𝗄,𝗁)=Cμ(1;𝗄)Cμ+(2𝗁1;𝗄)I^{\textup{SV}}_{\mu}(w(\mu),z(\mu);\mathsf{k,h})=C_{\mu}^{-}(1;\mathsf{k})C_{\mu}^{+}(2\mathsf{h}-1;\mathsf{k}). Moreover, {IμSV}\{I^{\textup{SV}}_{\mu}\} is a basis for Λρ\Lambda^{\rho}.

In the same work [SV09], a closely related polynomial is introduced

(11) IμSV(zi,wj;𝗄,𝗁):=dμI^μ(zi,wj;𝗄1,𝗁^),I^{\textup{SV}}_{\mu}(z_{i},w_{j};\mathsf{k},\mathsf{h}):=d_{\mu}\hat{I}_{\mu^{\prime}}(z_{i},w_{j};\mathsf{k}^{-1},\hat{\mathsf{h}}),

for

(12) dμ:=(1)|μ|𝗄2|μ|Cμ(1;𝗄)Cμ(𝗄;𝗄).d_{\mu}:=(-1)^{|\mu|}\mathsf{k}^{2|\mu|}\frac{C_{\mu}^{-}(1;\mathsf{k})}{C_{\mu}^{-}(-\mathsf{k};\mathsf{k})}.

A tableau formula is provided in [SV09, Proposition 6.4] as follows.

Theorem 2.6.

For λ(m,n)\lambda\in\mathscr{H}(m,n), we have

(13) I^λ(zi,wj;𝗄,𝗁)=TφT(𝗄)(i,j)λfT(i,j).\hat{I}_{\lambda}(z_{i},w_{j};\mathsf{k},\mathsf{h})=\sum_{T}\varphi_{T}(-\mathsf{k})\prod_{(i,j)\in\lambda}f_{T}(i,j).

Here TT is any reverse bitableau of type (n,m)(n,m) and shape λ(m,n)\lambda\in\mathscr{H}(m,n), with a filling by symbols 1<<n<1<<m1<\cdots<n<1^{\prime}<\cdots<m^{\prime} (see [SV09, Section 6]). The weight φT\varphi_{T} is defined as in [SV05, Eq. (41)], and fT(i,j)f_{T}(i,j) has leading term za2z_{a}^{2} if T(i,j)=aT(i,j)=a for a=1,,na=1,\dots,n and leading term 𝗄2wb2\mathsf{k}^{2}w_{b}^{2} if T(i,j)=bT(i,j)=b^{\prime} for b=1,,mb^{\prime}=1^{\prime},\dots,m^{\prime}. In fact, [SV09] uses Eq. (13) as the definition for I^λ\hat{I}_{\lambda} and Eq. (11) as a proposition. Here we present the tableau formula as a theorem in parallel with the following result regarding super Jack polynomials.

We recall the theory of super Jack polynomials from [SV05]. The (monic) Jack symmetric functions Pλ(x;θ)P_{\lambda}(x;\theta) are a linear basis for the ring Λ\Lambda of symmetric functions, and the power sums prp_{r} are free generators Λ\Lambda, see e.g. [Mac95]. Let ϕθ\phi_{\theta} be the homomorphism from Λ\Lambda to the polynomial ring [xi,yj]\mathbb{C}[x_{i},y_{j}] which is defined on the generators as follows

ϕθ(pr)=pr,θ(m,n)(xi,yj):=i=1mxir1θj=1nyjr.\phi_{\theta}(p_{r})=p^{(m,n)}_{r,\theta}(x_{i},y_{j}):=\sum_{i=1}^{m}x_{i}^{r}-\frac{1}{\theta}\sum_{j=1}^{n}y_{j}^{r}.

Then the super Jack polynomial SPλSP_{\lambda} is defined by

SPλ=SPλ(xi,yj;θ):=ϕθ(Pλ(x;θ)).SP_{\lambda}=SP_{\lambda}(x_{i},y_{j};\theta):=\phi_{\theta}(P_{\lambda}(x;\theta)).
Theorem 2.7.

We have

(14) SPλ(xi,yj;θ)=TφT(θ)(i,j)λxT(i,j)SP_{\lambda}(x_{i},y_{j};\theta)=\sum_{T}\varphi_{T}(\theta)\prod_{(i,j)\in\lambda}x_{T(i,j)}

where xjx_{j^{\prime}} is yjy_{j} for j=1,,nj=1,\dots,n and for φT\varphi_{T} defined as in [SV05, Eq. (41)].

We also record the following expansion ([Yan92, Sta89]) as in [SZ19] (note τ\tau is θ\theta here)

(15) 1m!(ixi2)m=|μ|=mCμ(1;θ)Pμ(xi2;θ).\frac{1}{m!}\left(\sum_{i}x_{i}^{2}\right)^{m}=\sum_{|\mu|=m}C_{\mu}^{-}(1;-\theta)P_{\mu}(x_{i}^{2};\theta).

Applying ϕθ\phi_{\theta} on both sides, we have

(16) 1m!(p2,θ(m,n)(xi,yj;θ))m=μm(m,n)Cμ(1;θ)SPμ(xi2,yj2;θ).\frac{1}{m!}\left(p_{2,\theta}^{(m,n)}(x_{i},y_{j};\theta)\right)^{m}=\sum_{\mu\in\mathscr{H}^{m}(m,n)}C_{\mu}^{-}(1;-\theta)SP_{\mu}(x_{i}^{2},y_{j}^{2};\theta).

3. Supersymmetric Shimura Operators

From now on, we fix 𝔤=𝔤𝔩(2p|2q)\mathfrak{g}=\mathfrak{gl}(2p|2q) and 𝔨=𝔤𝔩(p|q)𝔤𝔩(p|q)\mathfrak{k}=\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q). This section is devoted to the description of the pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) and important subspaces therein. We then define the supersymmetric Shimura operators and specialize related results introduced above. Throughout the section we set indices i=1,,p,j=1,qi=1,\dots,p,j=1\dots,q, and kk is either ii or jj depending on the context.

3.1. Realization

We fix the following embedding of 𝔨\mathfrak{k} into 𝔤\mathfrak{g}, and identify 𝔨\mathfrak{k} with its image

(17) ((Ap×pBp×qCq×pDq×q),(Ap×pBp×qCq×pDq×q))(Ap×p0p×pBp×q0p×q0p×pAp×p0p×qBp×qCq×p0q×pDq×q0q×q0q×pCq×p0q×qDq×q)\left(\left(\begin{array}[]{c|c}A_{p\times p}&B_{p\times q}\\ \hline\cr C_{q\times p}&D_{q\times q}\end{array}\right),\left(\begin{array}[]{c|c}A_{p\times p}^{\prime}&B_{p\times q}^{\prime}\\ \hline\cr C_{q\times p}^{\prime}&D_{q\times q}^{\prime}\end{array}\right)\right)\mapsto\left(\begin{array}[]{c c |c c }A_{p\times p}&0_{p\times p}&B_{p\times q}&0_{p\times q}\\ 0_{p\times p}&A_{p\times p}^{\prime}&0_{p\times q}&B_{p\times q}^{\prime}\\ \hline\cr C_{q\times p}&0_{q\times p}&D_{q\times q}&0_{q\times q}\\ 0_{q\times p}&C_{q\times p}^{\prime}&0_{q\times q}&D_{q\times q}^{\prime}\end{array}\right)

We let J:=12diag(Ip×p,Ip×p,Iq×q,Iq×q)J:=\frac{1}{2}\operatorname{diag}(I_{p\times p},-I_{p\times p},I_{q\times q},-I_{q\times q}), and θ:=Adexp(iπJ)\theta:=\operatorname{Ad}\exp(i\pi J). Then θ\theta has fixed point subalgebra 𝔨𝔤𝔩(p|q)𝔤𝔩(p|q)\mathfrak{k}\cong\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q). We also have the Harish-Chandra decomposition

𝔤=𝔭𝔨𝔭+\mathfrak{g}=\mathfrak{p}^{-}\oplus\mathfrak{k}\oplus\mathfrak{p}^{+}

where 𝔭+\mathfrak{p}^{+} (respectively 𝔭\mathfrak{p}^{-}) consists of matrices with non-zero entries only in the upper right (respectively bottom left) sub-blocks in each of the four blocks. Set 𝔭:=𝔭𝔭+\mathfrak{p}:=\mathfrak{p}^{-}\oplus\mathfrak{p}^{+}. Then 𝔤=𝔨𝔭\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{p}.

In our theory, we need to work with a θ\theta-stable Cartan subalgebra 𝔥\mathfrak{h} of 𝔤\mathfrak{g} extended from the toral subalgebra 𝔞𝔭0¯\mathfrak{a}\subseteq\mathfrak{p}_{\overline{0}}. Note in [AHZ10, Section 4], it is showed that (𝔤𝔩(r+p|s+q),𝔤𝔩(r|s)𝔤𝔩(p|q))(\mathfrak{gl}(r+p|s+q),\mathfrak{gl}(r|s)\oplus\mathfrak{gl}(p|q)) is of even type if and only if (rp)(sq)0(r-p)(s-q)\geq 0, satisfied by our pair with p=r,q=sp=r,q=s. We present a construction of 𝔥\mathfrak{h} and 𝔞\mathfrak{a} using a certain Cayley transform as follows.

We let ϵi+:=ϵi\epsilon_{i}^{+}:=\epsilon_{i}, ϵi:=ϵp+i\epsilon_{i}^{-}:=\epsilon_{p+i}, δj+:=δj\delta_{j}^{+}:=\delta_{j}, and δj:=δq+j\delta_{j}^{-}:=\delta_{q+j} be the characters on 𝔱\mathfrak{t}. Let

γiB:=ϵi+ϵi,γjF:=δj+δj.\gamma_{i}^{\textsc{B}}:=\epsilon_{i}^{+}-\epsilon_{i}^{-},\quad\gamma_{j}^{\textsc{F}}:=\delta_{j}^{+}-\delta_{j}^{-}.

These are the Harish-Chandra strongly orthogonal roots, and we denote the set of γkB/F\gamma_{k}^{\textsc{B/F}} 111Here B indicates the Boson–Boson block (top left) and F the Fermion–Fermion block (bottom right), c.f. [All12]. as Σ\Sigma_{\perp}. We set Dj,j:=E2p+j,2p+jD_{j,j^{\prime}}:=E_{2p+j,2p+j^{\prime}} for 1j,j2q1\leq j,j^{\prime}\leq 2q. Associated with each γiB\gamma^{\textsc{B}}_{i} is an 𝔰𝔩(2)\mathfrak{sl}(2)-triple spanned by Ei,iEp+i,p+i,Ei,p+iE_{i,i}-E_{p+i,p+i},E_{i,p+i} and Ep+i,iE_{p+i,i} (similarly for γjF\gamma^{\textsc{F}}_{j} with Dj,jD_{j,j^{\prime}}). It is not hard to see that all (p+q)(p+q) 𝔰𝔩(2)\mathfrak{sl}(2)-triples commute. We write 𝗂\mathsf{i} for the imaginary unit 1\sqrt{-1} to avoid confusion. Define

ciB:=Adexp(π4𝗂(Ei,p+iEp+i,i)),cjF:=Adexp(π4𝗂(Dj,q+jDq+j,j))c^{\textsc{B}}_{i}:=\operatorname{Ad}\exp\left({\frac{\pi}{4}}\mathsf{i}(-E_{i,p+i}-E_{p+i,i})\right),\quad c^{\textsc{F}}_{j}:=\operatorname{Ad}\exp\left({\frac{\pi}{4}}\mathsf{i}(-D_{j,q+j}-D_{q+j,j})\right)

The product

(18) c:=iciBjcjFc:=\prod_{i}c^{\textsc{B}}_{i}\prod_{j}c^{\textsc{F}}_{j}

is thus a well-defined automorphism on 𝔤\mathfrak{g} as all terms commute. We set

xi\displaystyle x_{i} :=𝗂(Ei,p+iEp+i,i),\displaystyle:=\mathsf{i}(E_{i,p+i}-E_{p+i,i}), xi\displaystyle\quad x^{\prime}_{i} :=Ei,i+Ep+i,p+i,\displaystyle:=E_{i,i}+E_{p+i,p+i}, x±i\displaystyle\quad x_{\pm i} :=12(xi±xi)\displaystyle:=\frac{1}{2}(x^{\prime}_{i}\pm x_{i})
yj\displaystyle y_{j} :=𝗂(Dj,q+jDq+j,j),\displaystyle:=\mathsf{i}(D_{j,q+j}-D_{q+j,j}), yj\displaystyle\quad y^{\prime}_{j} :=Dj,j+Dq+j,q+j,\displaystyle:=D_{j,j}+D_{q+j,q+j}, y±j\displaystyle\quad y_{\pm j} :=12(yj±yj)\displaystyle:=\frac{1}{2}(y^{\prime}_{j}\pm y_{j})

Then by a direct (rank 1) computation, we see that under cc:

(19) Ei,ix+i,Ep+i,p+ixi,Dj,jy+j,Dq+j,q+jyj.E_{i,i}\mapsto x_{+i},\quad E_{p+i,p+i}\mapsto x_{-i},\quad D_{j,j}\mapsto y_{+j},\quad D_{q+j,q+j}\mapsto y_{-j}.

We now define

(20) 𝔥:=c(𝔱),𝔞:=Span{xi,yj},𝔱+:=Span{xi,yj}.(Note 𝔥=𝔞𝔱+.)\mathfrak{h}:=c(\mathfrak{t}),\quad\mathfrak{a}:=\operatorname{Span}_{\mathbb{C}}\{x_{i},y_{j}\},\quad\mathfrak{t}_{+}:=\operatorname{Span}_{\mathbb{C}}\{x^{\prime}_{i},y^{\prime}_{j}\}.\quad(\text{Note }\mathfrak{h}=\mathfrak{a}\oplus\mathfrak{t}_{+}.)

In 𝔱\mathfrak{t}, the space 𝔱:=Span{Ei,iEp+i,p+i,Dj,jDq+j,q+j}\mathfrak{t}_{-}:=\operatorname{Span}_{\mathbb{C}}\{E_{i,i}-E_{p+i,p+i},D_{j,j}-D_{q+j,q+j}\} is the orthogonal complement of 𝔱+\mathfrak{t}_{+} with respect to the Killing form on 𝔤\mathfrak{g}. Also, on 𝔥\mathfrak{h} we let

αiB,αjF,τiB,τjF𝔥\alpha^{\textsc{B}}_{i},\alpha^{\textsc{F}}_{j},\tau_{i}^{\textsc{B}},\tau_{j}^{\textsc{F}}\in\mathfrak{h}^{*}

be dual to xi,yj,xi,yj𝔥x_{i},y_{j},x^{\prime}_{i},y^{\prime}_{j}\in\mathfrak{h} respectively. Then τkB/F\tau_{k}^{\textsc{B/F}} and αkB/F\alpha_{k}^{\textsc{B/F}} vanish on 𝔞\mathfrak{a} and 𝔱+\mathfrak{t}_{+} respectively. We identify αkB/F\alpha_{k}^{\textsc{B/F}} with its restriction to 𝔞\mathfrak{a}. We also have τiB=12(ϵi++ϵi)\tau_{i}^{\textsc{B}}=\frac{1}{2}(\epsilon_{i}^{+}+\epsilon_{i}^{-}), τjF=12(δj++δj)\tau_{j}^{\textsc{F}}=\frac{1}{2}(\delta_{j}^{+}+\delta_{j}^{-}) and we identify them with their restrictions on 𝔱+\mathfrak{t}_{+}. On 𝔞\mathfrak{a}^{*}, we set

(21) (αmB,αnB)=(αmF,αnF)=δmn,(αiB,αjF)=0,(\alpha_{m}^{\textsc{B}},\alpha_{n}^{\textsc{B}})=-(\alpha_{m}^{\textsc{F}},\alpha_{n}^{\textsc{F}})=\delta_{mn},\quad(\alpha_{i}^{\textsc{B}},\alpha_{j}^{\textsc{F}})=0,

which is induced from bb, the one-half of the supertrace form. For future purposes, we also consider the following basis for 𝔥\mathfrak{h}^{*}:

(22) χ±i:=τiB±αiB,η±j:=τjF±αjF.\chi_{\pm i}:=\tau_{i}^{\textsc{B}}\pm\alpha_{i}^{\textsc{B}},\quad\eta_{\pm j}:=\tau_{j}^{\textsc{F}}\pm\alpha_{j}^{\textsc{F}}.

3.2. Root Data

Recall the superdimension of a super vector space VV is denoted as sdimV=dimV0¯|dimV1¯\operatorname{sdim}V=\dim V_{\overline{0}}|\dim V_{\overline{1}}. Recall that for a (restricted) root space 𝔤α\mathfrak{g}_{\alpha} of αΣ(𝔤,)\alpha\in\Sigma(\mathfrak{g},\cdot), the root multiplicity is defined as mα:=dim(𝔤α)0¯dim(𝔤α)1¯,m_{\alpha}:=\dim(\mathfrak{g}_{\alpha})_{\overline{0}}-\dim(\mathfrak{g}_{\alpha})_{\overline{1}}, and the deformed root multiplicity is defined as

m(α):=12mα.\operatorname{\mathit{m}}(\alpha):=-\frac{1}{2}m_{\alpha}.

We first give an explicit description of Σ(𝔤,𝔥)\Sigma(\mathfrak{g},\mathfrak{h}) using the coordinates defined in Eq. (22). The order << on the indices {±i,±j}\{\pm i,\pm j\} is interpreted as the natural order on \mathbb{Z}, where +i+i is identified with ii. Then we have

(23) Σ0¯+(𝔤,𝔥)\displaystyle\Sigma_{\overline{0}}^{+}(\mathfrak{g},\mathfrak{h}) ={χkχl:k<l}{ηkηl:k<l},\displaystyle=\{\chi_{k}-\chi_{l}:k<l\}\cup\{\eta_{k}-\eta_{l}:k<l\},\quad mα=1\displaystyle m_{\alpha}=1
(24) Σ1¯+(𝔤,𝔥)\displaystyle\Sigma_{\overline{1}}^{+}(\mathfrak{g},\mathfrak{h}) ={χkηl:+1k+p,ql+q}\displaystyle=\{\chi_{k}-\eta_{l}:+1\leq k\leq+p,-q\leq l\leq+q\}
{ηlχk:ql+q,pk1},\displaystyle\cup\{\eta_{l}-\chi_{k}:-q\leq l\leq+q,-p\leq k\leq-1\},\quad mα=1\displaystyle m_{\alpha}=-1

This can be seen by (I) viewing 𝔥\mathfrak{h} as a Weyl group conjugate of the standard Cartan subalgebra 𝔱\mathfrak{t} and identifying the roots expressed using ϵ±\epsilon^{\pm} and δ±\delta^{\pm}, or (II) considering the Cayley transform which in turn gives a correspondence between positive roots on 𝔥\mathfrak{h} and those on 𝔱\mathfrak{t}. Alternatively, the chain in Eq. (35) gives the above choice of positive roots.

We now record restricted root data taken from [Zhu22]. The restricted root system Σ(𝔤,𝔞)\Sigma(\mathfrak{g},\mathfrak{a}) is of Type C. In the following table, we give the standard choice of positivity, together with the superdimensions of the restricted root spaces. The parameters 𝗄,𝗉,𝗊,𝗋,𝗌\mathsf{k,p,q,r,s} are m(α)\operatorname{\mathit{m}}(\alpha), and these are the 5 parameters first introduced by Sergeev and Veselov in [SV09] to study interpolation polynomials. Note only the last column gives the odd restricted roots.

αiB,𝗋\alpha^{\textsc{B}}_{i},\mathsf{r} 2αiB,𝗌2\alpha^{\textsc{B}}_{i},\mathsf{s} αiB±αkB,𝗄\alpha^{\textsc{B}}_{i}\pm\alpha^{\textsc{B}}_{k},\mathsf{k} αjF,𝗉\alpha^{\textsc{F}}_{j},\mathsf{p} 2αjF,𝗊2\alpha^{\textsc{F}}_{j},\mathsf{q} αjF±αlF,𝗄1\alpha^{\textsc{F}}_{j}\pm\alpha^{\textsc{F}}_{l},\mathsf{k}^{-1} αiB±αjF,1\alpha^{\textsc{B}}_{i}\pm\alpha^{\textsc{F}}_{j},1
/ 1|0,121|0,-\frac{1}{2} 2|0,12|0,-1 / 1|0,121|0,-\frac{1}{2} 2|0,12|0,-1 0|2,10|2,1
Table 1. Positive Restricted Roots and m(α)\operatorname{\mathit{m}}(\alpha). 1i<kp,1j<lq1\leq i<k\leq p,1\leq j<l\leq q

Note their ϵi\epsilon_{i} and δp\delta_{p} are our αjF\alpha^{\textsc{F}}_{j} and αiB\alpha^{\textsc{B}}_{i} respectively, due to a different choice of positivity (see [SV09, Eq. (71)]). A direct computation using Eqs. (23, 24) shows the Weyl vector in 𝔥\mathfrak{h}^{*} is

ρ𝔥\displaystyle\rho_{\mathfrak{h}} =i=1p((pi)+12q)(χ+iχi)+j=1q((qi)+12)(η+jηj)\displaystyle=\sum_{i=1}^{p}\left((p-i)+\frac{1}{2}-q\right)(\chi_{+i}-\chi_{-i})+\sum_{j=1}^{q}\left((q-i)+\frac{1}{2}\right)(\eta_{+j}-\eta_{-j})
(25) =i=1p(2(pi)+12q)αiB+j=1q(2(qj)+1)αjF.\displaystyle=\sum_{i=1}^{p}(2(p-i)+1-2q)\alpha^{\textsc{B}}_{i}+\sum_{j=1}^{q}(2(q-j)+1)\alpha^{\textsc{F}}_{j}.

The Weyl vector of restricted roots in 𝔞\mathfrak{a}^{*} is:

(26) ρ=i=1p(2(pi)+12q)αiB+j=1q(2(qj)+1)αjF,\rho=\sum_{i=1}^{p}(2(p-i)+1-2q)\alpha^{\textsc{B}}_{i}+\sum_{j=1}^{q}(2(q-j)+1)\alpha^{\textsc{F}}_{j},

as the restriction of ρ𝔥\rho_{\mathfrak{h}} on 𝔞\mathfrak{a}.

3.3. Supersymmetric Shimura operators

Throughout the subsection, we set i=1,,pi=1,\dots,p and j=1,,qj=1,\dots,q. Recall L(λ)L(\lambda^{\natural}) is the irreducible 𝔤𝔩(p|q)\mathfrak{gl}(p|q)-module of highest weight

λiϵi+λjpδj\sum\lambda_{i}\epsilon_{i}+\sum\langle\lambda^{\prime}_{j}-p\rangle\delta_{j}

with respect to 𝔟st\mathfrak{b}^{\operatorname{st}}, and it is of Type M ([CW12]). In this case, Schur’s Lemma indicates that dimEnd𝔨(L(λ))=1\dim\operatorname{End}_{\mathfrak{k}}\left(L(\lambda^{\natural})\right)=1, and L(λ)L(λ)L(\lambda^{\natural})\otimes L(\lambda^{\natural}) is actually irreducible as a 𝔨\mathfrak{k}-module as 𝔨=𝔤𝔩(p|q)𝔤𝔩(p|q)\mathfrak{k}=\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q). If we let 𝔤𝔩(p|q)\mathfrak{gl}(p|q) act on the second component contragrediently (via a negative supertranspose), then we define the irreducible 𝔨\mathfrak{k}-module L(λ)L(λ)L(\lambda^{\natural})\otimes L^{*}(\lambda^{\natural}) as WλW_{\lambda}. Note both 𝔭\mathfrak{p}^{-} and 𝔭+\mathfrak{p}^{+} are 𝔨\mathfrak{k}-modules by the short grading. We identify 𝔭\mathfrak{p}^{-} as (𝔭+)(\mathfrak{p}^{+})^{*} via the form bb also used for Eq. (21).

Proposition 3.1.

The symmetric superalgebras 𝔖(𝔭+)\mathfrak{S}(\mathfrak{p}^{+}) and 𝔖(𝔭)\mathfrak{S}(\mathfrak{p}^{-}) are completely reducible and multiplicity free as 𝔨\mathfrak{k}-modules. Specifically,

(27) 𝔖d(𝔭+)=λd(p,q)Wλ,𝔖d(𝔭)=λd(p,q)Wλ.\mathfrak{S}^{d}(\mathfrak{p}^{+})=\bigoplus_{\lambda\in\mathscr{H}^{d}(p,q)}W_{\lambda},\quad\mathfrak{S}^{d}(\mathfrak{p}^{-})=\bigoplus_{\lambda\in\mathscr{H}^{d}(p,q)}W_{\lambda}^{*}.
Proof.

By duality, it suffices to show the first equation. First we have 𝔭+m|n(m|n)\mathfrak{p}^{+}\cong\mathbb{C}^{m|n}\otimes(\mathbb{C}^{m|n})^{*}, by identifying m|n\mathbb{C}^{m|n} and (m|n)(\mathbb{C}^{m|n})^{*} as spaces of column and row vectors respectively. The contragredient 𝔨\mathfrak{k}-module structure on (m|n)(\mathbb{C}^{m|n})^{*} is obtained by applying the negative supertranspose on 𝔤𝔩(p|q)\mathfrak{gl}(p|q). Then Theorem 2.4 implies

𝔖d(𝔭+)=λd(p,q)L(λ)L(λ),\mathfrak{S}^{d}(\mathfrak{p}^{+})=\bigoplus_{\lambda\in\mathscr{H}^{d}(p,q)}L(\lambda^{\natural})\otimes L^{*}(\lambda^{\natural}),

proving the claim, c.f. [SSS20, Theorem 1.4] and notations therein. ∎

We write down the highest weight of WλW_{\lambda} with respect to the Borel subalgebra 𝔟st𝔟op\mathfrak{b}^{\operatorname{st}}\oplus\mathfrak{b}^{\operatorname{op}} of 𝔨\mathfrak{k}:

(28) 𝔱λ𝔱\displaystyle\mathfrak{t}^{*}\ni\lambda^{\natural}_{\mathfrak{t}} :=λiϵi++λjpδj+λiϵiλjpδj\displaystyle:=\sum\lambda_{i}\epsilon^{+}_{i}+\sum\langle\lambda_{j}^{\prime}-p\rangle\delta^{+}_{j}-\sum\lambda_{i}\epsilon^{-}_{i}-\sum\langle\lambda_{j}^{\prime}-p\rangle\delta^{-}_{j}
(29) =λiγiB+λjpγjF,γiB,γjFΣ.\displaystyle=\sum\lambda_{i}\gamma_{i}^{\textsc{B}}+\sum\langle\lambda_{j}^{\prime}-p\rangle\gamma_{j}^{\textsc{F}},\quad\gamma_{i}^{\textsc{B}},\gamma_{j}^{\textsc{F}}\in\Sigma_{\perp}.

The part with negative terms is indeed dominant since we take the opposite Borel subalgebra for the second copy of 𝔤𝔩(p|q)\mathfrak{gl}(p|q) in 𝔨=𝔤𝔩(p|q)𝔤𝔩(p|q)\mathfrak{k}=\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q).

As 𝔭±\mathfrak{p}^{\pm} are supercommutative, the respective universal enveloping algebras are just 𝔖(𝔭±)\mathfrak{S}(\mathfrak{p}^{\pm}). The direct summand WλWλW_{\lambda}^{*}\otimes W_{\lambda} embedded in 𝔖(𝔭)𝔖(𝔭+)\mathfrak{S}(\mathfrak{p}^{-})\otimes\mathfrak{S}(\mathfrak{p}^{+}) is then multiplied into 𝔘\mathfrak{U}. We write 1λ1_{\lambda} for the element in (WλWλ)𝔨\left(W_{\lambda}^{*}\otimes W_{\lambda}\right)^{\mathfrak{k}} corresponding to IdWλEnd𝔨(Wλ)\mathrm{Id}_{W_{\lambda}}\in\operatorname{End}_{\mathfrak{k}}(W_{\lambda}). Let 𝔇:=𝔘𝔨/(𝔘𝔨)𝔨\mathfrak{D}:=\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}.

Definition 3.2.

For each λ(p,q)\lambda\in\mathscr{H}(p,q), we let DλD_{\lambda} be the image corresponding to 1λ1_{\lambda} under the composition of the multiplication and the embedding

(WλWλ)𝔨(𝔖(𝔭)𝔖(𝔭+))𝔨\displaystyle\left(W_{\lambda}^{*}\otimes W_{\lambda}\right)^{\mathfrak{k}}\hookrightarrow\left(\mathfrak{S}(\mathfrak{p}^{-})\otimes\mathfrak{S}(\mathfrak{p}^{+})\right)^{\mathfrak{k}}\rightarrow 𝔘𝔨\displaystyle\mathfrak{U}^{\mathfrak{k}} 𝔇\displaystyle\rightarrow\mathfrak{D}
(30) 1λ\displaystyle 1_{\lambda}\xmapsto{\hphantom{\left(W_{\lambda}^{*}\otimes W_{\lambda}\right)^{\mathfrak{k}}\hookrightarrow\left(\mathfrak{S}(\mathfrak{p}^{-})\otimes\mathfrak{S}(\mathfrak{p}^{+})\right)^{\mathfrak{k}}\rightarrow\mathfrak{U}^{\mathfrak{k}}}} Dλ\displaystyle D_{\lambda} 𝒟λ\displaystyle\mapsto\mathscr{D}_{\lambda}

The element 𝒟λ\mathscr{D}_{\lambda} is called the supersymmetric Shimura operator associated with the partition λ\lambda.

Remark 2.

For an irreducible Hermitian symmetric space G/KG/K, such DλD_{\lambda} can be similarly defined (c.f. [SZ19]) in 𝔇(G)\mathfrak{D}(G), the algebra of differential operators on GG, identified with the universal enveloping algebra of the Lie algebra of GG. As DλD_{\lambda} commutes with KK, its right action descends to 𝒟λ𝔇\mathscr{D}_{\lambda}\in\mathfrak{D}, the algebra of differential operators on G/KG/K. These 𝒟λ\mathscr{D}_{\lambda} are the original Shimura operators. As we will study the action of DλD_{\lambda} on 𝔤\mathfrak{g}-modules, we call Dλ𝔘𝔨D_{\lambda}\in\mathfrak{U}^{\mathfrak{k}} as (supersymmetric) Shimura operators as well, by a slight abuse of name. Working with the lift Dλ𝔘𝔨D_{\lambda}\in\mathfrak{U}^{\mathfrak{k}} of 𝒟λ𝔘𝔨\mathscr{D}_{\lambda}\in\mathfrak{U}^{\mathfrak{k}} gives tremendous flexibility. By the definitions of Γ{\Gamma}, γ𝟶\gamma^{\mathtt{0}}, DλD_{\lambda} and 𝒟λ\mathscr{D}_{\lambda}, we see

(31) γ𝟶(𝒟λ)=Γ(Dλ).\gamma^{\mathtt{0}}(\mathscr{D}_{\lambda})={\Gamma}(D_{\lambda}).

3.4. Specialized Results

We now specialize some of the results we introduced in Section 2. First, we set the choice of positive roots Σ+(𝔤,𝔱)\Sigma^{+}(\mathfrak{g},\mathfrak{t}) according to the chain (c.f. Subsection 2.3)

(32) [ϵ1+ϵp+δ1+δq+δqδ1ϵpϵ1][\epsilon_{1}^{+}\cdots\epsilon_{p}^{+}\delta_{1}^{+}\cdots\delta_{q}^{+}\delta_{q}^{-}\cdots\delta_{1}^{-}\epsilon_{p}^{-}\cdots\epsilon_{1}^{-}]

which equivalently gives a Borel subalgebra of 𝔤\mathfrak{g}, denoted as 𝔟\mathfrak{b}^{\natural}. In fact, we have

(33) 𝔟=𝔟st𝔟op𝔭+.\mathfrak{b}^{\natural}=\mathfrak{b}^{\operatorname{st}}\oplus\mathfrak{b}^{\operatorname{op}}\oplus\mathfrak{p}^{+}.

Note the Cayley transform cc as in Eq. (18) allows us to send the positivity on one Cartan subalgebra to that of the other Cartan subalgebra. Specifically, we have c1:𝔥𝔱c^{-1}:\mathfrak{h}\rightarrow\mathfrak{t}, and the dual map c1:𝔱𝔥c^{-1}_{*}:\mathfrak{t}^{*}\rightarrow\mathfrak{h}^{*}. It is then a direct computation to check that

(34) c1:ϵi±χ±i,δj±η±j.c^{-1}_{*}:\epsilon^{\pm}_{i}\mapsto\chi_{\pm i},\quad\delta^{\pm}_{j}\mapsto\eta_{\pm j}.

Thus, the induced choice of positivity of Σ(𝔤,𝔥)\Sigma(\mathfrak{g},\mathfrak{h}) is given by

(35) [χ+1χ+pη+1η+qηqη1χpχ1].[\chi_{+1}\cdots\chi_{+p}\eta_{+1}\cdots\eta_{+q}\eta_{-q}\cdots\eta_{-1}\chi_{-p}\cdots\chi_{-1}].

We denote the corresponding Borel subalgebra as 𝔟\mathfrak{b}. Specifically, when restricted to 𝔞\mathfrak{a}^{*}, 2αiB,2αjF2\alpha^{\textsc{B}}_{i},2\alpha^{\textsc{F}}_{j}, αiB±αiB\alpha^{\textsc{B}}_{i}\pm\alpha^{\textsc{B}}_{i^{\prime}}, αjF±αjF\alpha^{\textsc{F}}_{j}\pm\alpha^{\textsc{F}}_{j^{\prime}}, and αiB±αjF\alpha^{\textsc{B}}_{i}\pm\alpha^{\textsc{F}}_{j} are positive restricted roots in Σ+=Σ+(𝔤,𝔞)\Sigma^{+}=\Sigma^{+}(\mathfrak{g},\mathfrak{a}).

For the Harish-Chandra isomorphism γ:=γ𝔥\gamma:=\gamma_{\mathfrak{h}} on the non-diagonal 𝔥\mathfrak{h}, we obtain it via the Cayley transform cc. By Eq. (19), cc sends the standard basis for 𝔱\mathfrak{t} the basis {x±i,y±j}\{x_{\pm i},y_{\pm j}\} for 𝔥\mathfrak{h}. Thus by a change of variable using cc, Theorem 2.3 and Eq. (8) imply

(36) γ:𝔓(𝔥)=I(x±i,y±j)=Λ(𝔥).\gamma:\mathfrak{Z}\rightarrow\mathfrak{P}(\mathfrak{h}^{*})=I_{\mathbb{C}}(x_{\pm i},y_{\pm j})=\Lambda(\mathfrak{h}^{*}).

The central character χλ\chi_{\lambda} for λ𝔥\lambda\in\mathfrak{h}^{*} is now given by

(37) χλ(z)=γ(z)(λ+ρ𝔥).\chi_{\lambda}(z)=\gamma(z)(\lambda+\rho_{\mathfrak{h}}).

For the Harish-Chandra isomorphism γ𝟶\gamma^{\mathtt{0}}, we simply apply Eqs. (3, 4) to the choice of 𝔤,𝔨,𝔞\mathfrak{g},\mathfrak{k},\mathfrak{a} and 𝔫\mathfrak{n}^{-} in this context. See Eqs. (3.2, 26) ρ𝔥\rho_{\mathfrak{h}} and ρ\rho, the Weyl vectors of Σ+(𝔤,𝔥)\Sigma^{+}(\mathfrak{g},\mathfrak{h}) and Σ+\Sigma^{+} respectively.

We also need the specialization of the spherical highest weights. Consider the weight λ𝔱\lambda^{\natural}_{\mathfrak{t}} (Eq. (28)) on 𝔱\mathfrak{t} and 𝔟\mathfrak{b}^{\natural} (Eq. (33)). For λ(p,q)\lambda\in\mathscr{H}(p,q), we define

V(λ):=V(λ𝔱,𝔟).V(\lambda^{\natural}):=V(\lambda^{\natural}_{\mathfrak{t}},\mathfrak{b}^{\natural}).

This is the VλV_{\lambda} in the introduction. To detect the 𝔨\mathfrak{k}-sphericity (used in Section 4), we specialize Theorem 2.1, which turns out to be quite simple (see [Zhu22, Proposition 4.4]).

Proposition 3.3.

Let λ(p,q)\lambda\in\mathscr{H}(p,q) and V(λ)V(\lambda^{\natural}) be as above. If λp>λ1p\lambda_{p}>\left\langle\lambda_{1}^{\prime}-p\right\rangle, then dimV(λ)𝔨=1\dim V(\lambda^{\natural})^{\mathfrak{k}}=1.

Proof.

It is showed in [Zhu22, Theorem 4.3] that V(λ)V(\lambda^{\natural}) is always finite dimensional via some combinatorial machinery. This allows us to apply the Cayley transform cc to rewrite 𝔱\mathfrak{t}-weights as 𝔥\mathfrak{h}-weights by [Zhu22, Proposition 3.3]. The resulting weight is

(38) 2λ𝔥:=i=1p2λiαiB+j=1q2λjpαjF𝔥2\lambda_{\mathfrak{h}}^{\natural}:=\sum_{i=1}^{p}2\lambda_{i}\alpha_{i}^{\textsc{B}}+\sum_{j=1}^{q}2\left\langle\lambda_{j}^{\prime}-p\right\rangle\alpha_{j}^{\textsc{F}}\in\mathfrak{h}^{*}

which vanishes on 𝔱+=𝔥𝔨\mathfrak{t}_{+}=\mathfrak{h}\cap\mathfrak{k}. The high enough condition degenerates to (2λ𝔥|𝔞,β)>0(2\lambda_{\mathfrak{h}}^{\natural}|_{\mathfrak{a}},\beta)>0 for β=αiB±αjF\beta=\alpha^{\textsc{B}}_{i}\pm\alpha^{\textsc{F}}_{j}. Then a computation shows the claim. ∎

Note that by definition of our notation, we have

(39) V(λ)=V(λ𝔱,𝔟)=V(2λ𝔥,𝔟).V(\lambda^{\natural})=V(\lambda^{\natural}_{\mathfrak{t}},\mathfrak{b}^{\natural})=V(2\lambda_{\mathfrak{h}}^{\natural},\mathfrak{b}).

We write 2λ¯2\overline{\lambda^{\natural}} for the restriction of 2λ𝔥2\lambda_{\mathfrak{h}}^{\natural} to 𝔞\mathfrak{a}. Then by Theorem 2.2, for spherical V(λ)V(\lambda^{\natural}), Dμ𝔘𝔨D_{\mu}\in\mathfrak{U}^{\mathfrak{k}} acts on V(λ)𝔨V(\lambda^{\natural})^{\mathfrak{k}} by the scalar Γ(Dμ)(2λ¯+ρ){\Gamma}(D_{\mu})(2\overline{\lambda^{\natural}}+\rho).

Remark 3.

For p=q=1p=q=1, it is proved that every λ(1,1)\lambda\in\mathscr{H}(1,1) corresponds to a spherical V(λ)V(\lambda^{\natural}) [Zhu22, Theorem 1] utilizing Kac induction which appears to be difficult for higher ranks.

4. Harish-Chandra Isomorphisms and a Commutative Diagram

Let γ\gamma, Γ{\Gamma}, and γ𝟶\gamma^{\mathtt{0}} be as in Section 3. We have Imγ=Λ(𝔥)\operatorname{Im}\gamma=\Lambda(\mathfrak{h}^{*}), the ring of supersymmetric polynomials on 𝔥\mathfrak{h}^{*}. The image of Γ{\Gamma} (equivalently of γ𝟶\gamma^{\mathtt{0}}), however, is less understood in such terms. Alldridge’s work [All12] formulates Imγ𝟶\operatorname{Im}\gamma^{\mathtt{0}} as the intersection of some subalgebras of 𝔖(𝔞)\mathfrak{S}(\mathfrak{a}), while the second author proved a Weyl groupoid invariance formulation in Type A in [Zhu22]. We will study the Imγ𝟶\operatorname{Im}\gamma^{\mathtt{0}} for 𝔤=𝔤𝔩(2p|2q)\mathfrak{g}=\mathfrak{gl}(2p|2q) and 𝔨=𝔤𝔩(p|q)𝔤𝔩(p|q)\mathfrak{k}=\mathfrak{gl}(p|q)\oplus\mathfrak{gl}(p|q) in this section. Specifically, we will give an independent proof that kerΓ=(𝔘𝔨)𝔨\ker{\Gamma}=(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}} and ImΓ=Imγ𝟶=Λ𝟶(𝔞)\operatorname{Im}{\Gamma}=\operatorname{Im}\gamma^{\mathtt{0}}=\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}).

4.1. Image of γ𝟶\gamma^{\mathtt{0}}

We introduce even supersymmetric polynomials on par with what we did in Subsection 2.2. Let {xi}:={xi}i=1m\{x_{i}\}:=\{x_{i}\}_{i=1}^{m} and {yj}:={yj}j=1n\{y_{j}\}:=\{y_{j}\}_{j=1}^{n} be two sets of independent variables, and set {xi,yj}:={xi}{yj}\{x_{i},y_{j}\}:=\{x_{i}\}\cup\{y_{j}\}. We write f(x1=t,y1=t)f(x_{1}=t,y_{1}=-t) for the polynomial obtained by the substitutions x1=tx_{1}=t and y1=ty_{1}=-t. A polynomial ff in {xi,yj}\{x_{i},y_{j}\} is said to be even supersymmetric if (1) ff is invariant under permutations of {xi}\{x_{i}\} and of {yj}\{y_{j}\} separately, and invariant under sign changes of {xi,yj}\{x_{i},y_{j}\} and (2) f(x1=t,y1=t)f(x_{1}=t,y_{1}=-t) is independent of tt. Denote the algebra of even supersymmetric polynomials in {xi,yj}\{x_{i},y_{j}\} as 𝒮m,n𝟶\mathscr{S}^{\mathtt{0}}_{m,n}. It is not hard to see that an even supersymmetric polynomial is supersymmetric in {xi2,yj2}\{x_{i}^{2},y_{j}^{2}\}, so 𝒮m,n𝟶\mathscr{S}^{\mathtt{0}}_{m,n} is generated by p2r(m,n)p_{2r}(m,n) in Eq. (7) for r>0r\in\mathbb{Z}_{>0}.

Let {ϵi}i=1m{δj}j=1n\{\epsilon_{i}\}_{i=1}^{m}\cup\{\delta_{j}\}_{j=1}^{n} be the standard basis for V=m+nV=\mathbb{C}^{m+n}, equipped with an inner product (,)(\cdot,\cdot) such that (ϵi,ϵj)=(δi,δj)=cδi,j(\epsilon_{i},\epsilon_{j})=-(\delta_{i},\delta_{j})=c\delta_{i,j} and (ϵi,δj)=0(\epsilon_{i},\delta_{j})=0 for some c>0c>0. Denote the coordinate functions for this basis as {xi}\{x_{i}\} and {yj}\{y_{j}\}. Then Condition (2) is equivalent to (see [Mus12]) (2’) f(X+ϵ1δ1)=f(X)f(X+\epsilon_{1}-\delta_{1})=f(X) if (X,ϵ1δ1)=0(X,\epsilon_{1}-\delta_{1})=0. As such, 𝒮m,n𝟶\mathscr{S}^{\mathtt{0}}_{m,n} is identified as a subalgebra of 𝔓(V)\mathfrak{P}(V^{*}), and denoted as Λ𝟶(V)\Lambda^{\mathtt{0}}(V^{*}) (c.f. Λ(V)\Lambda(V^{*})).

Let W0W_{0} be the Weyl group associated with the restricted root system Σ:=Σ(𝔤,𝔞)\Sigma:=\Sigma(\mathfrak{g},\mathfrak{a}) (see Subsection 3.2). Then W0W_{0} is in fact of Type BC and equals (𝒮p(/2)p)×(𝒮q(/2)q)(\mathscr{S}_{p}\ltimes(\mathbb{Z}/2\mathbb{Z})^{p})\times(\mathscr{S}_{q}\ltimes(\mathbb{Z}/2\mathbb{Z})^{q}). Originally in [All12], ImΓ\operatorname{Im}{\Gamma} is given by the intersection of certain subspaces using explicitly chosen elements. In [Zhu22], it is proved that this description is in fact independent of any choice, and we have

(40) ImΓ=Λ𝟶(𝔞),𝔘𝔨/(𝔘𝔨)𝔨γ𝟶Λ𝟶(𝔞).\operatorname{Im}{\Gamma}=\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}),\quad\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}\xrightarrow[\sim]{\gamma^{\mathtt{0}}}\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}).

We will offer a different proof that kerΓ=(𝔘𝔨)𝔨\ker{\Gamma}=(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}} and ImΓ=Imγ𝟶=Λ𝟶(𝔞)\operatorname{Im}{\Gamma}=\operatorname{Im}\gamma^{\mathtt{0}}=\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}) below in Subsection 4.2.

Remark 4.

See [SV11, Zhu22] for an alternative way of writing down (even) supersymmetric polynomials. Namely, we can consider a so-called Weyl groupoid that captures both the usual double symmetry on two sets of coordinates, and the translational invariance on hyperplanes for all isotropic (restricted) roots (as in Condition (2’)). If 𝔚\mathfrak{W} (respectively 𝔚0\mathfrak{W}_{0}) denotes the Weyl groupoid associated with Σ(𝔤,𝔥)\Sigma(\mathfrak{g},\mathfrak{h}) (respectively Σ(𝔤,𝔞)\Sigma(\mathfrak{g},\mathfrak{a})), then Imγ=𝔓(𝔥)𝔚\operatorname{Im}\gamma=\mathfrak{P}(\mathfrak{h}^{*})^{\mathfrak{W}} while Imγ𝟶=𝔓(𝔞)𝔚0\operatorname{Im}\gamma^{\mathtt{0}}=\mathfrak{P}(\mathfrak{a}^{*})^{\mathfrak{W}_{0}}.

In this section, we prove Theorem B. The strategy is to use a commutative diagram (1) and the surjectivity of a certain restriction map we introduce below. This map allows us to show that Λ(𝔥)\Lambda(\mathfrak{h}^{*}) surjects onto Λ𝟶(𝔞)\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}). Loosely speaking, the non-restricted supersymmetry should cover the restricted supersymmetry entirely.

Recall 𝔥=𝔱+𝔞\mathfrak{h}=\mathfrak{t}_{+}\oplus\mathfrak{a}. Let pp be the respective projection from 𝔥\mathfrak{h} to 𝔞\mathfrak{a}. Then we extend pp to a projection homomorphism from 𝔖(𝔥)\mathfrak{S}(\mathfrak{h}) to 𝔖(𝔞)\mathfrak{S}(\mathfrak{a}), again denoted as pp. Under the identification between 𝔖(V)\mathfrak{S}(V) and 𝔓(V)\mathfrak{P}(V^{*}), we let 𝚁𝚎𝚜\operatorname{\mathtt{Res}} be the “restriction” homomorphism corresponding to pp. Specifically, the adjoint map p:𝔞𝔥p^{*}:\mathfrak{a}^{*}\rightarrow\mathfrak{h}^{*} gives a pullback of a polynomial ff defined on 𝔥\mathfrak{h}^{*} to a polynomial gg defined on 𝔞\mathfrak{a}^{*}, and 𝚁𝚎𝚜(f):=g\operatorname{\mathtt{Res}}(f):=g. We denote its restriction on Λ(𝔥)\Lambda(\mathfrak{h}^{*}) as 𝚁𝚎𝚜\operatorname{\mathtt{Res}} too.

Proposition 4.1.

The restriction map 𝚁𝚎𝚜\operatorname{\mathtt{Res}} on Λ(𝔥)\Lambda(\mathfrak{h}^{*}) surjects onto Λ𝟶(𝔞)\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}).

Proof.

We show that 𝚁𝚎𝚜\operatorname{\mathtt{Res}} maps generators of Λ(𝔥)\Lambda(\mathfrak{h}^{*}) to all the generators of Λ𝟶(𝔞)\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}). For this, we choose coordinate systems for both 𝔓(𝔥)\mathfrak{P}(\mathfrak{h}^{*}) and 𝔓(𝔞)\mathfrak{P}(\mathfrak{a}^{*}) as in Section 3. Throughout the proof we set indices i=1,,p,j=1,qi=1,\dots,p,j=1\dots,q. Recall xix_{i}, yjy_{j}, xix_{i}^{\prime}, yjy_{j}^{\prime}, x±ix_{\pm i}, and y±jy_{\pm j} from Subsection 3.1. Also, x±ix_{\pm i} and y±jy_{\pm j} span 𝔥\mathfrak{h} while xix_{i} and yjy_{j} span 𝔞\mathfrak{a} by Eq. (20). Thus by definition, the homomorphism 𝚁𝚎𝚜\operatorname{\mathtt{Res}} on 𝔓(𝔥)=[x±i,y±j]\mathfrak{P}(\mathfrak{h}^{*})=\mathbb{C}[x_{\pm i},y_{\pm j}] makes the substitution

(41) x±i±12xi,y±j±12yj.x_{\pm i}\mapsto\pm\frac{1}{2}x_{i},\quad y_{\pm j}\mapsto\pm\frac{1}{2}y_{j}.

The generators in Eq. (7) of Λ(𝔥)\Lambda(\mathfrak{h}^{*}) become pr(2p,2q)(x±i,y±j)p_{r}^{(2p,2q)}(x_{\pm i},y_{\pm j}). By Eq. (41), we have

𝚁𝚎𝚜(pr(2p,2q)(x±i,y±j))={21rp2r(p,q)(xi,yj), if r is even 0, if r is odd.\operatorname{\mathtt{Res}}(p_{r}^{(2p,2q)}(x_{\pm i},y_{\pm j}))=\begin{cases}2^{1-r}p_{2r}^{(p,q)}(x_{i},y_{j}),&\text{ if }r\text{ is even }\\ 0,&\text{ if }r\text{ is odd}\end{cases}.

The preimage of p2r(p,q)(xi,yj)p_{2r}^{(p,q)}(x_{i},y_{j}) under 𝚁𝚎𝚜\operatorname{\mathtt{Res}} is 2r1p2r(2p,2q)(x±i,y±j)2^{r-1}p_{2r}^{(2p,2q)}(x_{\pm i},y_{\pm j}), meaning the generators of Λ𝟶\Lambda^{\mathtt{0}} are in the image of 𝚁𝚎𝚜\operatorname{\mathtt{Res}}. Therefore, Im𝚁𝚎𝚜=Λ𝟶(𝔞)\operatorname{Im}\operatorname{\mathtt{Res}}=\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}). ∎

The substitution in Eq. (41) also tells us that fΛ(𝔥)f\in\Lambda(\mathfrak{h}^{*}) and 𝚁𝚎𝚜(f)Λ𝟶(𝔞)\operatorname{\mathtt{Res}}(f)\in\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}) are related by

(42) f(iai(χ+iχi)+jbj(η+jηj))=(𝚁𝚎𝚜(f))(i2aiαiB+j2bjαjF).f\left(\sum_{i}a_{i}(\chi_{+i}-\chi_{-i})+\sum_{j}b_{j}(\eta_{+j}-\eta_{-j})\right)=(\operatorname{\mathtt{Res}}(f))\left(\sum_{i}2a_{i}\alpha_{i}^{\textsc{B}}+\sum_{j}2b_{j}\alpha_{j}^{\textsc{F}}\right).

4.2. A commutative diagram

As pointed out in the Introduction, the center \mathfrak{Z} acts on the unique up to constant spherical vector by the character via γ\gamma. Considered as a subalgebra of 𝔘𝔨\mathfrak{U}^{\mathfrak{k}}, there is another way to compute such a character using the Harish-Chandra map γ𝟶\gamma^{\mathtt{0}}. We aim to relate the two Harish-Chandra isomorphisms γ\gamma and γ𝟶\gamma^{\mathtt{0}}.

We denote the quotient map 𝔘𝔨𝔘𝔨/(𝔘𝔨)𝔨\mathfrak{Z}\hookrightarrow\mathfrak{U}^{\mathfrak{k}}\rightarrow\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}} as π\pi. Note by definition, (𝔘𝔨)𝔨𝔘𝔨(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}\subseteq\mathfrak{U}\mathfrak{k} is contained in the kernel of the Harish-Chandra projection, hence in kerΓ\ker{\Gamma}. So the homomorphism Γ:𝔇=𝔘𝔨/(𝔘𝔨)𝔨𝔓(𝔞){\Gamma}^{\prime}:\mathfrak{D}=\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}\rightarrow\mathfrak{P}(\mathfrak{a}^{*}) is well-defined. Then ImΓ=ImΓ\operatorname{Im}{\Gamma}=\operatorname{Im}{\Gamma}^{\prime}. Of course, the bijectivity of Γ{\Gamma}^{\prime} would imply Γ=γ𝟶{\Gamma}^{\prime}=\gamma^{\mathtt{0}}. We form the following diagram first.

(43) {\mathfrak{Z}}𝔇{\mathfrak{D}}Λ(𝔥){\Lambda(\mathfrak{h}^{*})}𝔓(𝔞){\mathfrak{P}(\mathfrak{a}^{*})}π\scriptstyle{\pi}γ\scriptstyle{\gamma}\scriptstyle{\wr}Γ\scriptstyle{{\Gamma}^{\prime}}𝚁𝚎𝚜\scriptstyle{\operatorname{\mathtt{Res}}}

Then Theorem B follows from the two assertions below.

Proposition 4.2.

Diagram (43) is commutative.

Proposition 4.3.

The homomorphism Γ{\Gamma}^{\prime} is the isomorphism γ𝟶\gamma^{\mathtt{0}} and Imγ𝟶=Λ𝟶(𝔞)\operatorname{Im}\gamma^{\mathtt{0}}=\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}).

Proof of Theorem B.

Combining Propositions 4.2 and 4.3, the above diagram (43) gives the commutative diagram (see (1)):

{\mathfrak{Z}}𝔇{\mathfrak{D}}Λ(𝔥){\Lambda(\mathfrak{h}^{*})}Λ𝟶(𝔞){\Lambda^{\mathtt{0}}(\mathfrak{a}^{*})}π\scriptstyle{\pi}γ\scriptstyle{\gamma}\scriptstyle{\wr}γ𝟶\scriptstyle{\gamma^{\mathtt{0}}}\scriptstyle{\wr}𝚁𝚎𝚜\scriptstyle{\operatorname{\mathtt{Res}}}

By Proposition 4.1, 𝚁𝚎𝚜\operatorname{\mathtt{Res}} is surjective. Thus the homomorphism π\pi is also surjective on 𝔇\mathfrak{D}. ∎

Now we supply the proofs of Propositions 4.2 and 4.3.

Proof of Proposition 4.2.

The set of all the hook partitions λ(p,q)\lambda\in\mathscr{H}(p,q) satisfying the condition in Proposition 3.3 give us a family of finite dimensional, irreducible, and spherical 𝔤\mathfrak{g}-module V(λ)V(\lambda^{\natural}). In notation of Eq. (22) (c.f. proof to Proposition 4.1), the corresponding highest weight (see Eq. (38)) is:

(44) 2λ𝔥=i=1pλi(χ+iχi)+j=1qλjp(η+jηj)𝔥,2\lambda^{\natural}_{\mathfrak{h}}=\sum_{i=1}^{p}\lambda_{i}(\chi_{+i}-\chi_{-i})+\sum_{j=1}^{q}\left\langle\lambda_{j}^{\prime}-p\right\rangle(\eta_{+j}-\eta_{-j})\in\mathfrak{h}^{*},

whose restriction to 𝔞\mathfrak{a} is

(45) 2λ𝔥|𝔞=2λ¯=i=1p2λiαiB+j=1q2λjpαjF𝔞.2\lambda^{\natural}_{\mathfrak{h}}|_{\mathfrak{a}}=2\overline{\lambda^{\natural}}=\sum_{i=1}^{p}2\lambda_{i}\alpha_{i}^{\textsc{B}}+\sum_{j=1}^{q}2\left\langle\lambda_{j}^{\prime}-p\right\rangle\alpha_{j}^{\textsc{F}}\in\mathfrak{a}^{*}.

Let us denote the set of weights in Eq. (45) as 𝕂\mathbb{K}. Now, considered as a subset of the integral lattice (2)p+q(2\mathbb{Z})^{p+q} in 𝔞p+q\mathfrak{a}^{*}\cong\mathbb{C}^{p+q}, 𝕂\mathbb{K} is fully determined by p+qp+q many inequalities

x1xp>y1yq0.x_{1}\geq\cdots\geq x_{p}>y_{1}\geq\cdots\geq y_{q}\geq 0.

Then 𝕂\mathbb{K} is Zariski dense in 𝔞\mathfrak{a}^{*} as it is the intersection of an open cone with (2)p+q(2\mathbb{Z})^{p+q}.

For z𝔘𝔨z\in\mathfrak{Z}\subseteq\mathfrak{U}^{\mathfrak{k}}, it acts by a scalar by Lemma 2.1, and by Theorem 2.2, this scalar is

Γ(z)(2λ¯+ρ)=Γ(π(z))(2λ¯+ρ).{\Gamma}(z)(2\overline{\lambda^{\natural}}+\rho)={\Gamma}^{\prime}(\pi(z))(2\overline{\lambda^{\natural}}+\rho).

On the other hand, by Eq. (37), zz acts on the entirety of V(λ)V(\lambda^{\natural}) by γ(z)(2λ𝔥+ρ𝔥)\gamma(z)(2\lambda^{\natural}_{\mathfrak{h}}+\rho_{\mathfrak{h}}). Hence

(46) γ(z)(2λ𝔥+ρ𝔥)=Γ(π(z))(2λ¯+ρ).\gamma(z)(2\lambda^{\natural}_{\mathfrak{h}}+\rho_{\mathfrak{h}})={\Gamma}^{\prime}(\pi(z))(2\overline{\lambda^{\natural}}+\rho).

But 2λ𝔥+ρ𝔥2\lambda^{\natural}_{\mathfrak{h}}+\rho_{\mathfrak{h}} vanishes on 𝔱+\mathfrak{t}_{+}. Therefore, the left side of Eq. (46) is just 𝚁𝚎𝚜(γ(z))(2λ¯+ρ)\operatorname{\mathtt{Res}}(\gamma(z))(2\overline{\lambda^{\natural}}+\rho). This can be verified directly using Eq. (42) with inputs Eqs. (3.2, 26, 44). Thus Eq. (46) implies

𝚁𝚎𝚜(γ(z))=Γ(π(z))\operatorname{\mathtt{Res}}(\gamma(z))={\Gamma}^{\prime}(\pi(z))

on 𝕂+ρ\mathbb{K}+\rho. As the two sides agree on a Zariski dense subset of 𝔞\mathfrak{a}^{*}, they must be equal everywhere. This proves the commutativity of (43). ∎

Proof of Proposition 4.3.

Now we show that ImΓ=ImΓ=Λ𝟶(𝔞)\operatorname{Im}{\Gamma}^{\prime}=\operatorname{Im}{\Gamma}=\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}), and Γ{\Gamma}^{\prime} is injective. For the sake of simplicity, write \mathscr{H} for (p,q)\mathscr{H}(p,q), and let d={λ:|λ|d}\mathscr{H}_{d}=\{\lambda\in\mathscr{H}:|\lambda|\leq d\}. We also let 𝔓2d(𝔞):={f𝔓(𝔞):degf2d}\mathfrak{P}_{2d}(\mathfrak{a}^{*}):=\{f\in\mathfrak{P}(\mathfrak{a}^{*}):\deg f\leq 2d\}, and Λd𝟶(𝔞):={fΛ𝟶(𝔞):degf2d}\Lambda^{\mathtt{0}}_{d}(\mathfrak{a}^{*}):=\{f\in\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}):\deg f\leq 2d\}.

Step 1. Let 𝔓:=𝔘(𝔭)=𝔖(𝔭)𝔖(𝔭+)\mathfrak{P}:=\mathfrak{U}(\mathfrak{p})=\mathfrak{S}(\mathfrak{p}^{-})\mathfrak{S}(\mathfrak{p}^{+}). By the Poincaré–Birkhoff–Witt Theorem, we have 𝔘=𝔘𝔨𝔓\mathfrak{U}=\mathfrak{U}\mathfrak{k}\oplus\mathfrak{P}. This is in fact a 𝔨\mathfrak{k}-module decomposition, and its 𝔨\mathfrak{k}-invariant counterpart is

𝔘𝔨=(𝔘𝔨)𝔨𝔓𝔨.\mathfrak{U}^{\mathfrak{k}}=(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}\oplus\mathfrak{P}^{\mathfrak{k}}.

By Definition 3.2, we see that Dλ(WλWλ)𝔨D_{\lambda}\in(W_{\lambda}^{*}W_{\lambda})^{\mathfrak{k}} constitute a basis for

(47) 𝔓𝔨=λ(WλWλ)𝔨\mathfrak{P}^{\mathfrak{k}}=\bigoplus_{\lambda\in\mathscr{H}}(W_{\lambda}^{*}W_{\lambda})^{\mathfrak{k}}

indexed by λ\lambda\in\mathscr{H}. Let 𝒟λ𝔘𝔨/(𝔘𝔨)𝔨=𝔇\mathscr{D}_{\lambda}\in\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}=\mathfrak{D} be the equivalent class of Dλ𝔘𝔨D_{\lambda}\in\mathfrak{U}^{\mathfrak{k}}. Identifying 𝔓𝔨𝔘𝔨/(𝔘𝔨)𝔨\mathfrak{P}^{\mathfrak{k}}\cong\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}}, we see that 𝒟λ\mathscr{D}_{\lambda} gives a basis for 𝔇\mathfrak{D}.

Step 2. Furthermore, we see that 𝔓𝔨=d0(𝔖d(𝔭)𝔖d(𝔭+))𝔨\mathfrak{P}^{\mathfrak{k}}=\bigoplus_{d\geq 0}(\mathfrak{S}^{d}(\mathfrak{p}^{-})\mathfrak{S}^{d}(\mathfrak{p}^{+}))^{\mathfrak{k}} is graded by non-negative integers according to |λ||\lambda| in Eq. (47) (c.f. Eq. (27)) as a vector superspace. The basis {𝒟λ}\{\mathscr{D}_{\lambda}\} is homogeneous. We further define

𝔇d:=Span{𝒟λ:λd}.\mathfrak{D}_{d}:=\operatorname{Span}\{\mathscr{D}_{\lambda}:\lambda\in\mathscr{H}_{d}\}.

This gives a vector superspace filtration. The reason why we consider filtration instead of grading is that Γ{\Gamma}^{\prime} only preserves filtration due to the ρ\rho shift in its definition. For each filtered degree, let

Γd:𝔇d𝔓2d(𝔞).{\Gamma}^{\prime}_{d}:\mathfrak{D}_{d}\rightarrow\mathfrak{P}_{2d}(\mathfrak{a}^{*}).

Step 3. By Proposition 4.1, 𝚁𝚎𝚜\operatorname{\mathtt{Res}} is surjective onto Λ𝟶(𝔞)\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}), from which we have

Λ𝟶(𝔞)=𝚁𝚎𝚜(γ())\Lambda^{\mathtt{0}}(\mathfrak{a}^{*})=\operatorname{\mathtt{Res}}(\gamma(\mathfrak{Z}))

We also showed that the diagram commutes. Thus the right side is Γ(π()){\Gamma}^{\prime}(\pi(\mathfrak{Z})), which gives Λ𝟶(𝔞)ImΓ\Lambda^{\mathtt{0}}(\mathfrak{a}^{*})\subseteq\operatorname{Im}{\Gamma}^{\prime}. By Step 2., the filtered version of the assertion is

(48) Λd𝟶(𝔞)ImΓd.\Lambda^{\mathtt{0}}_{d}(\mathfrak{a}^{*})\subseteq\operatorname{Im}{\Gamma}^{\prime}_{d}.

Step 4. By Proposition 6.1 below, the polynomials JλJ_{\lambda} with |λ|d|\lambda|\leq d give a basis for Λd𝟶(𝔞)\Lambda^{\mathtt{0}}_{d}(\mathfrak{a}^{*}). Such JλJ_{\lambda} are indexed again by λd\lambda\in\mathscr{H}_{d}. Therefore, dimΛd𝟶(𝔞)=|d|dimImΓd\dim\Lambda^{\mathtt{0}}_{d}(\mathfrak{a}^{*})=|\mathscr{H}_{d}|\geq\dim\operatorname{Im}{\Gamma}^{\prime}_{d}, proving

Λd𝟶(𝔞)=ImΓd\Lambda^{\mathtt{0}}_{d}(\mathfrak{a}^{*})=\operatorname{Im}{\Gamma}^{\prime}_{d}

by Eq. (48). This shows ImΓ=ImΓ=Λ𝟶(𝔞)\operatorname{Im}{\Gamma}^{\prime}=\operatorname{Im}{\Gamma}=\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}), as well as the injectivity of Γ{\Gamma}^{\prime}. Therefore, Γ=γ𝟶{\Gamma}^{\prime}=\gamma^{\mathtt{0}} is bijective, from which we have γ𝟶:𝔇Λ𝟶(𝔞)\gamma^{\mathtt{0}}:\mathfrak{D}\overset{\sim}{\rightarrow}\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}). ∎

5. Generalized Verma Modules

In this section, we study certain generalized Verma modules, denoted as IλI_{\lambda}. These are the 𝔤\mathfrak{g}-modules induced on WλW_{\lambda} (λ(p,q)\lambda\in\mathscr{H}(p,q)) from a parabolic subalgebra containing 𝔨\mathfrak{k} to 𝔤\mathfrak{g}. If we choose a Borel subalgebra of 𝔤\mathfrak{g} as the parabolic subalgebra, then we obtain the usual Verma module, which explains the name. We show that IλI_{\lambda} is spherical. We investigate a natural grading on IλI_{\lambda}, and use it to show that certain central elements ZμZ_{\mu} act trivially on IλI_{\lambda} trivially whenever |λ||μ||\lambda|\leq|\mu|. Further strengthening is also obtained, asserting the same vanishing action of ZμZ_{\mu} for λμ\lambda\nsupseteq\mu. This is Theorem C.

5.1. Basics

Consider the parabolic subalgebra 𝔮:=𝔨𝔭+\mathfrak{q}:=\mathfrak{k}\oplus\mathfrak{p}^{+} in 𝔤\mathfrak{g}. The associated set of roots is given by Σ(𝔨,𝔱)Σ(𝔭+,𝔱)\Sigma(\mathfrak{k},\mathfrak{t})\cup\Sigma(\mathfrak{p}^{+},\mathfrak{t}). Let WλW_{\lambda} be as in the Cheng–Wang decomposition. We extend the 𝔨\mathfrak{k}-action trivially to 𝔭+\mathfrak{p}^{+} to obtain a 𝔮\mathfrak{q}-module structure on WλW_{\lambda}. We define the generalized Verma module as

Iλ:=Ind𝔨+𝔭+𝔤Wλ=𝔘𝔘(𝔮)Wλ.I_{\lambda}:=\operatorname{Ind}_{\mathfrak{k}+\mathfrak{p}^{+}}^{\mathfrak{g}}W_{\lambda}=\mathfrak{U}\otimes_{\mathfrak{U}(\mathfrak{q})}W_{\lambda}.

By the Poincaré–Birkhoff–Witt theorem, we see that 𝔘=𝔖(𝔭)𝔘(𝔨)𝔖(𝔭+)\mathfrak{U}=\mathfrak{S}(\mathfrak{p}^{-})\mathfrak{U}(\mathfrak{k})\mathfrak{S}(\mathfrak{p}^{+}), and we have

Iλ𝔖(𝔭)WλI_{\lambda}\cong\mathfrak{S}(\mathfrak{p}^{-})\otimes W_{\lambda}

both as super vector spaces and as 𝔨\mathfrak{k}-modules. Clearly, IλI_{\lambda} is a weight module and the highest weight is Eq. (28), given with respect to 𝔟\mathfrak{b}^{\natural} (see Eq. (33)). Then V(λ)V(\lambda^{\natural}) is the irreducible quotient of IλI_{\lambda}.

Next, we introduce a grading on IλI_{\lambda}. Recall that in the one-dimensional center of 𝔨\mathfrak{k}, there is a special element JJ (see Section 3) such that the Harish-Chandra decomposition 𝔤=𝔭𝔨𝔭+\mathfrak{g}=\mathfrak{p}^{-}\oplus\mathfrak{k}\oplus\mathfrak{p}^{+} corresponds to (1,0,1)(-1,0,1)-eigenspaces of adJ\operatorname{ad}J, and gives rise to the short grading. We may extend the action of adJ\operatorname{ad}J to 𝔘=𝔖(𝔭)𝔘(𝔨)𝔖(𝔭+)\mathfrak{U}=\mathfrak{S}(\mathfrak{p}^{-})\mathfrak{U}(\mathfrak{k})\mathfrak{S}(\mathfrak{p}^{+}). Let {ξi},{xi}\{\xi_{i}\},\{x_{i}\} and {ηi}\{\eta_{i}\} be homogeneous bases for 𝔭\mathfrak{p}^{-}, 𝔨\mathfrak{k}, and 𝔭+\mathfrak{p}^{+} respectively. Then the Poincaré–Birkhoff–Witt basis is:

{ξi1ξimxj1xjlηk1ηkn:i1im,j1jl,k1kn}.\left\{\xi_{i_{1}}\cdots\xi_{i_{m}}x_{j_{1}}\cdots x_{j_{l}}\eta_{k_{1}}\cdots\eta_{k_{n}}:i_{1}\leq\cdots\leq i_{m},j_{1}\leq\cdots\leq j_{l},k_{1}\leq\cdots\leq k_{n}\right\}.

where a strict inequality of subindices occurs when the corresponding basis vectors are odd. We let

𝔘d:=Span{ξi1ξimxj1xjlηk1ηkn:nm=d}.\mathfrak{U}^{d}:=\operatorname{Span}\{\xi_{i_{1}}\cdots\xi_{i_{m}}x_{j_{1}}\cdots x_{j_{l}}\eta_{k_{1}}\cdots\eta_{k_{n}}:n-m=d\}.
Lemma 5.1.

The universal enveloping algebra 𝔘\mathfrak{U} is graded by adJ\operatorname{ad}J, and 𝔘=d𝔘d\mathfrak{U}=\bigoplus_{d\in\mathbb{Z}}\mathfrak{U}^{d}.

Proof.

The action of JJ on the basis ξi1ξimxj1xjlηk1ηkn\xi_{i_{1}}\cdots\xi_{i_{m}}x_{j_{1}}\cdots x_{j_{l}}\eta_{k_{1}}\cdots\eta_{k_{n}} is given by the scalar nm=dn-m=d in 𝔘d\mathfrak{U}^{d}. The bases for 𝔘d\mathfrak{U}^{d} for all d0d\geq 0 give a partition of the Poincaré–Birkhoff–Witt basis, and the sum is hence direct. Furthermore, the multiplication in 𝔘\mathfrak{U} respects the Lie superbracket, which in turn respects the short grading on 𝔤\mathfrak{g}. Therefore the grading is well-defined. ∎

Let l=|λ|l=|\lambda|. Then IλI_{\lambda} is also endowed with the adJ\operatorname{ad}J grading. Explicitly,

Iλld:=𝔖d(𝔭)WλI_{\lambda}^{l-d}:=\mathfrak{S}^{d}(\mathfrak{p}^{-})\otimes W_{\lambda}

has degree ldl-d for dd\in\mathbb{N}. In particular, its top homogeneous component is Wλ\mathbb{C}\otimes W_{\lambda} of degree ll, isomorphic to WλW_{\lambda}. By Lemma 5.1, since the 𝔤\mathfrak{g} acts by left multiplication on IλI_{\lambda}, we have 𝔘d.IλmIλd+m\mathfrak{U}^{d}.I_{\lambda}^{m}\subseteq I_{\lambda}^{d+m}. This leads to the following lemma.

Lemma 5.2.

Let u𝔖k(𝔭+)u\in\mathfrak{S}^{k}(\mathfrak{p}^{+}). Then u.IλmIλk+mu.I_{\lambda}^{m}\subseteq I_{\lambda}^{k+m}.

Next, we show that IλI_{\lambda} is in fact spherical.

Proposition 5.3.

The module IλI_{\lambda} is spherical. Furthermore, dimIλ𝔨=1\dim I_{\lambda}^{\mathfrak{k}}=1 and Iλ𝔨Iλ0I_{\lambda}^{\mathfrak{k}}\subseteq I_{\lambda}^{0}.

Proof.

As a 𝔨\mathfrak{k}-module, we have Iλν(p,q)WνWλI_{\lambda}\cong\bigoplus_{\nu\in\mathscr{H}(p,q)}W_{\nu}^{*}\otimes W_{\lambda} and each component can be identified as Hom(Wν,Wλ)\operatorname{Hom}(W_{\nu},W_{\lambda}). Only the degree 0 component WλWλEnd(Wλ)W_{\lambda}^{*}\otimes W_{\lambda}\cong\operatorname{End}\left(W_{\lambda}\right) has a one-dimensional 𝔨\mathfrak{k}-invariant subspace by Schur’s Lemma for Type M modules. Hence IλI_{\lambda} is spherical with a unique up to constant spherical vector. ∎

By the above Proposition, ω0\omega\neq 0 in Iλ𝔨I_{\lambda}^{\mathfrak{k}} is unique up to constant. Let us fix such ωIλ𝔨\omega\in I_{\lambda}^{\mathfrak{k}} in the following discussion.

5.2. Vanishing actions

Recall that the supersymmetric Shimura operator DμD_{\mu} in 𝔘𝔨\mathfrak{U}^{\mathfrak{k}} is defined as the image of 1μ1_{\mu} under the following composition of maps (Definition 3.2)

(WμWμ)𝔨(𝔖(𝔭)𝔖(𝔭+))𝔨\displaystyle\left(W_{\mu}^{*}\right.\otimes\left.W_{\mu}\right)^{\mathfrak{k}}\hookrightarrow\left(\mathfrak{S}(\mathfrak{p}^{-})\otimes\mathfrak{S}(\mathfrak{p}^{+})\right)^{\mathfrak{k}}\rightarrow 𝔘𝔨\displaystyle\mathfrak{U}^{\mathfrak{k}}
1μ\displaystyle 1_{\mu}\xmapsto{\hphantom{\left(W_{\mu}^{*}\right.\otimes\left.W_{\mu}\right)^{\mathfrak{k}}\hookrightarrow\left(\mathfrak{S}(\mathfrak{p}^{-})\otimes\mathfrak{S}(\mathfrak{p}^{+})\right)^{\mathfrak{k}}\rightarrow\mathfrak{U}^{\mathfrak{k}}}} Dμ\displaystyle D_{\mu}

In particular DμWμWμD_{\mu}\in W_{\mu}^{*}W_{\mu}. If we take a homogeneous basis {ηiμ}\{\eta^{\mu}_{i}\} for WμW_{\mu} and its dual basis {ξiμ}\{\xi^{-\mu}_{i}\} for WμW_{\mu}^{*}, then DμD_{\mu} can be written as

(49) Dμ=ξiμηiμ.D_{\mu}=\sum\xi^{-\mu}_{i}\eta^{\mu}_{i}.

Let m=|μ|m=|\mu| and l=|λ|l=|\lambda|. We have the following “vanishing” actions (Propositions 5.4 and 5.6).

Proposition 5.4.

If l<ml<m, then Dμ.v0=0D_{\mu}.v^{0}=0 for any v0Iλ0v^{0}\in I_{\lambda}^{0}.

Proof.

In DμD_{\mu}, every ηiμ\eta^{\mu}_{i} in Eq. (49) has degree mm as an element in 𝔖m(𝔭+)\mathfrak{S}^{m}(\mathfrak{p}^{+}). Hence ηiμ.Iλ0Iλm\eta^{\mu}_{i}.I_{\lambda}^{0}\subseteq I_{\lambda}^{m} by Lemma 5.2. The top degree of IλI_{\lambda} is ll. But l<ml<m. Thus Dμ.Iλ0=Iλm={0}D_{\mu}.I_{\lambda}^{0}=I_{\lambda}^{m}=\{0\}. ∎

Viewing Wμ𝔘W_{\mu}\subseteq\mathfrak{U}, we have the representation R:WμIλIλR:W_{\mu}\otimes I_{\lambda}\rightarrow I_{\lambda} given by ηvη.v\eta\otimes v\mapsto\eta.v for ηWμ\eta\in W_{\mu} and vIλv\in I_{\lambda}.

Lemma 5.5.

The image of RR restricted to WμIλ𝔨W_{\mu}\otimes I_{\lambda}^{\mathfrak{k}} is homomorphic to WμW_{\mu}.

Proof.

Let m=|μ|m=|\mu|. For each ηiμ.ωIλm\eta^{\mu}_{i}.\omega\in I_{\lambda}^{m}, if we apply a homogeneous X𝔨X\in\mathfrak{k}, we get

(50) X.(ηiμ.ω)=[X,ηiμ].ω+(1)|X||ηiμ|ηiμ.X.ω=(ad(X)(ηiμ)).ωX.(\eta^{\mu}_{i}.\omega)=[X,\eta^{\mu}_{i}].\omega+(-1)^{|X||\eta^{\mu}_{i}|}\eta^{\mu}_{i}.X.\omega=(\operatorname{ad}(X)(\eta^{\mu}_{i})).\omega

as X.ω=0X.\omega=0. Thus, the map ηiμηiμ.ω\eta^{\mu}_{i}\mapsto\eta^{\mu}_{i}.\omega intertwines with the actions of 𝔨\mathfrak{k} on WμW_{\mu} and Iλ𝔨I_{\lambda}^{\mathfrak{k}}.

Let SS be the subspace of IλmI_{\lambda}^{m} spanned by {ηiμ.ω}\{\eta^{\mu}_{i}.\omega\}. Then SS is a homomorphic image of WμW_{\mu} in IλI_{\lambda} by Eq. (50). ∎

Proposition 5.6.

If l=ml=m but λμ\lambda\neq\mu, then Dμ.ω=0D_{\mu}.\omega=0.

Proof.

Since Wμ𝔖l(𝔭+)W_{\mu}\subseteq\mathfrak{S}^{l}(\mathfrak{p}^{+}) and Iλ𝔨Iλ0I_{\lambda}^{\mathfrak{k}}\subseteq I_{\lambda}^{0}, by Lemma 5.2, we see that

WμIλ𝔨Iλl=Wλ.W_{\mu}\otimes I_{\lambda}^{\mathfrak{k}}\rightarrow I_{\lambda}^{l}=\mathbb{C}\otimes W_{\lambda}.

By Lemma 5.5, WμW_{\mu} is homomorphic to some SS in IλlI_{\lambda}^{l}. Since λμ\lambda\neq\mu, Iλl={0}I_{\lambda}^{l}=\{0\}. Hence, WμW_{\mu} acts on Iλ𝔨I_{\lambda}^{\mathfrak{k}} by 0, and so does DμD_{\mu}. ∎

Corollary 5.7.

For any λ(p,q)\lambda\in\mathscr{H}(p,q) such that |λ||μ||\lambda|\leq|\mu|, λμ\lambda\neq\mu, we have Dμ.ω=0D_{\mu}.\omega=0 in IλI_{\lambda}.

Proof.

It follows from Propositions 5.3, 5.4 for the case |λ|<|μ||\lambda|<|\mu|, and 5.6 for the case |λ|=|μ||\lambda|=|\mu|. ∎

Recall π(Dμ)=𝒟μ\pi(D_{\mu})=\mathscr{D}_{\mu}. By Theorem B, π\pi is surjective. Then it is possible to find ZμZ_{\mu}\in\mathfrak{Z} such that π(Zμ)=𝒟μ\pi(Z_{\mu})=\mathscr{D}_{\mu}. We show that ZμZ_{\mu} acts by 0 on IλI_{\lambda} when λ\lambda satisfies certain condition.

Proposition 5.8.

The central element ZμZ_{\mu} acts on IλI_{\lambda} by 0 when |λ||μ||\lambda|\leq|\mu|, and λμ\lambda\neq\mu.

Proof.

By Lemma 2.1, DμD_{\mu} acts on ωIλ\omega\in I_{\lambda} by a scalar cμ(λ)c_{\mu}(\lambda) for any λ\lambda. On the other hand, the action of 𝔘𝔨\mathfrak{U}^{\mathfrak{k}} on a spherical vector v𝔨v^{\mathfrak{k}} in a 𝔤\mathfrak{g}-module descends to an action of 𝔇=𝔘𝔨/(𝔘𝔨)𝔨\mathfrak{D}=\mathfrak{U}^{\mathfrak{k}}/(\mathfrak{U}\mathfrak{k})^{\mathfrak{k}} by setting π(X).v𝔨:=X.v𝔨\pi(X).v^{\mathfrak{k}}:=X.v^{\mathfrak{k}}. Hence, in IλI_{\lambda}, there is no ambiguity of writing

(51) cμ(λ)ω=Dμ.ω=𝒟μ.ω=Zμ.ω.c_{\mu}(\lambda)\omega=D_{\mu}.\omega=\mathscr{D}_{\mu}.\omega=Z_{\mu}.\omega.

Therefore, by Corollary 5.7,

(52) cμ(λ)=0,if |λ||μ|,λμ.c_{\mu}(\lambda)=0,\;\text{if }|\lambda|\leq|\mu|,\lambda\neq\mu.

Since ZμZ_{\mu}\in\mathfrak{Z} acts on the entirety of a highest weight module by a scalar and IλI_{\lambda} is a highest weight module, such scalar has to be cμ(λ)c_{\mu}(\lambda) which vanishes according to Eq. (52). ∎

5.3. A stronger result

What is proved in the previous subsection turns out to be enough for the purposes of this paper (Theorem A). Nonetheless, we give a better description of such vanishing actions, proving Theorem C.

First, we recall the following “weight decomposition” proposition from [Zhu22], parallel to Proposition 3.2 in [Kum10], and is first due to Kostant [Kos59]. We let 𝔏\mathfrak{L} be a Lie (super)algebra, 𝔟𝔏\mathfrak{b}\subseteq\mathfrak{L} be a Borel subalgebra.

Proposition 5.9.

Let L1L_{1} and L2L_{2} be two finite dimensional irreducible 𝔏\mathfrak{L}-modules with highest weight κ1\kappa_{1} and κ2\kappa_{2}, and VV be a weight module of 𝔏\mathfrak{L}. If Hom𝔏(L1,L2V){0}\operatorname{Hom}_{\mathfrak{L}}\left(L_{1},L_{2}\otimes V\right)\neq\{0\}, then κ1=κ2+ν\kappa_{1}=\kappa_{2}+\nu for some weight ν\nu of VV.

We would like to show the following branching statement, i.e., WμW_{\mu} occurs in IλI_{\lambda} as a homomorphic image if μiλi\mu_{i}\leq\lambda_{i}. This result is similar to the one regarding the irreducible representation V(λ)V(\lambda^{\natural}) in [Zhu22].

Proposition 5.10.

If Hom𝔨(Wμ,Iλ){0}\operatorname{Hom}_{\mathfrak{k}}\left(W_{\mu},I_{\lambda}\right)\neq\{0\}, then λμ\lambda\supseteq\mu.

Proof.

The 𝔨\mathfrak{k}-module structure on IλI_{\lambda} is given by 𝔖(𝔭)Wλ\mathfrak{S}(\mathfrak{p}^{-})\otimes W_{\lambda}. Let us denote the representation map on 𝔖(𝔭)Wλ\mathfrak{S}(\mathfrak{p}^{-})\otimes W_{\lambda} (respectively Wλ𝔖(𝔭)W_{\lambda}\otimes\mathfrak{S}(\mathfrak{p}^{-})) as π\pi (respectively π\pi^{\prime}). Then the braiding map s:𝔖(𝔭)WλWλ𝔖(𝔭)s:\mathfrak{S}(\mathfrak{p}^{-})\otimes W_{\lambda}\rightarrow W_{\lambda}\otimes\mathfrak{S}(\mathfrak{p}^{-}) gives a canonical module isomorphism, meaning that if WμW_{\mu} occurs in IλI_{\lambda}, then it also occurs in Wλ𝔖(𝔭)W_{\lambda}\otimes\mathfrak{S}(\mathfrak{p}^{-}).

By Proposition 5.9, ν=μ𝔱λ𝔱\nu=\mu^{\natural}_{\mathfrak{t}}-\lambda^{\natural}_{\mathfrak{t}} is a weight of 𝔖(𝔭)\mathfrak{S}(\mathfrak{p}^{-}). Here, μ𝔱\mu^{\natural}_{\mathfrak{t}} and λ𝔱\lambda^{\natural}_{\mathfrak{t}} are weights in 𝔱\mathfrak{t}^{*}. As both μ𝔱\mu^{\natural}_{\mathfrak{t}} and λ𝔱\lambda^{\natural}_{\mathfrak{t}} are integral combinations of γΣ\gamma\in\Sigma_{\perp}, so is ν\nu. Since ΣΣ(𝔭0¯+,𝔱)\Sigma_{\perp}\subseteq\Sigma(\mathfrak{p}^{+}_{\overline{0}},\mathfrak{t}), such integral coefficients in ν\nu must be non-positive. Equivalently, λμ\lambda\supseteq\mu. ∎

Proof of Theorem C.

As in the proof of Proposition 5.8, cμ(λ)c_{\mu}(\lambda) is the scalar by which DμD_{\mu} acts on Iλ𝔨I_{\lambda}^{\mathfrak{k}}. Again by Eq. (51) and the fact that ZμZ_{\mu} has to act uniformly by a scalar on IλI_{\lambda}, cμ(λ)c_{\mu}(\lambda) is the scalar by which DμD_{\mu} acts on the entirety of IλI_{\lambda}. By Lemma 5.5, the representation of WμW_{\mu} on Iλ𝔨I_{\lambda}^{\mathfrak{k}} always gives a homomorphic image of itself in IλI_{\lambda}. If λμ\lambda\nsupseteq\mu, then this image is zero by Proposition 5.10. Spelling out DμD_{\mu} as in Eq. (49), we see that Dμ.Iλ𝔨Wμ.Iλ𝔨={0}D_{\mu}.I_{\lambda}^{\mathfrak{k}}\subseteq W_{\mu}.I_{\lambda}^{\mathfrak{k}}=\{0\}. Thus,

(53) cμ(λ)=0,if λμ,c_{\mu}(\lambda)=0,\;\text{if }\lambda\nsupseteq\mu,

and this proves that ZμZ_{\mu} acts on IλI_{\lambda} by 0 unless λiμi\lambda_{i}\geq\mu_{i} for all ii. ∎

Eq. (53) gives more vanishing points of cμc_{\mu} than we actually need for proving Theorem A in Section 6. It is stronger than Proposition 5.8 which concerns only zeros at lower degrees. In fact, this is much closer to the original formulation of the vanishing properties, introduced in Theorem 2.5, c.f. Proposition 6.1 below.

6. Type BC Interpolation Polynomials

In this section, we specify Sergeev and Veselov’s results (Subsection 2.4) with all the parameters in Table 1 (Subsection 3.2), and then an explicit change of variables. For the sake of simplicity, write \mathscr{H} for (p,q)\mathscr{H}(p,q), d\mathscr{H}_{d} for {λ:|λ|d}\{\lambda\in\mathscr{H}:|\lambda|\leq d\}. Throughout the section, we set 1ip1\leq i\leq p, 1jq1\leq j\leq q.

We now specify using our restricted root system Σ=Σ(𝔤,𝔞)\Sigma=\Sigma(\mathfrak{g},\mathfrak{a}). In particular, we set m:=q,n:=pm:=q,n:=p, ϵj:=αjF\epsilon_{j}:=\alpha^{\textsc{F}}_{j}, δi:=αiB\delta_{i}:=\alpha^{\textsc{B}}_{i}, zi:=xiz_{i}:=x_{i}, wj:=yjw_{j}:=y_{j}. By Table 1, 𝗄=1\mathsf{k}=-1, 𝗊=𝗌=1/2\mathsf{q}=\mathsf{s}=-1/2 while 𝗉\mathsf{p} and 𝗋\mathsf{r} do not exist. By comparing Eq. (10) with (1/2)ρ(-1/2)\rho as in Eq. (26), we have 𝗁=qp+1/2\mathsf{h}=q-p+1/2. Thus we have the two parameters

𝗄=1,𝗁=qp+1/2\mathsf{k}=-1,\quad\mathsf{h}=q-p+1/2

for Λq,p(1,qp+1/2)\Lambda^{(-1,q-p+1/2)}_{q,p}. Note our form Eq. (21) is the negative of what is defined in Subsection 2.4. But this does not matter when 𝗄=1\mathsf{k}=-1. We set (c.f. Eq. (9))

λ¯:=λiαiB+λjmαjF.\overline{\lambda^{\natural}}:=\sum\lambda_{i}\alpha^{\textsc{B}}_{i}+\sum\langle\lambda_{j}^{\prime}-m\rangle\alpha^{\textsc{F}}_{j}.

We write Λ𝟶\Lambda^{\mathtt{0}} for Λ𝟶(𝔞)\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}), and define Λd𝟶:={fΛ𝟶:degf2d}\Lambda^{\mathtt{0}}_{d}:=\{f\in\Lambda^{\mathtt{0}}:\deg f\leq 2d\}. We define 2(d)+ρ:={2λ¯+ρ:λd}𝔞2(\mathscr{H}_{d})^{\natural}+\rho:=\left\{2\overline{\lambda^{\natural}}+\rho:\lambda\in\mathscr{H}_{d}\right\}\subseteq\mathfrak{a}^{*}.

Proposition 6.1.

For each μ\mu\in\mathscr{H}, there is a unique polynomial JμΛ𝟶J_{\mu}\in\Lambda^{\mathtt{0}} of degree 2|μ|2|\mu| such that

Jμ(2λ¯+ρ)=0,for all λμ,λJ_{\mu}(2\overline{\lambda^{\natural}}+\rho)=0,\quad\text{for all }\lambda\nsupseteq\mu,\,\lambda\in\mathscr{H}

and that Jμ(2μ¯+ρ)=Cμ(1;1)Cμ+(2q2p;1)J_{\mu}(2\overline{\mu^{\natural}}+\rho)=C_{\mu}^{-}(1;-1)C_{\mu}^{+}(2q-2p;-1). Moreover, {Jμ:μd}\{J_{\mu}:\mu\in\mathscr{H}_{d}\} is a basis for Λd𝟶\Lambda^{\mathtt{0}}_{d}.

Proof.

Let τ:xi12(xiρiB)\tau:x_{i}\mapsto\frac{1}{2}(x_{i}-\rho_{i}^{\textsc{B}}) and yj12(yjρjF)y_{j}\mapsto\frac{1}{2}(y_{j}-\rho_{j}^{\textsc{F}}) be the change of variables which is an isomorphism from Λp,q(1,qp+1/2)\Lambda^{(-1,q-p+1/2)}_{p,q} to Λ𝟶(𝔞)\Lambda^{\mathtt{0}}(\mathfrak{a}^{*}) (c.f. [Zhu22, Proposition 5.9]). Consider Iμ(X;1,qp+12)Λp,q(1,qp+1/2)I_{\mu^{\prime}}(X;-1,q-p+\frac{1}{2})\in\Lambda^{(-1,q-p+1/2)}_{p,q}. Since (z(λ),w(λ))=(λ)(z(\lambda),w(\lambda))=(\lambda^{\prime})^{\natural}, by Theorem 2.5, we see that Jμ:=τ(Iμ(X;1,qp+12))J_{\mu}:=\tau(I_{\mu^{\prime}}(X;-1,q-p+\frac{1}{2})) satisfies the vanishing properties

(54) Jμ(2λ¯+ρ)=0,for all λμ,λ.J_{\mu}(2\overline{\lambda^{\natural}}+\rho)=0,\quad\text{for all }\lambda\nsupseteq\mu,\,\lambda\in\mathscr{H}.

The normalization condition is straightforward. As {Iμ:μd}\{I_{\mu}^{\prime}:\mu\in\mathscr{H}_{d}\} is a basis for {fΛp,q(1,qp+1/2):degf2d}\{f\in\Lambda^{(-1,q-p+1/2)}_{p,q}:\deg f\leq 2d\}, {Jμ:μd}\{J_{\mu}:\mu\in\mathscr{H}_{d}\} becomes a basis for Λd𝟶\Lambda^{\mathtt{0}}_{d}. ∎

Proposition 6.2.

Every fΛd𝟶f\in\Lambda^{\mathtt{0}}_{d} is determined by its values on 2(d)+ρ2(\mathscr{H}_{d})^{\natural}+\rho.

Proof.

Let 𝒱d\mathcal{V}_{d} be the space of functions on 2(d)+ρ2(\mathscr{H}_{d})^{\natural}+\rho. Then dimΛd𝟶=dim𝒱d=|d|\dim\Lambda^{\mathtt{0}}_{d}=\dim\mathcal{V}_{d}=|\mathscr{H}_{d}|. In particular, 𝒱d\mathcal{V}_{d} has a Kronecker-delta basis {δλ:δλ(2λ¯+ρ)=1,δλ(2μ¯+ρ)=0,λ,μd}\{\delta_{\lambda}:\delta_{\lambda}(2\overline{\lambda^{\natural}}+\rho)=1,\delta_{\lambda}(2\overline{\mu^{\natural}}+\rho)=0,\lambda,\mu\in\mathscr{H}_{d}\}. Next, the evaluation of fΛd𝟶f\in\Lambda^{\mathtt{0}}_{d} on d\mathscr{H}_{d} gives a restriction map

𝚛𝚎𝚜:Λd𝟶𝒱d.\mathtt{res}:\Lambda^{\mathtt{0}}_{d}\rightarrow\mathcal{V}_{d}.

To prove the statement, we show that 𝚛𝚎𝚜\mathtt{res} is an isomorphism.

Fix a total order \succ on d\mathscr{H}_{d} such that μλ\mu\succ\lambda implies |μ||λ||\mu|\geq|\lambda|. Consider the matrix RR for 𝚛𝚎𝚜\mathtt{res} with respect to the bases {Jμ}\{J_{\mu}\} for Λd𝟶\Lambda^{\mathtt{0}}_{d} and {δλ}\{\delta_{\lambda}\} for 𝒱d\mathcal{V}_{d} arranged by \succ. Since Jμ(2μ¯+ρ)0J_{\mu}(2\overline{\mu^{\natural}}+\rho)\neq 0, and Jμ(2λ¯+ρ)=0J_{\mu}(2\overline{\lambda^{\natural}}+\rho)=0 for any λ\lambda such that μλ\mu\succ\lambda, we see that RR is upper triangular with non-zero diagonal entries. Therefore RR is invertible, proving the statement. ∎

To prove Theorem A, we first show that γ𝟶(𝒟μ)\gamma^{\mathtt{0}}(\mathscr{D}_{\mu}) is proportional to JμJ_{\mu} defined above. Then we pin down the scalar

(55) kμ:=(1)|μ|Cμ(1;1).k_{\mu}:=(-1)^{|\mu|}C_{\mu}^{-}(1;-1).

by comparing the top homogeneous degrees of both sides. Here

Cμ(1;1)=(i,j)μ(μij+μji+1)C_{\mu}^{-}(1;-1)=\prod_{(i,j)\in\mu}\left(\mu_{i}-j+\mu^{\prime}_{j}-i+1\right)

(c.f. Subsection 2.4). The factor of (1)|μ|(-1)^{|\mu|} is due to the definition of dμd_{\mu}, which is a result of the definition of φT\varphi_{T} in [SV05, Eq. (41)].

Proof of Theorem A.

We show that

γ𝟶(𝒟μ)=kμJμ.\gamma^{\mathtt{0}}(\mathscr{D}_{\mu})=k_{\mu}J_{\mu}.

Step 1. From the definition of Γ{\Gamma} and γ𝟶\gamma^{\mathtt{0}}, we see that degγ𝟶(𝒟μ)2|μ|\deg\gamma^{\mathtt{0}}(\mathscr{D}_{\mu})\leq 2|\mu|. By (1), we have γ𝟶(π(Zμ))=𝚁𝚎𝚜(γ(Zμ))\gamma^{\mathtt{0}}(\pi(Z_{\mu}))=\operatorname{\mathtt{Res}}(\gamma(Z_{\mu})). The left side is just γ𝟶(𝒟μ)\gamma^{\mathtt{0}}(\mathscr{D}_{\mu}). The right side, when applied to 2λ¯+ρ2\overline{\lambda^{\natural}}+\rho is just γ(Zμ)(2λ𝔥+ρ𝔥)\gamma(Z_{\mu})(2\lambda^{\natural}_{\mathfrak{h}}+\rho_{\mathfrak{h}}) by Eq. (42) (c.f. Eq. (46)). Hence

(56) γ𝟶(𝒟μ)(2λ¯+ρ)=γ(Zμ)(2λ𝔥+ρ).\gamma^{\mathtt{0}}(\mathscr{D}_{\mu})(2\overline{\lambda^{\natural}}+\rho)=\gamma(Z_{\mu})(2\lambda^{\natural}_{\mathfrak{h}}+\rho).

By Theorem C, ZμZ_{\mu} acts by 0 on IλI_{\lambda} for all λμ\lambda\nsupseteq\mu, λ\lambda\in\mathscr{H}, which exactly gives γ(Zμ)(2λ𝔥+ρ)=0\gamma(Z_{\mu})(2\lambda^{\natural}_{\mathfrak{h}}+\rho)=0. Thus

γ𝟶(𝒟μ)(2λ¯+ρ)=0\gamma^{\mathtt{0}}(\mathscr{D}_{\mu})(2\overline{\lambda^{\natural}}+\rho)=0

for all λμ\lambda\nsupseteq\mu, λ\lambda\in\mathscr{H}, proving the vanishing properties.

Alternatively, by Proposition 5.8, we have the “relaxed” vanishing properties γ𝟶(𝒟μ)(2λ¯+ρ)=0\gamma^{\mathtt{0}}(\mathscr{D}_{\mu})(2\overline{\lambda^{\natural}}+\rho)=0 whenλμ\lambda\neq\mu, |λ||μ||\lambda|\leq|\mu|. By Proposition 6.2, this also implies the full “extra” vanishing properties.

Therefore, γ𝟶(𝒟μ)\gamma^{\mathtt{0}}(\mathscr{D}_{\mu}) is proportional to JμJ_{\mu} by its vanishing properties.

Step 2. We first pick a basis for 𝔭+\mathfrak{p}^{+} extended from

{𝗂Ei,p+i,𝗂Dj,q+j},\{\mathsf{i}E_{i,p+i},\mathsf{i}D_{j,q+j}\},

and a dual basis for 𝔭(𝔭+)\mathfrak{p}^{-}\cong(\mathfrak{p}^{+})^{*} (identified via the bilinear form bb) extended from

{2𝗂Ei,p+i,2𝗂Dj,q+j}.\{-2\mathsf{i}E_{i,p+i},2\mathsf{i}D_{j,q+j}\}.

Then the identity map 1𝔭+End𝔨(𝔭+)1_{\mathfrak{p}^{+}}\in\operatorname{End}_{\mathfrak{k}}(\mathfrak{p}^{+}) corresponds to

(57) i=1p𝗂Ei,p+i(2𝗂Ei,p+i)+j=1q𝗂Dj,p+j(2𝗂Dj,p+j)+other terms\sum_{i=1}^{p}\mathsf{i}E_{i,p+i}\otimes(-2\mathsf{i}E_{i,p+i})+\sum_{j=1}^{q}\mathsf{i}D_{j,p+j}\otimes(2\mathsf{i}D_{j,p+j})+\text{other terms}

in 𝔭𝔭+\mathfrak{p}^{-}\otimes\mathfrak{p}^{+}. On the other hand, we may extend the basis {xi,yj}\{x_{i},y_{j}\} for 𝔞\mathfrak{a} (see Subsection 3.1) to a basis for 𝔤\mathfrak{g} according to 𝔤=𝔫𝔞𝔨\mathfrak{g}=\mathfrak{n}^{-}\oplus\mathfrak{a}\oplus\mathfrak{k}. This is used to define Γ{\Gamma} and γ𝟶\gamma^{\mathtt{0}}. Then it is a direct computation to see that the top homogeneous degree of

Γ(D(1))=γ𝟶(𝒟(1)){\Gamma}(D_{(1)})=\gamma^{\mathtt{0}}(\mathscr{D}_{(1)})

is precisely

12i=1pxi212j=1qyj2.\frac{1}{2}\sum_{i=1}^{p}x_{i}^{2}-\frac{1}{2}\sum_{j=1}^{q}y_{j}^{2}.

Note the ρ\rho shift does not change this top degree.

Now we let μm\mu\in\mathscr{H}^{m}. The identity map 1μ1_{\mu} on WμW_{\mu} corresponds to an element in WμWμW_{\mu}^{*}\otimes W_{\mu}. Thus the sum μm1μ\sum_{\mu\in\mathscr{H}^{m}}1_{\mu} corresponds to an element in 𝔖m(𝔭)𝔖m(𝔭+)\mathfrak{S}^{m}(\mathfrak{p}^{-})\otimes\mathfrak{S}^{m}(\mathfrak{p}^{+}) by the decomposition in Proposition 3.1. On 𝔖m(𝔭)Tm(𝔭)\mathfrak{S}^{m}(\mathfrak{p})\subseteq T^{m}(\mathfrak{p}), the bilinear form is induced from bb on 𝔭\mathfrak{p} by normalizing bmb^{\otimes m} by 1/m!1/m!. Then similarly, we see that the top homogeneous degree of γ𝟶(μm𝒟μ)\gamma^{\mathtt{0}}\left(\sum_{\mu\in\mathscr{H}^{m}}\mathscr{D}_{\mu}\right) is given by

(58) 1m!(12i=1pxi212j=1qyj2)m=12mm!(p2(m,n)(xi,yj))m.\frac{1}{m!}\left(\frac{1}{2}\sum_{i=1}^{p}x_{i}^{2}-\frac{1}{2}\sum_{j=1}^{q}y_{j}^{2}\right)^{m}=\frac{1}{2^{m}m!}\left(p_{2}^{(m,n)}(x_{i},y_{j})\right)^{m}.

Note dλd_{\lambda} (Eq. (12)) specializes to (1)|λ|(-1)^{|\lambda|}. From Eq. (11), Theorems 2.6 and 2.7, we see that

IμSV(xi,yj;1,qp+12)=(1)|μ|I^μ(xi,yj;1,12(qp)).I^{\textup{SV}}_{\mu^{\prime}}\left(x_{i},y_{j};-1,q-p+\frac{1}{2}\right)=(-1)^{|\mu|}\hat{I}_{\mu}\left(x_{i},y_{j};-1,\frac{1}{2}-(q-p)\right).

and the top homogeneous degree of is (1)|μ|SPμ(xi2,yj2;1)(-1)^{|\mu|}SP_{\mu}(x_{i}^{2},y_{j}^{2};1). For JμJ_{\mu}, this becomes

(1/2)|μ|SPμ(xi2,yj2;1).(-1/2)^{|\mu|}SP_{\mu}(x_{i}^{2},y_{j}^{2};1).

Thus by setting θ=1\theta=1 in Eq. (16),

(59) 12mm!(i=1pxi2j=1qyj2)m=μm2mCμ(1;1)(2)|μ|((1/2)|μ|SPμ(xi2,yj2;1)).\frac{1}{2^{m}m!}\left(\sum_{i=1}^{p}x_{i}^{2}-\sum_{j=1}^{q}y_{j}^{2}\right)^{m}=\sum_{\mu\in\mathscr{H}^{m}}2^{-m}C_{\mu}^{-}(1;-1)(-2)^{|\mu|}\left((-1/2)^{|\mu|}SP_{\mu}(x_{i}^{2},y_{j}^{2};1)\right).

Comparing the coefficients, we have

(60) γ𝟶(𝒟μ)=(1)|μ|Cμ(1;1)Jμ.\gamma^{\mathtt{0}}(\mathscr{D}_{\mu})=(-1)^{|\mu|}C_{\mu}^{-}(1;-1)J_{\mu}.

Therefore the coefficient is precisely kμk_{\mu} given by Eq. (55). ∎

Additionally, we can answer the interesting question of by which scalar DμD_{\mu} acts on the spherical vector ωIμ𝔨\omega\in I_{\mu}^{\mathfrak{k}}, without any direct consideration on the module IμI_{\mu} itself!

Corollary 6.3.

The operator Dμ𝔘𝔨D_{\mu}\in\mathfrak{U}^{\mathfrak{k}} acts on Iμ𝔨I_{\mu}^{\mathfrak{k}} by (1)|μ|Cμ(1;1)2Cμ+(2q2p;1)(-1)^{|\mu|}C_{\mu}^{-}(1;-1)^{2}C_{\mu}^{+}(2q-2p;-1).

Proof.

This scalar is the value of Γ(Dμ)=γ𝟶(𝒟μ){\Gamma}(D_{\mu})=\gamma^{\mathtt{0}}(\mathscr{D}_{\mu}) at 2μ¯+ρ2\overline{\mu^{\natural}}+\rho which equals the value of kμIμSVk_{\mu}I^{\textup{SV}}_{\mu^{\prime}} at (z(μ),w(μ))(z(\mu^{\prime}),w(\mu^{\prime})):

γ𝟶(𝒟μ)(2μ¯+ρ)\displaystyle\gamma^{\mathtt{0}}(\mathscr{D}_{\mu})(2\overline{\mu^{\natural}}+\rho) =kμCμ(1;1)Cμ+(2q2p;1)\displaystyle=k_{\mu}C_{\mu^{\prime}}^{-}(1;-1)C_{\mu^{\prime}}^{+}(2q-2p;-1)
=(1)|μ|Cμ(1;1)2Cμ+(2q2p;1).\displaystyle=(-1)^{|\mu|}C_{\mu}^{-}(1;-1)^{2}C_{\mu}^{+}(2q-2p;-1).

References

  • [AHZ10] Alexander Alldridge, Joachim Hilgert, and Martin R. Zirnbauer, Chevalley’s restriction theorem for reductive symmetric superpairs, J. Algebra 323 (2010), no. 4, 1159–1185. MR 2578599
  • [All12] A. Alldridge, The Harish-Chandra isomorphism for reductive symmetric superpairs, Transform. Groups 17 (2012), no. 4, 889–919. MR 3000475
  • [AS15] Alexander Alldridge and Sebastian Schmittner, Spherical representations of Lie supergroups, J. Funct. Anal. 268 (2015), no. 6, 1403–1453. MR 3306354
  • [ASS18] Alexander Alldridge, Siddhartha Sahi, and Hadi Salmasian, Schur QQ-functions and the Capelli eigenvalue problem for the Lie superalgebra 𝔮(n)\mathfrak{q}(n), Representation theory and harmonic analysis on symmetric spaces, Contemp. Math., vol. 714, Amer. Math. Soc., [Providence], RI, [2018] ©2018, pp. 1–21. MR 3847241
  • [CW01] Shun-Jen Cheng and Weiqiang Wang, Howe duality for Lie superalgebras, Compositio Math. 128 (2001), no. 1, 55–94. MR 1847665
  • [CW12] by same author, Dualities and representations of Lie superalgebras, Graduate Studies in Mathematics, vol. 144, American Mathematical Society, Providence, RI, 2012. MR 3012224
  • [FK90] J. Faraut and A. Korányi, Function spaces and reproducing kernels on bounded symmetric domains, J. Funct. Anal. 88 (1990), no. 1, 64–89. MR 1033914
  • [Gor04] Maria Gorelik, The Kac construction of the centre of U(𝔤)U(\mathfrak{g}) for Lie superalgebras, J. Nonlinear Math. Phys. 11 (2004), no. 3, 325–349. MR 2084313
  • [Hel00] Sigurdur Helgason, Groups and geometric analysis, Mathematical Surveys and Monographs, vol. 83, American Mathematical Society, Providence, RI, 2000, Integral geometry, invariant differential operators, and spherical functions, Corrected reprint of the 1984 original. MR 1790156
  • [Hum78] James E. Humphreys, Introduction to Lie algebras and representation theory, Graduate Texts in Mathematics, vol. 9, Springer-Verlag, New York-Berlin, 1978, Second printing, revised. MR 499562
  • [Kac84] Victor G. Kac, Laplace operators of infinite-dimensional Lie algebras and theta functions, Proc. Nat. Acad. Sci. U.S.A. 81 (1984), no. 2, 645–647. MR 735060
  • [Kos59] Bertram Kostant, A formula for the multiplicity of a weight, Trans. Amer. Math. Soc. 93 (1959), 53–73. MR 109192
  • [KS93] Bertram Kostant and Siddhartha Sahi, Jordan algebras and Capelli identities, Invent. Math. 112 (1993), no. 3, 657–664. MR 1218328
  • [Kum10] Shrawan Kumar, Tensor product decomposition, Proceedings of the International Congress of Mathematicians. Volume III, Hindustan Book Agency, New Delhi, 2010, pp. 1226–1261. MR 2827839
  • [LSS22a] Gail Letzter, Siddhartha Sahi, and Hadi Salmasian, The Capelli eigenvalue problem for quantum groups, 2022, arXiv.
  • [LSS22b] by same author, Weyl algebras for quantum homogeneous spaces, 2022, arXiv.
  • [LSS23] by same author, Quantized Weyl algebras, the double centralizer property, and a new First Fundamental Theorem for Uq(𝔤𝔩n)U_{q}(\mathfrak{gl}_{n}), 2023, arXiv.
  • [Mac95] I. G. Macdonald, Symmetric functions and Hall polynomials, second ed., Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1995, With contributions by A. Zelevinsky, Oxford Science Publications. MR 1354144
  • [Mus12] Ian M. Musson, Lie superalgebras and enveloping algebras, Graduate Studies in Mathematics, vol. 131, American Mathematical Society, Providence, RI, 2012. MR 2906817
  • [Oko98] A. Okounkov, BC{\rm BC}-type interpolation Macdonald polynomials and binomial formula for Koornwinder polynomials, Transform. Groups 3 (1998), no. 2, 181–207. MR 1628453
  • [OO06] Andrei Okounkov and Grigori Olshanski, Limits of BCBC-type orthogonal polynomials as the number of variables goes to infinity, Jack, Hall-Littlewood and Macdonald polynomials, Contemp. Math., vol. 417, Amer. Math. Soc., Providence, RI, 2006, pp. 281–318. MR 2284134
  • [Sah94] Siddhartha Sahi, The spectrum of certain invariant differential operators associated to a Hermitian symmetric space, Lie theory and geometry, Progr. Math., vol. 123, Birkhäuser Boston, Boston, MA, 1994, pp. 569–576. MR 1327549
  • [Sch70] Wilfried Schmid, Die Randwerte holomorpher Funktionen auf hermitesch symmetrischen Räumen, Invent. Math. 9 (1969/70), 61–80. MR 259164
  • [Ser99] Alexander Sergeev, The invariant polynomials on simple Lie superalgebras, Represent. Theory 3 (1999), 250–280. MR 1714627
  • [She22] Alexander Sherman, Iwasawa decomposition for Lie superalgebras, J. Lie Theory 32 (2022), no. 4, 973–996. MR 4512155
  • [Shi90] Goro Shimura, Invariant differential operators on Hermitian symmetric spaces, Ann. of Math. (2) 132 (1990), no. 2, 237–272. MR 1070598
  • [SS16] Siddhartha Sahi and Hadi Salmasian, The Capelli problem for 𝔤𝔩(m|n)\mathfrak{gl}(m|n) and the spectrum of invariant differential operators, Adv. Math. 303 (2016), 1–38. MR 3552519
  • [SSS20] Siddhartha Sahi, Hadi Salmasian, and Vera Serganova, The Capelli eigenvalue problem for Lie superalgebras, Math. Z. 294 (2020), no. 1-2, 359–395. MR 4054814
  • [Sta89] Richard P. Stanley, Some combinatorial properties of Jack symmetric functions, Adv. Math. 77 (1989), no. 1, 76–115. MR 1014073
  • [Ste85] John R. Stembridge, A characterization of supersymmetric polynomials, J. Algebra 95 (1985), no. 2, 439–444. MR 801279
  • [SV05] A. N. Sergeev and A. P. Veselov, Generalised discriminants, deformed Calogero-Moser-Sutherland operators and super-Jack polynomials, Adv. Math. 192 (2005), no. 2, 341–375. MR 2128703
  • [SV09] Alexander N. Sergeev and Alexander P. Veselov, BCBC_{\infty} Calogero-Moser operator and super Jacobi polynomials, Adv. Math. 222 (2009), no. 5, 1687–1726. MR 2555909
  • [SV11] by same author, Grothendieck rings of basic classical Lie superalgebras, Ann. of Math. (2) 173 (2011), no. 2, 663–703. MR 2776360
  • [SZ19] Siddhartha Sahi and Genkai Zhang, Positivity of Shimura operators, Math. Res. Lett. 26 (2019), no. 2, 587–626. MR 3999556
  • [Yan92] Zhi Min Yan, A class of generalized hypergeometric functions in several variables, Canad. J. Math. 44 (1992), no. 6, 1317–1338. MR 1192421
  • [Zhu22] Songhao Zhu, Shimura operators for certain Hermitian symmetric superpairs, arXiv, submitted, 2022.