Hostname: page-component-cb9f654ff-w5vf4 Total loading time: 0 Render date: 2025-08-04T10:35:49.852Z Has data issue: false hasContentIssue false

Quasi-polynomial representations of double affine Hecke algebras

Published online by Cambridge University Press:  16 April 2025

Siddhartha Sahi
Affiliation:
Department of Mathematics, Rutgers University, 110 Frelinghhuysen Rd, Piscataway, NJ 08854-8019, USA; E-mail: sahi@math.rutgers.edu
Jasper Stokman*
Affiliation:
KdV Institute for Mathematics, University of Amsterdam, Science Park 105-107, 1098 XG, Amsterdam, The Netherlands
Vidya Venkateswaran
Affiliation:
Center for Communications Research, 805 Bunn Dr, Princeton, NJ 08540, USA; E-mail: vvenkat@idaccr.org
*
E-mail: j.v.stokman@uva.nl (corresponding author)

Abstract

We introduce an explicit family of representations of the double affine Hecke algebra $\mathbb {H}$ acting on spaces of quasi-polynomials, defined in terms of truncated Demazure-Lusztig type operators. We show that these quasi-polynomial representations provide concrete realisations of a natural family of cyclic Y-parabolically induced $\mathbb {H}$-representations. We recover Cherednik’s well-known polynomial representation as a special case.

The quasi-polynomial representation gives rise to a family of commuting operators acting on spaces of quasi-polynomials. These generalize the Cherednik operators, which are fundamental in the study of Macdonald polynomials. We provide a detailed study of their joint eigenfunctions, which may be regarded as quasi-polynomial, multi-parametric generalisations of nonsymmetric Macdonald polynomials. We also introduce generalizations of symmetric Macdonald polynomials, which are invariant under a multi-parametric generalization of the standard Weyl group action.

We connect our results to the representation theory of metaplectic covers of reductive groups over non-archimedean local fields. We introduce root system generalizations of the metaplectic polynomials from our previous work by taking a suitable restriction and reparametrization of the quasi-polynomial generalizations of Macdonald polynomials. We show that metaplectic Iwahori-Whittaker functions can be recovered by taking the Whittaker limit of these metaplectic polynomials.

Information

Type
Algebra
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1 Introduction

Nonsymmetric Macdonald polynomials [Reference Macdonald29, Reference Cherednik8, Reference Sahi37, Reference Haiman19, Reference Stokman, Koornwinder and Stokman44] are a remarkable family of functions that depend on several parameters – usually two or three, but as many as six for the Koornwinder setting. For suitable choices of parameters, they specialise to many important families of special functions arising in the representation theory of reductive groups, including spherical functions and Iwahori-Whittaker functions. In turn, Macdonald polynomials can be understood in terms of the representation theory of Cherednik’s double affine Hecke algebra $\mathbb {H}$ . More precisely, they are simultaneous eigenfunctions of a certain commutative subalgebra $\mathcal {P}_Y\subset \mathbb {H}$ in Cherednik’s polynomial representation of $\mathbb {H}$ on the group algebra $\mathcal {P}$ of a coroot lattice $Q^{\vee }$ .

In this paper, we discuss a generalisation of Macdonald polynomials, which can be understood via a family of representations $\pi _{c,\mathfrak {t}}$ of $\mathbb {H}$ on the subspace $\mathcal {P}^{(c)}$ of the group algebra of $Q^{\vee }\otimes _{\mathbb Z} {\mathbb R}$ generated by an affine Weyl group orbit $\mathcal {O}_c$ in $Q^{\vee }\otimes _{\mathbb {Z}}\mathbb {R}$ . Here, c is the representative of the orbit that lies in the fundamental alcove, and $\mathfrak {t}$ represents a number of additional representation parameters.

We refer to elements in the representation space as quasi-polynomials, and to $\pi _{c,\mathfrak {t}}$ as the quasi-polynomial representation. The simultaneous eigenfunctions of the commuting operators $\pi _{c,\mathfrak {t}}(h)$ ( $h\in \mathcal {P}_Y$ ) provide our quasi-polynomial generalisations of Macdonald polynomials. They depend on the usual Macdonald parameters as well as on the additional representation parameters $\mathfrak {t}$ .

We establish an explicit connection to the representation theory of metaplectic covers of reductive groups over non-archimedean local fields and the associated metaplectic Whittaker functions [Reference Chinta and Offen12, Reference McNamara30, Reference Chinta, Gunnells and Puskas15, Reference Patnaik and Puskas34, Reference Patnaik and Puskas35]. In our previous work [Reference Sahi, Stokman and Venkateswaran38], we introduced a metaplectic affine Hecke algebra action via explicit formulas and provided a direct construction as an induced representation. We showed that the Chinta-Gunnells [Reference Chinta and Gunnells13, Reference Chinta and Gunnells14] action of the Weyl group, originally used to construct Weyl group multiple Dirichlet series, can be recovered from this affine Hecke algebra action through localization. We also related this affine Hecke algebra action to the metaplectic Demazure-Lusztig operators from [Reference Patnaik and Puskas34, Reference Chinta, Gunnells and Puskas15, Reference Patnaik and Puskas35]. This is related to the context of the present paper as follows. If the orbit $\mathcal {O}_c$ is contained in the rational vector space $Q^{\vee }\otimes _{\mathbb {Z}}\mathbb {Q}$ , then the quasi-polynomials in $\mathcal {P}^{(c)}$ can be regarded as polynomials under an appropriate reparametrization. Moreover, under a suitable transformation, the restriction of the representation $\pi _{c, \mathfrak {t}}$ to the affine Hecke algebra generated by $T_1, \cdots , T_r$ and $\mathcal {P}$ recovers the metaplectic affine Hecke algebra representation of [Reference Sahi, Stokman and Venkateswaran38]. Under these identifications, the quasi-polynomial generalizations of Macdonald polynomials extend the metaplectic polynomials introduced in [Reference Sahi, Stokman and Venkateswaran38], and studi ed further in [Reference Saied39], from type A to an arbitrary root system. We also recover the metaplectic Iwahori-Whittaker and spherical Whittaker functions [Reference Patnaik and Puskas34, Reference McNamara30] as Whittaker limits of these metaplectic polynomials and their antisymmetric versions.

In the following six subsections, we provide an overview of our main results on the quasi-polynomial representations and associated generalisations of the Macdonald polynomials relative to adjoint root datum.

1.1 Double affine Weyl group actions on quasi-polynomials

To set the stage, it is instructive to first introduce the quasi-polynomial representations of the double affine Weyl group.

Let $\Phi _0$ be a reduced irreducible root system, realised within the linear dual of a Euclidean space E, and normalised so that long roots have squared length $2/m^2$ for some natural number $m\in \mathbb {Z}_{>0}$ . Let $\Phi =\Phi _0\times \mathbb {Z}$ be the associated irreducible affine root system of untwisted type. The coroot lattice $\widehat {Q}^{\vee }$ of $\Phi $ is isomorphic to $Q^{\vee }+m^2{\mathbb Z} K$ , where $Q^{\vee }$ is the coroot lattice of $\Phi _0$ and K is the central element in the associated untwisted affine Lie algebra. The affine Weyl group W of $\Phi $ is a semidirect product $W=W_0\ltimes Q^{\vee }$ , where $W_0$ is the Weyl group of $\Phi _0$ . The double affine Weyl group is the semidirect product

(1.1) $$ \begin{align} \mathbb{W}=W\ltimes \widehat{Q}^{\vee}=(W_0\ltimes Q^{\vee})\ltimes \widehat{Q}^{\vee}, \\[-24pt]\nonumber\end{align} $$

where the W-action $(w,\widehat {\mu })\mapsto w\cdot \widehat {\mu }$ on $\widehat {Q}^{\vee }$ is determined by the pairing of $Q^{\vee }$ into the center $m^2{\mathbb Z} K$ of $\mathbb {W}$ induced by the inner product $\langle \cdot ,\cdot \rangle $ on E.

We fix simple roots $\alpha _0,\alpha _1,\ldots , \alpha _r$ for $\Phi $ such that $\alpha _1, \ldots , \alpha _r$ are simple roots for $\Phi _0$ . Let $\alpha _j^{\vee }$ be the corresponding simple coroots. Then $\alpha _0=(-\varphi ,1)$ , where $\varphi $ is the highest root for $\Phi _0$ , and $\alpha _0^{\vee }=m^2K-\varphi ^{\vee }$ . The simple reflection associated to $\alpha _j$ is denoted by $s_j\in W$ . The affine root hyperplane configuration in E provides a decomposition of E in closed alcoves, each closed alcove being a fundamental domain for the action of W on E by reflections and translations. The choice of simple roots singles out the fundamental open alcove $C_+$ . We write $\mathcal {O}_c$ for the W-orbit in E containing $c\in \overline {C}_+$ .

The face decomposition of $\overline {C}_+$ is the disjoint union

$$\begin{align*}\overline{C}_+=\bigsqcup_{J\subsetneq \{0,\ldots,r\}}C^J \end{align*}$$

with $C^J$ the vectors c in $\overline {C}_+$ satisfying $\alpha _j(c)=0$ if and only if $j\in J$ .

Let $\mathbf {F}$ be a field of characteristic zero and fix $q\in \mathbf {F}^\times $ not a root of unity. Let $T:=\text {Hom}(Q^{\vee },\mathbf {F}^\times )$ be the $\mathbf {F}$ -torus consisting of multiplicative characters $\mathfrak {t}: Q^{\vee }\rightarrow \mathbf {F}^\times $ . We will denote the value of $\mathfrak {t}$ at $\mu \in Q^{\vee }$ by $\mathfrak {t}^\mu $ . Consider the associated affine subtorus

$$\begin{align*}T_J:=\{\mathfrak{t}\in T \,\, | \,\, \mathfrak{t}^{\alpha_j^{\vee}}=1 \,\,\, \forall\, j\in J\} \end{align*}$$

of dimension $r-\#J$ , where $\mathfrak {t}^{\mu +\ell K}:=q^\ell \mathfrak {t}^\mu $ . The field $\mathbf {F}$ should contain a sufficiently large root of q to ensure that $T_J\not =\emptyset $ for subsets J containing $0$ .

Denote by

$$\begin{align*}\mathbf{F}[E]=\bigoplus_{y\in E}\mathbf{F}x^y \end{align*}$$

the group algebra of E, viewed as additive group. It splits up as the direct sum of the $\mathcal {P}$ -submodules $\mathcal {P}^{(c)}:=\bigoplus _{y\in \mathcal {O}_c}\mathbf {F}x^y$ ( $c\in \overline {C}_+$ ).

For $c\in C^J$ , the space $\mathcal {P}^{(c)}$ of quasi-polynomials admits a natural family of $\mathbb {W}$ -actions $\cdot _{\mathfrak {t}}$ depending on q and $\mathfrak {t}\in T_J$ . The action of $W_0\ltimes \widehat {Q}^{\vee }$ is by reflections and translations of the exponents of the quasi-monomials $x^y$ , with $m^2K$ acting as multiplication by $q^{m^2}$ . In particular, this part of the action is independent of $\mathfrak {t}$ . The action of the standard abelian subgroup $\{\tau (\mu )\}_{\mu \in Q^{\vee }}$ in $W=W_0\ltimes Q^{\vee }$ is defined by

(1.2) $$ \begin{align} \tau(\mu)_{\mathfrak{t}}x^y:=\mathfrak{t}_y^{-\mu}x^y\qquad\quad (\mu\in Q^{\vee},\, y\in\mathcal{O}_c), \end{align} $$

where $\mathfrak {t}_y\in T$ is the multiplicative character $\mu \mapsto \mathfrak {t}^{w_y^{-1}\cdot \mu }$ with $w_y\in W$ the element of shortest length such that $w_yc=y$ (actually, one may take here any $w\in W$ satisfying $wc=y$ ). Note that $s_{j,\mathfrak {t}}x^y:=(s_j)_{\mathfrak {t}}x^y$ for $y\in \mathcal {O}_c$ is explicitly given by

$$\begin{align*}s_{i,\mathfrak{t}}x^y=x^{s_iy}\quad (1\leq i\leq r),\qquad\quad s_{0,\mathfrak{t}}x^y=\mathfrak{t}_y^{\varphi^{\vee}}x^{s_\varphi y} \end{align*}$$

with $s_\varphi \in W_0$ the reflection associated to the highest root $\varphi \in \Phi _0$ .

In case $J=\{1,\ldots ,r\}$ , $c=0$ and $\mathfrak {t}=1_T$ , we reobtain the standard action of $\mathbb {W}$ on $\mathcal {P}$ by polynomial q-difference reflection operators, since

$$\begin{align*}\tau(\mu)_{1_T}(x^\nu)=q^{-\langle\mu,\nu\rangle}x^\nu\qquad\quad (\mu,\nu\in Q^{\vee}). \end{align*}$$

In this case, we write $w(x^\mu )$ for $w_{1_T}x^\mu $ ( $w\in W$ ).

In this paper, we deform the $\mathbb {W}$ -action $\cdot _{\mathfrak {t}}$ on $\mathcal {P}^{(c)}$ to an action of the double affine Hecke algebra $\mathbb {H}$ . The resulting quasi-polynomial representation of $\mathbb {H}$ plays a central role in this paper.

1.2 The quasi-polynomial representation

Cherednik’s [Reference Cherednik8] double affine Hecke algebra (DAHA) is a certain flat deformation $\mathbb {H}=\mathbb {H}(\mathbf {k},q)$ of $\mathbf {F}[\mathbb {W}]/(m^2K-q^{m^2})\simeq W\ltimes \mathcal {P}$ depending on a multiplicity function $\mathbf {k}=(k_a)_{a\in \Phi }$ . We consider in this paper the case that $\mathbf {k}$ is invariant for the extended affine Weyl group. This means that $k_a\in \mathbf {F}^\times $ only depends on the length $\|\alpha \|$ of the gradient $\alpha \in \Phi _0$ of the affine root $a=(\alpha ,\ell )\in \Phi $ . In particular, there are at most two distinct values of the $k_a$ . We write $k_j=k_{\alpha _j}$ .

The double affine Hecke algebra $\mathbb {H}$ has an explicit definition in terms of generators and relations (see Definition 2.12). The generators are $T_j$ ( $0\leq j\leq r$ ) and $x^\mu $ ( $\mu \in Q^{\vee }$ ). The subalgebra generated by $T_j$ ( $0\leq j\leq r$ ) is the affine Hecke algebra $H=H(\mathbf {k})$ of W. Its defining relations are the $(W,\{s_0,\ldots ,s_r\})$ -braid relations and the Hecke relations $(T_j-k_j)(T_j+k_j^{-1})=0$ . The subalgebra generated by $x^\mu $ ( $\mu \in Q^{\vee }$ ) is $\mathcal {P}$ . The commutation relations between the $T_j$ and $x^\mu $ are $\mathbf {k}$ -deformations of $s_jx^\mu =s_{j}(x^\mu )s_j$ in $W\ltimes \mathcal {P}$ , given explicitly by

(1.3) $$ \begin{align} T_jx^\mu-s_{j}(x^\mu)T_j=(k_j-k_j^{-1})\left(\frac{x^\mu-s_{j}(x^\mu)}{1-x^{\alpha_j^{\vee}}}\right). \end{align} $$

Note here that $x^{\alpha _0^{\vee }}=x^{m^2K^2-\varphi ^{\vee }}=q^{m^2}x^{-\varphi ^{\vee }}$ . The right-hand side can be alternatively written as

$$\begin{align*}(k_j-k_j^{-1})\left(\frac{1-x^{-D\alpha_j(\mu)\alpha_j^{\vee}}}{1-x^{\alpha_j^{\vee}}}\right)x^\mu \end{align*}$$

with $D\alpha _j\in \Phi _0$ the gradient of $\alpha _j$ . In particular, (1.3) makes sense in $\mathbb {H}$ , since $D\alpha _j(\mu )\in \mathbb {Z}$ . Together with the well-known fact that $H\otimes \mathcal {P}\simeq \mathbb {H}$ as vector spaces by multiplication (Cherednik’s [Reference Cherednik8] PBW theorem for $\mathbb {H}$ ), this gives an explicit generalisation of the first isomorphism in (1.1).

Also, the decomposition $W\simeq W_0\ltimes Q^{\vee }$ has a natural analog in H, known as the Bernstein-Zelevinsky decomposition of H (see [Reference Lusztig25]). It provides a subalgebra $\mathcal {P}_Y=\bigoplus _{\mu \in Q^{\vee }}\mathbf {F}Y^\mu $ of H, isomorphic to $\mathcal {P}$ , such that $H_0\otimes \mathcal {P}_Y\simeq H$ as vector spaces, with $H_0$ the subalgebra of $\mathbb {H}$ generated by $T_i$ ( $1\leq i\leq r$ ). The duality anti-involution $\delta $ of $\mathbb {H}$ interchanges the two copies of $\mathcal {P}$ inside $\mathbb {H}$ . Concretely, it satisfies $\delta (Y^\mu )=x^{-\mu }$ and $\delta (T_i)=T_i$ for $\mu \in Q^{\vee }$ and $1\leq i\leq r$ .

Let $\chi _{B}: \mathbb {R}\rightarrow \{0,1\}$ for a subset $B\subset \mathbb {R}$ be the characteristic function of B, and let $\lfloor \cdot \rfloor : \mathbb {R}\rightarrow \mathbb {Z}$ be the floor-function (i.e., $\lfloor d\rfloor $ for $d\in \mathbb {R}$ is the largest integer $\leq d$ ). For $0\leq j\leq r$ , the truncated divided-difference operator $\nabla _j$ is the linear operator on $\mathbf {F}[E]$ defined by

$$\begin{align*}\nabla_j(x^y):=\left(\frac{1-x^{-\lfloor D\alpha_j(y)\rfloor\alpha_j^{\vee}}}{1-x^{\alpha_j^{\vee}}}\right)x^y\qquad\quad (y\in E). \end{align*}$$

Its restriction to $\mathcal {P}$ is the usual divided-difference operator.

The first main result of this paper is the following DAHA analog of the $\mathbb {W}$ -module $(\mathcal {P}^{(c)},\cdot _{\mathfrak {t}})$ ( $c\in C^J$ , $\mathfrak {t}\in T_J$ ), in which $T_j$ acts by a Demazure-Lusztig type operator involving the truncated divided-difference operator $\nabla _j$ . We call $\pi _{c,\mathfrak {t}}$ the quasi-polynomial representation of $\mathbb {H}$ .

Theorem 1.1. Let $J\subsetneq \{0,\ldots ,r\}$ , $c\in C^J$ and $\mathfrak {t}\in T_J$ . The formulas

$$ \begin{align*} \begin{aligned} \pi_{c,\mathfrak{t}}(T_j)x^y&:=k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}s_{j,\mathfrak{t}}x^y+(k_j-k_j^{-1}) \nabla_j(x^y),\\ \pi_{c,\mathfrak{t}}(x^\mu)x^y&:=x^{y+\mu} \end{aligned} \end{align*} $$

for $0\leq j\leq r$ , $\mu \in Q^{\vee }$ and $y\in \mathcal {O}_c$ define a representation $\pi _{c,\mathfrak {t}}: \mathbb {H}\rightarrow \text {End}(\mathcal {P}^{(c)})$ .

The representation $\pi :=\pi _{0,1_T}: \mathbb {H}\rightarrow \text {End}(\mathcal {P})$ is Cherednik’s [Reference Cherednik8] polynomial representation of $\mathbb {H}$ . In this case, $J=\{1,\ldots ,r\}$ , and the commuting operators $\pi (Y^\mu )$ ( $\mu \in Q^{\vee }$ ) are the Cherednik operators. The polynomial representation $\pi $ can alternatively be obtained by inducing the trivial one-dimensional H-representation to $\mathbb {H}$ . As a second main result of this paper we extend this result to the quasi-polynomial representation $\pi _{c,\mathfrak {t}}$ ( $c\in C^J$ ) by inducing a one-dimensional representation of the Y-parabolic subalgebra $\mathbb {H}_J^Y$ of $\mathbb {H}$ generated by $\delta (T_j)$ ( $j\in J$ ) and $\mathcal {P}_Y$ .

The relevant one-dimensional $\mathbb {H}_J^Y$ -representations are parametrised by the affine subtorus

$$\begin{align*}L_J:=\{\mathfrak{t}\in T \,\, | \,\, \mathfrak{t}^{\alpha_j^{\vee}}=k_j^{-2}\quad \forall\, j\in J\} \end{align*}$$

of T. The one-dimensional $\mathbb {H}_J^Y$ -representation associated to $t\in L_J$ is then defined by

$$\begin{align*}\chi_{J,t}(Y^\mu):=t^{-\mu}\quad (\mu\in Q^{\vee}),\qquad \chi_{J,t}(\delta(T_j)):=k_j\quad (j\in J). \end{align*}$$

We write $\mathbb {M}_{J,t}$ for the resulting induced $\mathbb {H}$ -module, and $m_{J,t}$ for its canonical cyclic vector.

Consider

$$\begin{align*}\mathfrak{s}_y:=\prod_{\alpha\in\Phi_0^+}k_\alpha^{\eta(\alpha(y))\alpha}\in T\qquad (y\in E), \end{align*}$$

with $\eta :=\chi _{\mathbb {Z}_{>0}}-\chi _{\mathbb {Z}_{\leq 0}}: \mathbb {R}\rightarrow \{-1,0,1\}$ (a discrete analog of the Heaviside function). The map $y\mapsto \mathfrak {s}_y$ is constant on faces, and we write $\mathfrak {s}_J\in T$ for its value on $C^J$ . We then have $L_J=\mathfrak {s}_JT_J$ and

Theorem 1.2. Let $c\in C^J$ and $\mathfrak {t}\in T_J$ . Then $\mathbb {M}_{J,\mathfrak {s}_J\mathfrak {t}}\simeq (\mathcal {P}^{(c)},\pi _{c,\mathfrak {t}})$ as $\mathbb {H}$ -modules, with the isomorphism determined by $m_{J,\mathfrak {s}_J\mathfrak {t}}\mapsto x^c$ .

In particular, $\pi _{c,\mathfrak {t}}\simeq \pi _{c^\prime ,\mathfrak {t}}$ if c and $c^\prime $ lie in the same face of $\overline {C}_+$ .

Quite remarkably, for all $\mu \in Q^{\vee }$ and $v\in W_0$ , the vector $x^\mu T_vm_{J,\mathfrak {s}_J\mathfrak {t}}\in \mathbb {M}_{J,\mathfrak {s}_J\mathfrak {t}}$ is mapped to an explicit constant multiple of the quasi-monomial $x^{\mu +vc}$ (see Theorem 4.5(3)). Actually, it is this formula which forms the starting point of the proof of both Theorem 1.1 and Theorem 1.2.

1.3 Quasi-polynomial eigenfunctions

The quasi-polynomial representations $\pi _{c,\mathfrak {t}}$ give rise to a family of quasi-polynomials, which generalize the nonsymmetric Macdonald polynomials. These quasi-polynomials enjoy many properties paralleling the classical case; for example, they are joint eigenfunctions of $\pi _{c,\mathfrak {t}}(Y^{\mu })$ for $\mu \in Q^{\vee }$ , satisfy a triangularity condition, and can be constructed explicitly via intertwiners. Their definition is as follows.

Let $c\in C^J$ . Define a partial order on $\mathcal {O}_c$ by

$$\begin{align*}y\leq y^\prime\quad \Leftrightarrow \quad w_y\leq_B w_{y^\prime}, \end{align*}$$

with $\leq _B$ the Bruhat order on $(W,\{s_0,\ldots ,s_r\})$ . Denote by $T_J^\prime \subseteq T_J$ the set of characters $\mathfrak {t}\in T_J$ for which the $\mathfrak {s}_y\mathfrak {t}_y$ ( $y\in \mathcal {O}_c$ ) are pairwise different in T (it does not depend on the choice of $c\in C^J$ ).

Theorem 1.3. Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ . For each $y\in \mathcal {O}_c$ , there exists a unique joint eigenfunction $E_y^J(x;\mathfrak {t})\in \mathcal {P}^{(c)}$ of the commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) satisfying

(1.4) $$ \begin{align} E_y^J(x;\mathfrak{t})=x^y+\sum_{y^\prime<y}e_{y,y^\prime;\mathfrak{t}}^Jx^{y^\prime}\qquad\quad (e_{y,y^\prime;\mathfrak{t}}^J\in\mathbf{F}). \end{align} $$

Furthermore,

(1.5) $$ \begin{align} \pi_{c,\mathfrak{t}}(Y^\mu)E_y^J(x;\mathfrak{t})=(\mathfrak{s}_y\mathfrak{t}_y)^{-\mu}E_y^J(x;\mathfrak{t})\qquad\quad \forall\, \mu\in Q^{\vee}. \end{align} $$

In particular, $\{E_y^J(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ is a basis of $\mathcal {P}^{(c)}$ consisting of joint eigenfunctions of the commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ). Furthermore, the quasi-polynomial eigenfunction of smallest degree is trivial,

$$\begin{align*}E_c^J(x;\mathfrak{t})=x^c. \end{align*}$$

The choice of the vector $c\in C^J$ only influences the quasi-exponents in the quasi-monomial expansion of the $E_y^J(x;\mathfrak {t})$ . In other words, the coefficients

$$\begin{align*}e_{w,w^\prime;\mathfrak{t}}^J:=e_{wc,w^\prime c;\mathfrak{t}}^J\qquad\quad (w,w^\prime\in W) \end{align*}$$

are independent of $c\in C^J$ (see Theorem 6.3). Furthermore, if $J^\prime \supseteq J$ , then $C^{J^\prime }$ lies in the closure of $C^J$ , and the quasi-polynomial $E^{J^\prime }_{y^\prime }(x;\mathfrak {t})$ for $\mathfrak {t}\in T_{J^\prime }^\prime $ can be recovered from $E^J_y(x;\mathfrak {s}_J^{-1}\mathfrak {s}_{J^\prime }\mathfrak {t})$ (see Proposition 6.16).

Example 1.4. Consider $\mathbf {F}=\mathbb {Q}(q,\mathbf {k},z_1,\ldots ,z_r)$ with indeterminates $q,k_a,z_1,\ldots ,z_r$ , where $k_a=k_b$ if $\|Da\|=\|Db\|$ . For $I\subseteq \{1,\ldots ,r\}$ , let $I^{\text {co}}$ be its complement. Then

$$\begin{align*}\prod_{i\in I^{\text{co}}}z_i^{\varpi_i}\in T_I^\prime, \end{align*}$$

where $\varpi _i$ denotes the $i^{\text {th}}$ fundamental weight of $\Phi _0$ . Then

$$\begin{align*}E_y^I(x;z_1,\ldots,z_r):=E_y^I\left(x;\prod_{i \in I^{\text{co}}}z_i^{\varpi_i}\right)\qquad\quad (y\in\mathcal{O}_c) \end{align*}$$

depend on the indeterminates $q, \mathbf {k}$ and $z_i$ ( $i\in I^{\text {co}}$ ). Quasi-polynomial eigenfunctions over $\mathbb {Q}(q,\mathbf {k})$ can be obtained by specialisation of the $z_i$ ’s.

The situation is a bit more intricate when $0\in J$ since the $z_i$ ’s will no longer be independent; see Subsection 6.1.

If $1_T\in T_{\{1,\ldots ,r\}}^\prime $ (which imposes generic conditions on q and $\mathbf {k}$ ), then

$$\begin{align*}E_\mu(x):=E_\mu^{\{1,\ldots,r\}}(x;1_T)\in\mathcal{P} \end{align*}$$

is the monic nonsymmetric Macdonald polynomial of degree $\mu \in Q^{\vee }$ . The Macdonald polynomials and their normalised and (anti)symmetric variants have remarkable properties. They satisfy orthogonality relations with explicit quadratic norm formulas [Reference Macdonald28, Reference Cherednik6, Reference Cherednik8, Reference Macdonald29], have explicit evaluation formulas, and are self-dual (see, for example, [Reference Macdonald29, Reference Cherednik8] and references therein). They admit combinatorial formulas [Reference Ram and Yip36, Reference Haglund, Haiman and Loehr18], arise as spherical functions on quantum groups (see, for example, [Reference Noumi33, Reference Letzter24, Reference Etingof and Kirilliv17]), admit interpretations as partition functions on integrable lattice models [Reference Borodin and Wheeler2], and degenerate to Heckman-Opdam polynomials [Reference Macdonald27], spherical functions on p-adic reductive groups [Reference Macdonald27], and Iwahori-Whittaker functions [Reference Ion22, Reference Cherednik and Ma11]. In this paper, we discuss generalisations of a number of these properties to the quasi-polynomial setup:

  1. 1. We define a normalized version $P_y^J(x;\mathfrak {t})$ of $E_y^J(x;\mathfrak {t})$ using the action of normalised Y-intertwiners on the lowest-degree quasi-polynomial eigenfunction $x^c$ . It leads to an explicit expression of $E_y^J(x;\mathfrak {t})$ in terms of Y-intertwiners (Theorem 6.13). This entails a quasi-polynomial analog of the evaluation formula for Macdonald polynomials.

  2. 2. We obtain a weak form of duality for $P_y^J(x;\mathfrak {t})$ , called pseudo-duality, by showing that the H-action on $P_y^J(x;\mathfrak {t})$ can be alternatively described in terms of an H-action on $y\in \mathcal {O}_c$ by discrete Demazure-Lusztig operators (Theorem 6.26).

  3. 3. We establish orthogonality relations and quadratic norm formulas for the quasi-polynomial eigenfunctions $E_y^J(x;\mathfrak {t})$ and $P_y^J(x;\mathfrak {t})$ (Theorem 6.43).

  4. 4. We construct (anti)symmetric variants $E_y^{J,\pm }(x;\mathfrak {t})$ of the quasi-polynomial eigenfunctions $E_y^J(x;\mathfrak {t})$ by (anti)symmetrisation with respect to the trivial (resp., sign) idempotent of $H_0$ . They generalize the (anti)symmetric Macdonald polynomials. The symmetric variants $E_{y}^{J, +}(x; \mathfrak {t})$ are invariant under the $W_0$ -action obtained by localizing $\pi _{c, \mathfrak {t}}$ (see Theorem 4.9 and Lemma 6.32). We obtain explicit expansion formulas of $E_y^{J,\pm }(x;\mathfrak {t})$ in terms of the quasi-polynomials $\{E_{y^\prime }^J(x;\mathfrak {t})\}_{y^\prime \in W_0y}$ (Corollary 6.38 and Proposition 6.41).

  5. 5. We establish the existence of the Whittaker limit $q\rightarrow \infty $ of $E_y^J(x;\mathfrak {t})$ and $E_y^{J,-}(x;\mathfrak {t})$ (Proposition 6.47 and Corollary 6.52). For y in the closure of the negative Weyl chamber and $y^\prime \in W_0y$ , we obtain an explicit formula for the limit $q \rightarrow \infty $ of $E_{y^\prime }^J(x;\mathfrak {t})$ in terms of the $\pi _{c,\mathfrak {t}}\vert _{H_0}$ -action on $x^y$ (Theorem 6.51). It generalises a formula due to Patnaik-Puskas [Reference Patnaik and Puskas34, Cor. 5.4], which expresses the metaplectic Iwahori-Whittaker function in terms of metaplectic Demazure-Lusztig operators. We will say more about the link with metaplectic representation theory in Subsection 1.5.

We also consider the theory for reductive root data, when both $\mathcal {P}$ and $\mathcal {P}_Y$ in $\mathbb {H}$ are extended to group algebras of a finitely generated abelian subgroup $\Lambda \subset E$ satisfying

(1.6) $$ \begin{align} Q^{\vee}\subseteq\Lambda\,\, \mbox{ and }\,\, \alpha(\Lambda)\subseteq\mathbb{Z}\,\,\,\, \forall\alpha\in\Phi_0. \end{align} $$

When only $\mathcal {P}_Y$ is extended to the group algebra of such a lattice $\Lambda ^\prime $ , then the associated Y-extended double affine Hecke algebra $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ is a smashed product algebra $\Omega _{\Lambda ^\prime }\ltimes \mathbb {H}$ , with $\Omega _{\Lambda ^\prime }\simeq \Lambda ^\prime /Q^{\vee }$ the subgroup of the extended affine Weyl group $W_{\Lambda ^\prime }:=W_0\ltimes \Lambda ^\prime $ consisting of elements of length zero. The associated extended quasi-polynomial representation $\pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime }:\mathbb {H}_{Q^{\vee },\Lambda ^\prime }\rightarrow \text {End}(\mathcal {P}^{(c)})$ now depends on parameters $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ , where $T_{\Lambda ^\prime ,J}$ is the affine subtorus of $T_{\Lambda ^\prime }=\text {Hom}(\Lambda ^\prime ,\mathbf {F}^\times )$ consisting of the elements $\mathfrak {t}\in T_{\Lambda ^\prime }$ satisfying $\mathfrak {t}^{\alpha _j^{\vee }}=1$ for all $j\in J$ .

The extra representation parameters in $\pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime }$ compared to $\pi _{c,\mathfrak {t}\vert _{Q^{\vee }}}$ are captured by the kernel $T_{\Lambda ^\prime ,\{1,\ldots ,r\}}$ of the restriction map $T_{\Lambda ^\prime }\rightarrow T$ . It identifies with the character group of $\Omega _{\Lambda ^\prime }$ and naturally embeds into the automorphism group of $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }\simeq \Omega _{\Lambda ^\prime }\ltimes \mathbb {H}$ fixing $\mathbb {H}$ point-wise. We call these extra parameters twist parameters because $\pi _{c,\mathfrak {t}\mathfrak {t}^\prime }^{Q^{\vee },\Lambda ^\prime }$ with $\mathfrak {t}^\prime \in T_{\Lambda ^\prime ,\{1,\ldots ,r\}}$ coincides with $\pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime }$ twisted by the automorphism of $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ associated to $\mathfrak {t}^\prime $ . Since these automorphisms rescale the $Y^\mu $ ( $\mu \in Q^{\vee }$ ), the quasi-polynomials $E_y^J(x;\mathfrak {t}\vert _{Q^{\vee }})$ ( $y\in \mathcal {O}_c$ ) are eigenfunctions of all the commuting operators $\pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime }(Y^\mu )$ ( $\mu \in \Lambda ^\prime $ ) under suitable generic conditions on the parameters (Theorem 7.26).

The full double affine Hecke algebra $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ depends on two lattices $\Lambda ,\Lambda ^\prime $ . The extension of the quasi-polynomial representation $\pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime }$ to $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ requires adding $(c,\Lambda )$ -dependent constraints on $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ , as well as enlarging the representation space. It will be discussed in detail in Section 7. In case of the $\text {GL}_{r+1}$ root datum, the quasi-polynomial eigenfunctions give rise to the metaplectic polynomials introduced in our previous work [Reference Sahi, Stokman and Venkateswaran38, §5.4]. We will say more about the root system generalisations of these metaplectic polynomials in Subsection 1.5.

1.4 The uniform quasi-polynomial representation

The quasi-polynomial representations defined in Theorem 1.1 can be naturally combined into a family of $\mathbb {H}$ -representations on $\mathbf {F}[E]$ . The parametrisation of this family of uniform quasi-polynomial representations requires that $q\in \mathbf {F}^\times $ has a $(2h)^{th}$ root $q^{\frac {1}{2h}}$ in $\mathbf {F}$ , where h is the Coxeter number of $W_0$ .

To motivate the parametrisation, it is instructive to consider an important special case first, which requires the stronger assumption that $q\in \mathbf {F}^\times $ is part of an injective group homomorphism

(1.7) $$ \begin{align} \mathbb{R}\rightarrow\mathbf{F}^\times,\qquad d\mapsto q^d. \end{align} $$

Then

(1.8) $$ \begin{align} \bigoplus_{c\in\overline{C}_+}\pi_{c,q^c}: \mathbb{H}\rightarrow\text{End}(\mathbf{F}[E]) \end{align} $$

is the prototypical example of a uniform quasi-polynomial representation, where $q^y\in T$ ( $y\in E$ ) is the character $\mu \mapsto q^{\langle y,\mu \rangle }$ of $Q^{\vee }$ . Note that for the underlying $\mathbb {W}$ -action on $\mathbf {F}[E]$ (given by (1.8) for $\mathbf {k}\equiv 1$ ), the translations $\tau (\mu )\in W$ are simply acting by q-dilations $x^y\mapsto q^{-\langle \mu ,y\rangle }x^y$ ( $y\in E$ ). Furthermore, for $c^\prime \in C^J$ and $\mathfrak {t}\in T_J$ , we have $\pi _{c^\prime ,\mathfrak {t}}\simeq \pi _{c,q^c}$ for some $c\in C^J$ if $\mathfrak {t}$ takes values in $\{q^d\}_{d\in \mathbb {R}}$ . The new set of parameters that we will assign to the uniform quasi-polynomial representation (1.8) is the tuple $\{\widehat {p}_\alpha \}_{\alpha \in \Phi _0}$ with $\widehat {p}_\alpha : \mathbb {R}\rightarrow \mathbf {F}^\times $ given by $\widehat {p}_\alpha (d):=q_\alpha ^{d/h}$ , where $q_\alpha :=q^{2/\|\alpha \|^2}$ . The link with the original representation parameters of the quasi-polynomial representation is through the formula

$$\begin{align*}q^c=\prod_{\alpha\in\Phi_0^+} \widehat{p}_{\alpha}(\alpha(c))^{\alpha} \end{align*}$$

for the character $q^c\in T$ .

In general, we will consider the group $\mathcal {G}^{\text {amb}}$ of tuples $\mathbf {f}=(f_\alpha )_{\alpha \in \Phi _0}$ consisting of $\mathbf {F}^\times $ -valued functions $f_\alpha $ on $\mathbb {R}$ . Let $\mathcal {G}$ be the subgroup of $\mathcal {G}^{\text {amb}}$ consisting of the tuples $\mathbf {g}=(g_\alpha )_{\alpha \in \Phi _0}$ satisfying

(1.9) $$ \begin{align} g_{v\alpha}(d+\ell)=g_\alpha(d)\qquad (v\in W_0,\,\ell\in\mathbb{Z}), \end{align} $$

$g_\alpha (-d)=g_\alpha (d)^{-1}$ and $g_\alpha (0)=1$ . If we replace (1.9) by the quasi-invariance condition

$$\begin{align*}\widehat{g}_{v\alpha}(d+\ell)=q_\alpha^{\ell/h}\widehat{g}_\alpha(d)\qquad (v\in W_0,\,\ell\in\mathbb{Z}), \end{align*}$$

we obtain a $\mathcal {G}$ -coset in $\mathcal {G}^{\text {amb}}$ , which we will denote by $\widehat {\mathcal {G}}$ .

For a tuple $\mathbf {f}=(f_\alpha )_{\alpha \in \Phi _0}\in \mathcal {G}^{\text {amb}}$ and $y\in E$ , we set

(1.10) $$ \begin{align} \mathfrak{t}_y(\mathbf{f}):=\prod_{\alpha\in\Phi_0^+}f_\alpha(\alpha(y))^\alpha\in T. \end{align} $$

Then $\mathfrak {t}_c(\widehat {\mathbf {g}})\in T_J$ for $c\in C^J$ when $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ .

Theorem 1.5. Let $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ and $\mathbf {g}\in \mathcal {G}$ . The formulas

$$ \begin{align*} \begin{aligned} \pi_{\mathbf{g},\widehat{\mathbf{p}}}(T_i)x^y&:=k_i^{\chi_{\mathbb{Z}}(\alpha_i(y))}g_{\alpha_i}(\alpha_i(y))^{-1}x^{s_iy}+(k_i-k_i^{-1})\nabla_i(x^y) \qquad (1\leq i\leq r),\\ \pi_{\mathbf{g},\widehat{\mathbf{p}}}(T_0)x^y&:=k_0^{\chi_{\mathbb{Z}}(\alpha_0(y))}g_{\varphi}(\varphi(y))\mathfrak{t}_y(\widehat{\mathbf{p}})^{\varphi^{\vee}}x^{s_\varphi y} +(k_0-k_0^{-1})\nabla_0(x^y) \end{aligned} \end{align*} $$

and $\pi _{\mathbf {g},\widehat {\mathbf {p}}}(x^\lambda )x^y:=x^{y+\lambda }$ for $\lambda \in Q^{\vee }$ define a $\mathbb {H}$ -representation on $\mathbf {F}[E]$ , which is isomorphic to $\bigoplus _{c\in \overline {C}_+}\pi _{c,\mathfrak {t}_c(\widehat {\mathbf {p}}\cdot \mathbf {g})}$ .

In the main text, we prove a version of this theorem for the extended double affine Hecke algebra $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ ; see Theorem 8.19. In that case, the extra twisting parameters are added by replacing $\mathfrak {t}_y(\widehat {\mathbf {p}})$ , viewed as a multiplicative character of $\Lambda ^\prime $ , by

$$\begin{align*}\mathfrak{t}_y(\widehat{\mathbf{p}},\mathfrak{c}):=\mathfrak{t}_y(\widehat{\mathbf{p}})\mathfrak{c}(\text{pr}_{E_{\text{co}}}(y)). \end{align*}$$

Here, $\text {pr}_{E_{\text {co}}}$ is the orthogonal projection of E onto the orthocomplement $E_{\text {co}}$ of $\mathbb {R}\Phi _0^{\vee }$ in E, and $\mathfrak {c}: E_{\text {co}}\rightarrow T_{\Lambda ^\prime ,[1,r]}$ satisfies $\mathfrak {c}(y+\mu )=q^\mu \mathfrak {c}(y)$ for $\mu \in \text {pr}_{E_{\text {co}}}(\Lambda ^\prime )$ .

The isomorphism $\pi _{\mathbf {g},\widehat {\mathbf {p}}}\simeq \bigoplus _{c\in \overline {C}_+}\pi _{c,\mathfrak {t}_c(\widehat {\mathbf {p}}\cdot \mathbf {g})}$ in (the extended version of) Theorem 1.5 will be realised by an explicit $\mathbf {g}$ -dependent linear automorphism of $\mathbf {F}[E]$ , which is diagonalised by the quasi-monomial basis $\{x^y\}_{y\in E}$ of $\mathbf {F}[E]$ . As a consequence, we obtain simultaneous quasi-polynomial eigenfunctions $\mathcal {E}_y(x;\mathbf {g},\widehat {\mathbf {p}})$ of $\pi _{\mathbf {g},\widehat {\mathbf {p}}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) by appropriately rescaling the quasi-monomials in the quasi-monomial expansion (1.4) of $E_y^J(x;\mathfrak {t}_c(\widehat {\mathbf {p}}\cdot \mathbf {g}))$ (see Corollary 8.24).

If $L\subset E$ is a $W_0$ -invariant, finitely generated abelian subgroup containing $\Lambda $ , then $\mathcal {P}_L:=\bigoplus _{\lambda \in L}\mathbf {F}x^\lambda $ is a $\mathbb {H}$ -subrepresentation of $\pi _{\mathbf {g},\widehat {\mathbf {p}}}$ containing the quasi-polynomial eigenfunctions $\mathcal {E}_\lambda (x;\mathbf {g},\widehat {\mathbf {p}})$ of degree $\lambda \in L$ . Since $\mathcal {P}_L$ may be viewed as the space of polynomials on the torus $T_L:=\text {Hom}(L,\mathbf {F}^\times )$ , we thus obtain a polynomial representation of $\mathbb {H}$ with $\pi _{\mathbf {g},\widehat {\mathbf {p}}}(T_j)\vert _{\mathcal {P}_L}$ acting by truncated Demazure-Lusztig operators (these operators are still truncated since the roots $\alpha \in \Phi _0$ are allowed to take non-integral values on L). The decomposition of L into W-orbits provides a decomposition of $\pi _{\mathbf {g},\widehat {\mathbf {p}}}(\cdot )\vert _{\mathcal {P}_L}$ as a direct sum of quasi-polynomial representations, with the number of free parameters in $\pi _{\mathbf {g},\widehat {\mathbf {p}}}(\cdot )\vert _{\mathcal {P}_L}$ depending on L. This representation plays an important role in establishing the connection to metaplectic representation theory, which we discuss in the next subsection.

1.5 The metaplectic theory

The metaplectic root system $\Phi _0^m$ and the associated metaplectic double affine Hecke algebra $\mathbb {H}^m$ are defined in terms of a metaplectic datum. A metaplectic datum is a pair $(n,\mathbf {Q})$ consisting of a positive integer n and a rational $W_0$ -invariant quadratic form $\mathbf {Q}$ on $\Lambda $ taking integral values on $Q^{\vee }$ . The associated symmetric bilinear form on $\Lambda $ will be denoted by $\mathbf {B}$ . The metaplectic root system $\Phi _0^m:=\{\alpha ^m\}_{\alpha \in \Phi _0}$ consists of the rescaled roots $\alpha ^m=m(\alpha )^{-1}\alpha $ where

$$\begin{align*}m(\alpha):=\frac{n}{\text{gcd}(n,\mathbf{Q}(\alpha^{\vee}))}. \end{align*}$$

This is a root system with basis $\{\alpha _1^m,\ldots ,\alpha _r^m\}$ , which is isomorphic to either $\Phi _0$ or $\Phi _0^{\vee }$ . We denote by $\vartheta \in \Phi _0^+$ the root such that $\vartheta ^m$ is the long root of $\Phi _0^m$ . The simple affine root for the associated metaplectic affine root system $\Phi ^m$ is $\alpha _0^m=(-\vartheta ^m,1)$ . The co-root lattice $Q^{m\vee }$ of $\Phi _0^m$ is contained in $\Lambda $ , and $\alpha ^m(\Lambda )\subseteq m(\alpha )^{-1}\mathbb {Z}$ for $\alpha \in \Phi _0$ . The metaplectic double affine Hecke algebra $\mathbb {H}^m$ is the double affine Hecke algebra relative to the metaplectic root system $\Phi _0^m$ . It depends on a choice $\mathbf {k}=(k_{a^m})_{a^m\in \Phi ^m}$ of a multiplicity function of $\Phi ^m$ .

To connect the uniform quasi-polynomial representation and the associated quasi-polynomial eigenfunctions with the metaplectic representation theory and the associated metaplectic polynomials from [Reference Sahi, Stokman and Venkateswaran38], we consider the polynomial representation $\pi _{\mathbf {g},\widehat {\mathbf {p}}}(\cdot )\vert _{\mathcal {P}_L}$ with $\Phi _0$ and $\mathbb {H}$ replaced by their metaplectic variants $\Phi _0^m$ and $\mathbb {H}^m$ , with L a lattice $\Lambda $ satisfying (1.6) relative to the original, nonmetaplectic root system $\Phi _0$ , and with base-point $\widehat {\mathbf {p}}$ satisfying $\mathfrak {t}_\lambda (\widehat {\mathbf {p}})=q^\lambda $ for all $\lambda \in L$ . The remaining parameter dependencies are captured by the set $\mathcal {M}_{(n,\mathbf {Q})}$ of tuples $\underline {h}=(h_s(\alpha ))_{s\in \mathbf {Q}(\alpha ^{\vee })\mathbb {Z},\alpha \in \Phi _0}$ consisting of $W_0$ -invariant functions $h_s: \Phi _0\rightarrow \mathbf {F}^\times $ ( $s\in \mathbf {Q}(\alpha ^{\vee })\mathbb {Z}$ ) satisfying the additional requirements that $s\mapsto h_s(\alpha )$ is a $\text {lcm}(n,\mathbf {Q}(\alpha ^{\vee }))$ -periodic function, $h_0(\alpha )=-1$ , and

$$\begin{align*}h_s(\alpha)h_{-s}(\alpha)=k_{\alpha^m}^{-2}\qquad\qquad \forall\, s\in\mathbf{Q}(\alpha^{\vee})\mathbb{Z}\setminus\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}. \end{align*}$$

The explicit relation with the parameter set $\mathcal {G}$ for the uniform quasi-polynomial representation is described in Lemma 9.2. This specialisation leads to the following metaplectic basic representation of $\mathbb {H}^m$ . Let $r_\ell (s)\in \{0,\ldots ,\ell -1\}$ be the remainder of $s\in \mathbb {Z}$ modulo $\ell \in \mathbb {Z}_{>0}$ , and write $k_j=k_{\alpha _j^m}$ for $0\leq j\leq r$ .

Theorem 1.6. Let $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ . The formulas

$$ \begin{align*} \pi_\Lambda^m(T_0)x^\lambda&=-k_0h_{\mathbf{B}(\lambda,\vartheta^{\vee})}(\vartheta)q_\vartheta^{m(\vartheta)\vartheta(\lambda)}x^{s_{\vartheta}\lambda}\\ &\quad+ (k_0-k_0^{-1})\left(\frac{1-(q_\vartheta^{m(\vartheta)}x^{-\vartheta^{\vee}})^{(r_{m(\vartheta)}(-\vartheta(\lambda))+\vartheta(\lambda))}}{1-(q_\vartheta^{m(\vartheta)}x^{-\vartheta^{\vee}})^{m(\vartheta)}}\right)x^\lambda,\\ \pi_\Lambda^m(T_i)x^\lambda&=-k_ih_{-\mathbf{B}(\lambda,\alpha_i^{\vee})}(\alpha_i)x^{s_i\lambda}+(k_i-k_i^{-1})\left(\frac{1-x^{(r_{m(\alpha_i)}(\alpha_i(\lambda))-\alpha_i(\lambda))\alpha_i^{\vee}}}{1-x^{m(\alpha_i)\alpha_i^{\vee}}}\right)x^\lambda,\\ \pi_\Lambda^m(x^\mu)x^\lambda&=x^{\lambda+\mu} \end{align*} $$

for $1\leq i\leq r$ , $\mu \in Q^{m\vee }$ and $\lambda \in \Lambda $ define a representation $\pi _\Lambda ^m: \mathbb {H}^m\rightarrow \text {End}(\mathcal {P}_\Lambda )$ .

The metaplectic basic representation $\pi _\Lambda ^m$ naturally extends to the metaplectic extended double affine Hecke algebra; see Theorem 9.5.

At this stage, we can make the link with the (announced) results in [Reference Sahi, Stokman and Venkateswaran38]. The representation $\pi _\Lambda ^m$ , restricted to the metaplectic affine Hecke algebra generated by $\mathcal {P}$ and $T_1,\ldots ,T_r$ is the metaplectic affine Hecke algebra representation from [Reference Sahi, Stokman and Venkateswaran38, Thm. 3.7]. When the multiplicity functions $\mathbf {k}$ and $h_s$ are constant, it recovers, after localisation, Chinta’s and Gunnells’ [Reference Chinta and Gunnells13, Reference Chinta and Gunnells14] $W_0$ -action on a suitable subspace of rational functions on $T_\Lambda $ , which plays an important role in the construction of the local parts of Weyl group multiple Dirichlet series. By X-localisation of the double affine Hecke algebra, Theorem 1.6 gives rise to families of W-actions which generalise the Chinta-Gunnells action; see Corollary 9.16 (compare also with [Reference Patnaik and Puskas35, §3.3] and [Reference Lee and Zhang23] where the Chinta-Gunnells action is extended to an arbitrary Coxeter group). These W-actions actually extend to actions of the double affine Weyl group $\mathbb {W}$ , with the second coroot lattice $Q^{\vee }$ acting by translations of the exponents of the monomials. The localisation procedure can also be applied to the (uniform) quasi-polynomial representations $\pi _{c,\mathfrak {t}}$ and $\pi _{\mathbf {g},\widehat {\mathbf {p}}}$ ; see Theorem 4.9 and Theorem 8.27 for the resulting $\mathbb {W}$ -actions.

Rescaling the quasi-monomials in the expansion (1.4) gives rise to simultaneous eigenfunctions $E_\lambda ^m(x)\in \mathcal {P}_\Lambda $ of the commuting operators $\pi _{\Lambda }^m(Y^\mu )$ ( $\mu \in Q^{m\vee }$ ) satisfying

$$\begin{align*}E_\lambda^m(x)=x^\lambda+\sum_{\lambda^\prime<\lambda}e_{\lambda,\lambda^\prime}^m x^{\lambda^\prime}\qquad (e_{\lambda,\lambda^\prime}^m\in\mathbf{F}) \end{align*}$$

under suitable generic conditions on the metaplectic parameters $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ ; see Theorem 9.18. An antisymmetric version of the metaplectic polynomial $E_\lambda ^m(x)$ , denoted $E_{\lambda }^{m, -}(x)$ , is obtained by acting by the sign-idempotent of the finite Hecke algebra in $\mathbb {H}^m$ on $E_\lambda ^m(x)$ . For the root datum of type $\text {GL}_{r+1}$ , the representation $\pi _\Lambda ^m$ of the extended metaplectic double affine Hecke algebra recovers the representation introduced in [Reference Sahi, Stokman and Venkateswaran38, Thm. 5.4], and the $E_\lambda ^m(x)$ are the metaplectic polynomials introduced in [Reference Sahi, Stokman and Venkateswaran38, Thm. 5.7].

We consider the $q\rightarrow \infty $ limit of the metaplectic polynomials, for which we need to assume that $\mathbf {F}=\mathbf {K}(q^{\frac {1}{2h}})$ and that the remaining parameters $\mathbf {k}$ and $\underline {h}$ take values in the multiplicative group $\mathbf {K}^\times $ of the field $\mathbf {K}$ of characteristic zero. The Whittaker limit

$$\begin{align*}\overline{E}_\lambda^m(x):=E_\lambda^m(x)\vert_{q^{-\frac{1}{2h}}=0}\qquad (\lambda\in\Lambda) \end{align*}$$

of $E_\lambda ^m(x)$ then defines a Laurent polynomial in $\mathcal {P}_\Lambda $ with coefficients in $\mathbf {K}$ . They are related to metaplectic Iwahori-Whittaker functions [Reference Patnaik and Puskas34] as follows.

Suppose that $\mathbf {K}$ is a non-archimedean local field containing the $n^{\text {th}}$ roots of unity, with the cardinality $\upsilon $ of its residue field congruent $1$ module $2n$ . In [Reference Patnaik and Puskas34], Patnaik and Puskas introduced Iwahori-Whittaker functions for metaplectic covers of reductive groups over $\mathbf {K}$ . They expressed them in terms of certain metaplectic Demazure-Lusztig operators which depend on certain Gauss sums $\mathbf {g}_s$ , as well as on $\upsilon $ ; see [Reference Patnaik and Puskas34, Cor. 5.4]. Comparing this with the analogous formula for $\overline {E}_{\lambda }^{m}(x)$ in terms of truncated Demazure-Lusztig operators (Theorem 6.51), we establish in Subsection 10 the following link between $\overline {E}_\lambda ^m(x)$ and metaplectic Iwahori-Whittaker functions. Let I the linear automorphism of $\mathcal {P}_\Lambda $ satisfying $I(x^\mu )=x^{-\mu }$ for $\mu \in \Lambda $ , let $\overline {E}_\lambda ^{m,-}(x)$ be the Whittaker limit of $E_\lambda ^{m,-}(x)$ , and denote by $\rho ^{\vee }$ the half-sum of positive co-roots.

Theorem 1.7. Suppose that the multiplicity function $\mathbf {k}$ is constantly equal to $\upsilon ^{-2}$ and the multiplicity functions $h_s$ of the metaplectic parameters $\underline {h}$ are constantly equal to $\mathbf {g}_s$ . For $v\in W_0$ and for $\lambda \in \Lambda $ lying in the closure of the positive Weyl chamber,

$$\begin{align*}x^{-\rho^{\vee}}I\big(\overline{E}^m_{-v(\lambda+\rho^{\vee})}(x)\big)\quad \mbox{ and }\quad x^{-\rho^{\vee}}I\big(\overline{E}^{m,-}_{-\lambda-\rho^{\vee}}(x)\big) \end{align*}$$

are constant multiples of metaplectic Iwahori-Whittaker functions and metaplectic spherical Whittaker functions, respectively.

In Subsection 10, we also recover McNamara’s [Reference McNamara30] Casselman-Shalika type formula for the metaplectic spherical Whittaker function as a consequence of the explicit expansion formula of $E_y^{J,-}(x;\mathfrak {t})$ as a linear combination of the quasi-polynomial eigenfunctions $E_{y^\prime }^J(x;\mathfrak {t})$ ( $y^\prime \in W_0 y$ ), derived earlier in Subsection 6.6.

1.6 Structure of the paper

In Section 2, we collect basic properties of the affine Weyl group W and the double affine Hecke algebra $\mathbb {H}$ .

In Section 3, we introduce the class of Y-parabolically induced cyclic $\mathbb {H}$ -modules $\mathbb {M}_t^J$ .

In Section 4, we introduce the quasi-polynomial representation $\pi _{c,\mathfrak {t}}$ . We state the main theorem that the quasi-polynomial representation is well defined and isomorphic to $\mathbb {M}_{\mathfrak {s}_J\mathfrak {t}}^J$ in Subsection 4.3. We discuss the dependence of $\pi _{c,\mathfrak {t}}$ on $c\in C^J$ and J in Subsection 4.4, and localise it to obtain a nontrivial W-action on quasi-rational functions in Subsection 4.5.

Section 5 is devoted to the proof of the main theorem. As part of the proof, we establish in Subsection 5.5 triangularity properties of the commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) with respect to the basis of quasi-monomials, ordered by the parabolic Bruhat order on their exponents.

In Section 6, we define the quasi-polynomial generalisations $E_y^J(x;\mathfrak {t})$ of the monic Macdonald polynomials for adjoint root data. We show how they can be explicitly created from the quasi-monomial $x^c$ by acting by Y-intertwiners in Subsection 6.3. Their dependence on $J\subsetneq \{0,\ldots ,r\}$ is discussed in Subsection 6.4. Pseudo-duality for the normalised versions $P_y^J(x;\mathfrak {t})$ of the quasi-polynomials $E_y^J(x;\mathfrak {t})$ is derived in Subsection 6.5, quasi-polynomial generalisations $E_y^{J,\pm }(x;\mathfrak {t})$ of the (anti)symmetric Macdonald polynomials are discussed in Subsection 6.6, and orthogonality relations and quadratic norm formulas are obtained in Subsection 6.7. The Whittaker limit $q\rightarrow \infty $ of $E_y^J(x;\mathfrak {t})$ and $E_y^{J,-}(x;\mathfrak {t})$ is considered in Subsection 6.8.

In Section 7, we generalise the results of Sections 4-6 to reductive root data (including the $\text {GL}_{r+1}$ root datum as a special case). We reinterpret the role of the extra parameters in terms of twisting by automorphisms of the extended double affine Hecke algebra in Subsection 7.4. In Subsection 7.5, we introduce the analogs of the quasi-polynomial eigenfunctions $E_y^J(x;\mathfrak {t})$ for reductive root datum. We furthermore show that, under suitable generic conditions on the parameters, they are independent of the twist parameters and reduce to the quasi-polynomial eigenfunctions for the underlying adjoint root datum.

In Section 8, we introduce the uniform quasi-polynomial representation (Subsection 8.2), the associated uniform quasi-polynomial eigenfunctions (Subsection 8.3), and we explicitly relate the latter to the quasi-polynomial eigenfunctions $E_y^J(x;\mathfrak {t})$ , using an automorphism that rescales the quasi-monomials.

In Section 9, we attach to a metaplectic datum a polynomial representation of the metaplectic double affine Hecke algebra, called the metaplectic basic representation (Subsection 9.2), and we show that it is isomorphic to a suitable sub-representation of the uniform quasi-polynomial representation. We obtain an affine version of the Chinta-Gunnells’ $W_0$ -action on rational functions by X-localising the metaplectic basic representation in Subsection 9.3. We introduce metaplectic polynomials, which are simultaneous eigenfunctions for the action of $\mathcal {P}_Y$ under the metaplectic basic representation, and relate them to the quasi-polynomial eigenfunctions $E_y^J(x;\mathfrak {t})$ in Subsection 9.4.

In Section 10, we show the existence of the Whittaker limit of the metaplectic polynomial. In the context of representation theory of metaplectic covers of reductive groups over non-archimedean local fields, we prove that the Whittaker limit of the metaplectic polynomial is a metaplectic Iwahori-Whittaker function up to some elementary twists. We rederive McNamara’s [Reference McNamara30] Casselman-Shalika type formula for the metaplectic spherical Whittaker function from the Whittaker limit of an explicit expansion formula for the anti-symmetrised version of the quasi-polynomial eigenfunction derived in Subsection 6.6.

1.7 Conventions

We take $\mathbf {F}$ to be a field of characteristic zero, and we fix $q\in \mathbf {F}^\times $ not a root of unity (it contains the case of formal variable q by taking $\mathbf {F}=\mathbf {K}(q)$ for some field $\mathbf {K}$ of characteristic zero). For a commutative ring R and an abelian group A, we write $R[A]= \bigoplus _{y\in A}Rx^y$ for the group algebra of A over R (i.e., $x^yx^{y^\prime }=x^{y+y^\prime }$ for $y,y^\prime \in A$ and $x^0=1$ ). We will often apply this to the case that A is a Euclidean space.

For groups G and H, we write $\text {Hom}(G,H)$ for the class of group homomorphisms $G\rightarrow H$ . For a finitely generated free abelian group $\Lambda $ , we write $T_\Lambda :=\text {Hom}(\Lambda ,\mathbf {F}^\times )$ for the $\mathbf {F}$ -torus of multiplicative characters $\Lambda \rightarrow \mathbf {F}^\times $ , and $\mathcal {P}_\Lambda =\mathbf {F}[\Lambda ]$ for the group algebra of $\Lambda $ over $\mathbf {F}$ . Alternatively, it is the algebra of regular $\mathbf {F}$ -valued functions on $T_\Lambda $ , where the monomial $x^\lambda $ maps $t\in T_\Lambda $ to its value $t^\lambda $ at $\lambda \in \Lambda $ . Finally, for $\ell ,m\in \mathbb {Z}$ with $\ell \leq m$ , we write $[\ell ,m]:=\{\ell ,\ell +1,\ldots ,m\}$ .

2 Preliminaries on double affine Weyl and Hecke algebras

In this section, we fix the notations for root systems, reflection groups and double affine Weyl and Hecke algebras. For ease of reference, we also list well-known properties that will be regularly used in subsequent sections. For further details, see, for instance, [Reference Humphreys21, Reference Macdonald29] for root systems and reflection groups, and [Reference Cherednik8, Reference Macdonald29] for double affine Weyl and Hecke algebras.

2.1 Root systems and Weyl groups

Let E be an Euclidean space with scalar product $\langle \cdot ,\cdot \rangle $ and corresponding norm $\|\cdot \|$ . Let $E^*$ be its linear dual. We turn $E^*$ into an Euclidean space by transporting the scalar product of E through the linear isomorphism $E\overset {\sim }{\longrightarrow } E^*$ , $y\mapsto \langle y,\cdot \rangle $ . The resulting scalar product and norm on $E^*$ are denoted by $\langle \cdot ,\cdot \rangle $ and $\|\cdot \|$ again.

Let $\Phi _0\subset E^*$ be a reduced irreducible root system in $\text {span}_{\mathbb {R}}\{\alpha \, | \, \alpha \in \Phi _0\}$ . We normalise $\Phi _0$ in such a way that long roots have squared length equal to $2/m^2$ for some $m\in \mathbb {Z}_{>0}$ . Set $E_{\text {co}}:=\cap _{\alpha \in \Phi _0}\text {Ker}(\alpha )$ , and write $E^\prime \subseteq E$ for the orthogonal complement of $E_{\text {co}}$ in E, so that $E=E^\prime \oplus E_{\text {co}}$ (orthogonal direct sum). Write $\text {pr}_{E^\prime }$ and $\text {pr}_{E_{\text {co}}}$ for the corresponding projections onto $E^\prime $ and $E_{\text {co}}$ , respectively.

For $\alpha \in \Phi _0$ , write $s_\alpha \in \text {GL}(E^*)$ for the reflection

$$\begin{align*}s_\alpha(\xi):=\xi-\xi(\alpha^{\vee})\alpha\qquad (\xi\in E^*), \end{align*}$$

where $\alpha ^{\vee }\in E$ is the unique vector such that $\langle y,\alpha ^{\vee }\rangle = 2\alpha (y)/\|\alpha \|^2$ for all $y\in E$ . The Weyl group $W_0$ of $\Phi _0$ is the subgroup of $\text {GL}(E^*)$ generated by $s_\alpha $ ( $\alpha \in \Phi _0$ ). The root system $\Phi _0\subset E^*$ is $W_0$ -invariant. We denote by $Q\subset E^*$ the root lattice of $\Phi _0$ .

Write $\Phi _0^{\vee }= \{\alpha ^{\vee }\}_{\alpha \in \Phi _0}\subset E$ for the associated coroot system. Note that $\Phi _0^{\vee }\subset E^\prime $ , and note furthermore that $E^\prime =\text {span}_{\mathbb {R}}\{\alpha ^{\vee } \,\, | \,\, \alpha \in \Phi _0\}$ . The coroot lattice and co-weight lattice of $\Phi _0$ are

$$\begin{align*}Q^{\vee}:=\mathbb{Z}\Phi_0^{\vee},\qquad P^{\vee}:=\{\lambda\in E^\prime \,\, | \,\, \alpha(\lambda)\in\mathbb{Z}\quad \forall\, \alpha\in\Phi_0\}. \end{align*}$$

They are full sub-lattices of $E^\prime $ , and $Q^{\vee }\subseteq P^{\vee }$ . By the choice of normalisation of the roots in $\Phi _0$ , we have $Q^{\vee }\subseteq m^2Q$ upon identifying $E^*\simeq E$ via the scalar product $\langle \cdot ,\cdot \rangle $ on E.

Consider the action of $W_0$ on E with $s_\alpha $ ( $\alpha \in \Phi _0$ ) acting as the orthogonal reflection in the hyperplane $\alpha ^{-1}(0)$ ,

$$\begin{align*}s_\alpha(y):=y-\alpha(y)\alpha^{\vee}\qquad (y\in E). \end{align*}$$

Then $(v\xi )(y)=\xi (v^{-1}y)$ for $v\in W_0$ , $\xi \in E^*$ and $y\in E$ ; hence,

$$\begin{align*}v(\alpha^{\vee})=(v\alpha)^{\vee}\qquad\quad (v\in W_0,\,\alpha\in\Phi_0). \end{align*}$$

In particular, $\Phi _0^{\vee }\subset E$ is $W_0$ -invariant, and hence, so are the lattices $Q^{\vee }$ and $P^{\vee }$ .

Fix a set $\Delta _0:=\{\alpha _1,\ldots ,\alpha _r\}$ of simple roots and write $\Phi _0=\Phi ^+_0\cup \Phi ^-_0$ for the resulting natural division of the root system in positive and negative roots. Let $\Delta _0^{\vee }:=\{\alpha _1^{\vee },\ldots ,\alpha _r^{\vee }\}$ be the associated set of simple coroots for $\Phi _0^{\vee }$ and $\Phi _0^{\vee ,\pm }$ the associated sets of positive and negative coroots. We have $P^{\vee }=\bigoplus _{i=1}^r\mathbb {Z}\varpi _i^{\vee }$ with $\varpi _i^{\vee }\in E^\prime $ ( $1\leq i\leq r$ ) the fundamental weights with respect to $\Delta _0$ , characterised by $\alpha _i(\varpi _j^{\vee })=\delta _{i,j}$ ( $1\leq i,j\leq r$ ). Set $P^{\vee ,\pm }:=\pm \sum _{i=1}^r\mathbb {Z}_{\geq 0}\varpi _i^{\vee }$ for the cones of dominant and anti-dominant co-weights in $P^{\vee }$ , respectively.

The Weyl group $W_0$ is a Coxeter group, with Coxeter generators the simple reflections $s_i:=s_{\alpha _i}$ ( $1\leq i\leq r$ ). We write $w_0\in W_0$ for the longest Weyl group element, and $\varphi \in \Phi _0^+$ for the highest root. Note that $E^\prime $ is $W_0$ -invariant, and $W_0$ acts trivially on $E_{\text {co}}$ .

We write

(2.1) $$ \begin{align} E^{\text{reg}}:=\{y\in E \,\,\, | \,\,\, \alpha(y)\not=0 \quad \forall\, \alpha\in\Phi_0\} \end{align} $$

for the regular elements in E, and $E_+\subset E^{\text {reg}}$ for the fundamental Weyl chamber with respect to $\Delta _0$ ,

$$\begin{align*}E_+=\{y\in E \,\, | \,\, \alpha(y)>0\quad \forall\, \alpha\in\Phi_0^+\}. \end{align*}$$

The Weyl chamber opposite to $E_+$ will be denoted by $E_-$ . Note that $E_{\pm }$ are $E_{\text {co}}$ -translation invariant, and $E^{\prime }_{\pm }:=E^\prime \cap E_{\pm }$ form complete sets of representatives of the $E_{\text {co}}$ -orbits. Furthermore, $\overline {E}_{\pm }$ are fundamental domains for the $W_0$ -action on E. For $y\in E$ , we denote by $y_\pm \in \overline {E}_{\pm }$ the unique element such that $W_0y\cap \overline {E}_{\pm }=\{y_{\pm }\}$ . Note that $P^{\vee ,\pm }=P^{\vee }\cap \overline {E}_{\pm }$ .

2.2 Affine root systems and affine Weyl groups

The affine Weyl group is the semi-direct product group $W:=W_0\ltimes Q^{\vee }$ . We extend the $W_0$ -action on E to a faithful W-action on E by affine linear transformations, with $\mu \in Q^{\vee }$ acting by

$$\begin{align*}\tau(\mu)y:=y+\mu\qquad (y\in E). \end{align*}$$

We will regularly identify W with its realisation as a subgroup of the group of affine linear automorphisms of E.

Note that $W_0\subset W$ is the subgroup of elements $w\in W$ fixing the origin $0\in E$ . More generally, we write

(2.2) $$ \begin{align} W_y:=\{w\in W \,\,\, | \,\,\, wy=y\}\qquad (y\in E), \end{align} $$

which is always a finite subgroup of W.

The affine Weyl group W contains, besides the orthogonal reflections $s_\alpha $ ( $\alpha \in \Phi _0$ ), orthogonal reflections in affine hyperplanes. The affine hyperplanes can be described by an affine root system $\Phi $ of untwisted type in the following manner.

We identify $E^*\oplus \mathbb {R}$ with the space of affine linear functionals on E by interpreting $(\xi ,d)\in E^*\oplus \mathbb {R}$ as the affine linear functional $y\mapsto \xi (y)+d$ ( $y\in E$ ). The affine root system $\Phi $ is then defined by

$$\begin{align*}\Phi:=\{(\alpha,\ell)\,\,\, | \,\,\, \alpha\in\Phi_0\,\, \& \,\, \ell\in\mathbb{Z} \}\subset E^*\oplus\mathbb{R}. \end{align*}$$

For an affine root $a=(\alpha ,\ell )\in \Phi $ , we denote by $s_a$ the orthogonal reflection in the affine hyperplane $a^{-1}(0)\subset E$ . We then have for $a\in \Phi $ ,

$$\begin{align*}s_a(y)=y-a(y)\alpha^{\vee}. \end{align*}$$

In particular,

$$\begin{align*}s_a=\tau(-\ell\alpha^{\vee})s_\alpha\in W\qquad\quad (a=(\alpha,\ell)\in\Phi), \end{align*}$$

showing that the affine Weyl group W is generated by the reflections $s_a$ ( $a\in \Phi $ ).

The formulas

(2.3) $$ \begin{align} v(\xi,d)=(v\xi,d),\qquad \tau(\mu)(\xi,d)=(\xi,d-\xi(\mu)) \end{align} $$

for $v\in W_0$ , $\mu \in Q^{\vee }$ and $(\xi ,d)\in E^*\oplus \mathbb {R}$ define a linear W-action on the space $E^*\oplus \mathbb {R}$ of affine linear functionals on E. It is contragredient to the action of W on E by reflections and translations,

$$\begin{align*}(wf)(y)=f(w^{-1}y)\qquad\quad (w\in W,\, f\in E^*\oplus\mathbb{R},\, y\in E). \end{align*}$$

Note that $\Phi \subset E^*\oplus \mathbb {R}$ is W-stable.

We identify $E^*$ with the subspace $\{(\xi ,0)\,\, | \,\, \xi \in E^*\}$ of $E^*\oplus \mathbb {R}$ via the embedding $\xi \mapsto (\xi ,0)$ ( $\xi \in E^*$ ). This is an embedding of $W_0$ -modules. The second formula of (2.3) then gives

$$\begin{align*}\tau(\mu)\xi=(\xi,-\xi(\mu))\qquad (\xi\in E^*,\, \mu\in Q^{\vee}). \end{align*}$$

The set $\Delta :=\{\alpha _0,\ldots ,\alpha _r\}$ , with the simple affine root $\alpha _0$ defined by

(2.4) $$ \begin{align} \alpha_0:=(-\varphi,1)\in\Phi, \end{align} $$

forms a basis of the affine root system $\Phi $ . It gives rise to the division $\Phi =\Phi ^+\cup \Phi ^-$ of $\Phi $ in positive and negative roots, with

(2.5) $$ \begin{align} \Phi^+:=\mathbb{Z}_{\geq 0}\Delta\cap\Phi=\Phi^+_0\cup\{(\alpha,\ell) \,\,\, | \,\,\, (\alpha,\ell)\in\Phi_0\times\mathbb{Z}_{>0}\} \end{align} $$

and $\Phi ^-:=-\Phi ^+$ . The reflection

(2.6) $$ \begin{align} s_0:=s_{\alpha_0}=\tau(\varphi^{\vee})s_{\varphi}\in W \end{align} $$

is called the simple affine reflection of W. The affine Weyl group W is a Coxeter group with Coxeter generators $s_0,\ldots ,s_r$ .

Let $D: E^*\oplus \mathbb {R}\rightarrow E^*$ be the projection on the first component. Note that $Df$ is the gradient of $f\in E^*\oplus \mathbb {R}$ , viewed as affine linear functional on E. It induces a group epimorphism $W\twoheadrightarrow W_0$ , which we also denote by D, by requiring that

$$\begin{align*}D(wf)=(Dw)Df\qquad (w\in W,\, f\in E^*\oplus\mathbb{R}). \end{align*}$$

Concretely, $D(v\tau (\mu ))=v$ for $v\in W_0$ and $\mu \in Q^{\vee }$ .

The action of W on E preserves

(2.7) $$ \begin{align} E^{a,\text{reg}}:=\{y\in E \,\, | \,\, a(y)\not=0\quad \forall a\in\Phi \}. \end{align} $$

The connected components of $E^{a,\text {reg}}$ are called alcoves. The resulting W-action on the set of alcoves is simply transitive. The alcove

(2.8) $$ \begin{align} \begin{aligned} C_+&:=\{ y\in E \,\, | \,\, a(y)>0\quad \forall\, a\in\Phi^+ \}\\ &=\{ y\in E \,\, | \,\, 0<\alpha(y)<1\quad \forall\, \alpha\in\Phi_0^+ \} \end{aligned} \end{align} $$

is the fundamental alcove in E. For $y\in E$ , we write

(2.9) $$ \begin{align} \mathcal{O}_y:=Wy \end{align} $$

for the W-orbit of y in E. We thus have

(2.10) $$ \begin{align} E^{a,\text{reg}}=\bigsqcup_{c\in C_+}\mathcal{O}_c,\qquad\quad E=\bigsqcup_{c\in\overline{C}_+}\mathcal{O}_c \end{align} $$

(disjoint unions).

2.3 Length function and parabolic subgroups

We recall some properties of the length function

(2.11) $$ \begin{align} \ell(w):=\text{Card}\big(\Pi(w)\big),\qquad \Pi(w):=\Phi^+\cap w^{-1}\Phi^- \end{align} $$

that we will frequently use.

Lemma 2.1. Let $0\leq j\leq r$ and $w\in W$ .

  1. 1. $|\ell (s_jw)-\ell (w)|=1$ .

  2. 2. We have

    $$\begin{align*}\ell(s_jw)=\ell(w)+1 \quad\Leftrightarrow \quad w^{-1}\alpha_j\in \Phi^+, \end{align*}$$
    and then $\Pi (s_jw)=\{w^{-1}\alpha _j\}\cup \Pi (w)$ (disjoint union).

Proof. See, for example, [Reference Macdonald29, (2.2.8)].

The length $\ell (w)$ of $w\in W$ is the smallest nonnegative integer $\ell $ for which there exists an expression $w=s_{j_1}\cdots s_{j_\ell }$ of w as product of $\ell $ simple reflections ( $0\leq j_i\leq r$ ). Such shortest length expressions are called reduced expressions. If $w=s_{j_1}\cdots s_{j_\ell }$ is a reduced expression, then

(2.12) $$ \begin{align} \Pi(w)= \{b_1,\ldots,b_\ell\}\,\,\, \mbox{ with }\, b_m:=s_{j_\ell}\cdots s_{j_{m+1}}(\alpha_{j_m})\,\,\, (1\leq m<\ell),\quad b_\ell:=\alpha_{j_\ell}. \end{align} $$

Restricting $\ell $ to $W_0$ gives the length function of $W_0$ relative to $\{s_1,\ldots ,s_r\}$ . The following important length identity

(2.13) $$ \begin{align} \ell(\tau(\mu)v)=\sum_{\alpha\in\Phi_0^+\cap v\Phi_0^-}|\alpha(\mu)-1|+\sum_{\alpha\in\Phi_0^+\cap v\Phi_0^+}|\alpha(\mu)|\qquad (\mu\in Q^{\vee},\, v\in W_0), \end{align} $$

is obtained by explicitly describing the affine roots in $\Pi (\tau (\mu )v)$ (see, for example, [Reference Macdonald29, (2.4.1)]).

Fix a subset $J\subseteq [0,r]$ . The associated parabolic subgroup $W_J\subseteq W$ is the subgroup of W generated by $s_j$ ( $j\in J$ ). Write $W^{J}$ for the set of elements $w\in W$ such that

$$\begin{align*}\ell(ww^\prime)=\ell(w)+\ell(w^\prime)\qquad \forall\, w^\prime\in W_{J}. \end{align*}$$

Write $\Phi _{J}:=\Phi \cap \text {span}_{\mathbb {Z}}\{\alpha _j\}_{j\in J}$ and $\Phi _J^{\pm }:=\Phi ^{\pm }\cap \Phi _J$ ; then

(2.14) $$ \begin{align} W^{J}=\{w\in W \,\, | \,\, w\Phi_J^+\subseteq\Phi^+ \}. \end{align} $$

Lemma 2.2. Fix a subset $J\subseteq [0,r]$ .

  1. 1. $W^J$ is a complete set of representatives of $W/W_J$ (the elements in $W^J$ are called the minimal coset representatives of $W/W_J$ ).

  2. 2. Let $w\in W^J$ and $0\leq j\leq r$ . Then

    $$\begin{align*}s_jw\not\in W^J\,\,\,\Leftrightarrow\,\,\, s_jwW_J=wW_J, \end{align*}$$
    and then $s_jw=ws_{j^\prime }$ for some $j^\prime \in J$ (in particular, $\ell (s_jw)=\ell (w)+1$ ).

Proof. (1) See, for example, [Reference Bourbaki3].

(2) This follows from [Reference Deodhar16, Lem. 3.1 & Lem. 3.2].

The stabiliser subgroup $W_c$ of $c\in \overline {C}_+$ is parabolic,

$$\begin{align*}W_c=W_{\mathbf{J}(c)} \end{align*}$$

with $\mathbf {J}(c)\subseteq [0,r]$ given by

(2.15) $$ \begin{align} \mathbf{J}(c):=\{j\in [0,r] \,\, | \,\, s_jc=c\}. \end{align} $$

Definition 2.3. For $y\in E$ , we write

  1. 1. $c_y$ for the unique element in $\overline {C}_+$ such that $y\in Wc_y$ ,

  2. 2. $w_y$ for the unique element in $W^{\mathbf {J}(c_y)}$ such that $w_yc_y=y$ .

  3. 3. $\mu _y:=w_y(0)\in Q^{\vee }$ and $v_y:=w_y^{-1}\tau (\mu _y)\in W_0$ .

Fix $\nu \in Q^{\vee }$ . Then we have $c_\nu =0$ and $\mu _\nu =\nu $ . In particular, $w_\nu $ is the unique element in $\tau (\nu )W_0$ of minimal length and

(2.16) $$ \begin{align} \ell(w_\nu v)=\ell(w_\nu)+\ell(v)\qquad \forall\, v\in W_0. \end{align} $$

By [Reference Macdonald29, §2.4], $v_\nu \in W_0$ is the unique element of minimal length in $W_0$ such that $v_\nu \nu =\nu ^-$ , and by (2.13), we have

(2.17) $$ \begin{align} w_\nu=\tau(\nu)\qquad \forall\, \nu\in Q^{\vee}\cap\overline{E}_-. \end{align} $$

For $\nu =\varphi ^{\vee }$ , we have

$$\begin{align*}w_{\varphi^{\vee}}=s_0,\qquad\quad v_{\varphi^{\vee}}=s_\varphi, \end{align*}$$

and hence, $\ell (\tau (\varphi ^{\vee }))=1+\ell (v_{\varphi ^{\vee }})$ . In addition, we have

(2.18) $$ \begin{align} \begin{aligned} &\Pi(s_{\varphi})=\{\alpha\in\Phi^+_0 \,\, | \,\, \alpha(\varphi^{\vee})=1\}\cup\{\varphi\},\\ &\Phi^+_0\setminus\Pi(s_\varphi)=\{\alpha\in\Phi^+_0 \,\, | \,\, \alpha(\varphi^{\vee})=0\}. \end{aligned} \end{align} $$

For details see, for example, [Reference Macdonald29, §2.4].

The following proposition reduces to [Reference Cherednik7, Thm. 1.4] when $y\in Q^{\vee }$ .

Proposition 2.4. For $y\in E$ , we have

$$\begin{align*}\Pi(w_y^{-1})=\{a\in\Phi^+ \,\,\, | \,\,\, a(y)<0\}. \end{align*}$$

Proof. Let $a\in \Phi ^+$ with $a(y)<0$ . Then $(w_y^{-1}a)(c_y)=a(y)<0$ and $c_y\in \overline {C}_+$ , so $w_y^{-1}a\in \Phi ^-$ . In particular, $a\in \Pi (w_y^{-1})$ .

Conversely, if $a\in \Pi (w_y^{-1})$ , then $w_y^{-1}a\in \Phi ^-$ ; hence, $a(y)=(w_y^{-1}a)(c_y)\leq 0$ . If $a(y)=0$ , then $s_ay=y$ ; hence, $s_aw_y\in w_yW_{c_y}$ . But $w_y$ is the element of minimal length in $w_yW_{c_y}$ , so $\ell (s_aw_y)>\ell (w_y)$ , and hence, $\ell (w_y^{-1}s_a)>\ell (w_y^{-1})$ . This implies $w_y^{-1}a\in \Phi ^+$ by [Reference Humphreys21, Prop. 5.7], which contradicts the fact that $a\in \Pi (w_y^{-1})$ . Hence, $a(y)<0$ , which completes the proof.

2.4 The Coxeter complex of W

Let $\Sigma (\Phi )$ be the poset consisting of faces of the affine hyperplane arrangement $\{a^{-1}(0)\,\, | \,\, a\in \Phi \}$ in E. Concretely,

$$\begin{align*}\Sigma(\Phi)=\{wC^J\,\, | \,\, J\subsetneq [0,r]\,\, \& \,\, w\in W^J\} \end{align*}$$

with

$$\begin{align*}C^J:=\{c\in \overline{C}_+ \,\, | \,\, \alpha_j(c)=0\,\,\, (j\in J)\,\, \& \,\, \alpha_j(c)>0\,\,\, (j\in [0,r]\setminus J)\}. \end{align*}$$

Note that the faces are $E_{\text {co}}$ -translation invariant, $C^{[1,r]}=E_{\text {co}}$ and

(2.19) $$ \begin{align} E=\bigsqcup_{F\in\Sigma(\Phi)}F \end{align} $$

(disjoint union). The poset structure on $\Sigma (\Phi )$ is defined in terms of inclusion of closures of faces: $F\leq F^\prime $ if $\overline {F}\subseteq \overline {F}^\prime $ . The closure $C_J$ of $C^J$ is given by

(2.20) $$ \begin{align} C_J=\bigsqcup_{J\subseteq K\subsetneq [0,r]}C^K \end{align} $$

(disjoint union), so that $C^{J^\prime }\leq C^J$ iff $J\subseteq J^\prime \subsetneq [0,r]$ .

We obtain an abstract simplicial complex by adding $\emptyset $ to $\Sigma (\Phi )$ . It is isomorphic to the Coxeter complex

$$\begin{align*}\{wW_J\,\, | \,\, J\subseteq [0,r]\,\, \& \,\, w\in W\}, \end{align*}$$

partially ordered by opposite inclusion.

For a set X, write $\mathcal {F}_\Sigma (E,X)$ for the set of functions $E\rightarrow X$ which are constant on faces. It canonically identifies with the space $\mathcal {F}(\Sigma (\Phi ),X)$ of functions $\Sigma (\Phi )\rightarrow X$ . The following lemma provides a large class of examples when $X=\mathbb {R}$ .

Lemma 2.5. Suppose that $g:\mathbb {R}\rightarrow \mathbb {R}$ is a function such that $g|_{(m,m+1)}$ is constant for all $m\in \mathbb {Z}$ . Let $a\in \Phi $ . Then

(2.21) $$ \begin{align} y\mapsto g(a(y))\qquad (y\in E) \end{align} $$

lies in $\mathcal {F}_\Sigma (E,\mathbb {R})$ .

Proof. Let $a\in \Phi $ with $Da\in \Phi _0^+$ . It suffices to check that (2.21) is constant on $C^J$ ( $J\subsetneq [0,r]$ ). Fix $y\in C^J\subseteq \overline {C}_+$ . Note that $0\leq Da(y)\leq 1$ by (2.8).

Suppose that $a(y)\in \mathbb {Z}$ . If $Da(y)=0$ , then $Da\in \Phi _0^+\cap \Phi _J$ ; hence, $Da\in \text {span}_{\mathbb {R}}\{D\alpha _j\}_{j\in J}$ . If $Da(y)=1$ , then necessarily $\varphi (y)=1$ (i.e., $0\in J$ ). It again follows that $Da\in \text {span}\{D\alpha _j\}_{j\in J}$ . In particular, in both cases, $a\vert _{C^J}$ is constant.

If $a(y)\not \in \mathbb {Z}$ , then $Da\not \in \text {span}\{D\alpha _j\}_{j\in J}$ , and hence, $a(y^\prime )\not \in \mathbb {Z}$ for all $y^\prime \in C^J$ . Since $C^J$ is convex and a is an affine linear function on E, this forces $a\vert _{C^J}$ to take values in an interval $(m,m+1)$ for some $m\in \mathbb {Z}$ .

In case $X=W$ , we have the following example.

Lemma 2.6. The function $y\mapsto w_y$ lies in $\mathcal {F}_\Sigma (E,W)$ .

Proof. Fix a face $F=wC^J$ ( $J\subsetneq [0,r]$ , $w\in W^J$ ). For $y\in F$ , we have $c_y\in C^J$ ; hence, $w_y\in W^J$ . Furthermore, $F=w_yC^J$ since $y=w_yc_y$ . Consequently $w_y=w$ .

2.5 The double affine Weyl group

The dual affine root system is

$$\begin{align*}\Phi^{\vee}:=\{a^{\vee}\,\, | \,\, a\in\Phi\}\subset E\oplus\mathbb{R}, \end{align*}$$

with $a^{\vee }$ the affine coroot

(2.22) $$ \begin{align} a^{\vee}:=(\alpha^{\vee},2\ell/\|\alpha\|^2)\qquad\quad (a=(\alpha,\ell)\in\Phi). \end{align} $$

Note that $2/\|\alpha \|^2\in \mathbb {Z}_{>0}$ ( $\alpha \in \Phi _0$ ) by our convention on the length of the roots, and

$$\begin{align*}\alpha_0^{\vee}= m^2K-(\varphi^{\vee},0) \end{align*}$$

with $K:=(0,1)\in E\oplus \mathbb {Z}$ , in view of the convention that long roots have squared norm $2/m^2$ . Hence, $\Phi ^{\vee }$ generates the affine coroot lattice

$$\begin{align*}\widehat{Q}^{\vee}:=Q^{\vee}\oplus m^2\mathbb{Z}K\subset E\oplus\mathbb{R}, \end{align*}$$

and the simple coroots $\alpha _j^{\vee }$ ( $0\leq j\leq r$ ) form a basis of $\widehat {Q}^{\vee }$ .

The affine Weyl group W acts linearly on $E\oplus \mathbb {R}$ by

(2.23) $$ \begin{align} v\cdot(y,d):=(vy,d),\qquad \tau(\mu)\cdot (y,d)=(y,d-\langle \mu,y\rangle) \end{align} $$

for $v\in W_0$ , $\mu \in Q^{\vee }$ and $(y,d)\in E\oplus \mathbb {R}$ . We denote the action (2.23) with a dot to avoid confusion with the W-action on E by translations and reflections. Note that the linear isomorphism $E\oplus \mathbb {R}\overset {\sim }{\longrightarrow } E^*\oplus \mathbb {R}$ , $(y,d)\mapsto (\langle y,\cdot \rangle ,d)$ , intertwines the W-actions (2.3) and (2.23).

For $a\in \Phi $ and $w\in W$ , we have

(2.24) $$ \begin{align} s_a\cdot(y,d)=(y,d)-Da(y)a^{\vee},\qquad\quad w\cdot a^{\vee}=(wa)^{\vee}. \end{align} $$

In particular, $\Phi ^{\vee }$ and $\widehat {Q}^{\vee }$ are W-invariant. Explicitly, $w\cdot K=K$ for all $w\in W$ and

$$\begin{align*}v\cdot\mu=v\mu,\qquad \tau(\nu)\cdot \mu=\mu-\langle\nu,\mu\rangle K \end{align*}$$

for $v\in W_0$ and $\mu ,\nu \in Q^{\vee }$ .

Definition 2.7. The double affine Weyl group is the semidirect product

(2.25) $$ \begin{align} \mathbb{W}:=W\ltimes\widehat{Q}^{\vee}. \end{align} $$

The action (2.23) of W on $E\oplus \mathbb {R}$ then extends to a $\mathbb {W}$ -action by

(2.26) $$ \begin{align} (\mu+\ell K)\cdot (y,d):=(y+\mu,d+\ell) \end{align} $$

for $\mu \in Q^{\vee }$ , $\ell \in m^2\mathbb {Z}$ , $y\in E$ and $d\in \mathbb {R}$ (which no longer is a linear action). Note that $m^2K$ generates the center of $\mathbb {W}$ .

We next consider, for a special class of finitely generated abelian subgroups $\widehat {\Lambda }$ in $E\oplus \mathbb {R}$ containing $\widehat {Q}^{\vee }$ , the $\mathbf {F}$ -linear extension of the $\mathbb {W}$ -action (2.23) and (2.26) on $\widehat {\Lambda }$ to the group algebra $\mathcal {P}_{\widehat {\Lambda }}=\bigoplus _{\widehat {\lambda }\in \widehat {\Lambda }}\mathbf {F}x^{\widehat {\lambda }}$ of $\widehat {\Lambda }$ over $\mathbf {F}$ (see Subsection 1.7).

Definition 2.8. Denote by $\mathcal {L}$ the set of finitely generated abelian subgroups $\Lambda \subset E$ satisfying

(2.27) $$ \begin{align} Q^{\vee}\subseteq\Lambda\quad \&\quad \alpha(\Lambda)\subseteq\mathbb{Z}\quad \forall\, \alpha\in\Phi_0. \end{align} $$

Note that if $\Lambda \in \mathcal {L}$ , then L is $W_0$ -invariant. Furthermore, $\Lambda \cap E^\prime ,\text {pr}_{E^\prime }(\Lambda )\in \mathcal {L}$ and

$$\begin{align*}Q^{\vee}\subseteq \Lambda\cap E^\prime\subseteq\text{pr}_{E^\prime}(\Lambda)\subseteq P^{\vee}. \end{align*}$$

As a consequence, we have the following:

Corollary 2.9. For $\Lambda \in \mathcal {L}$ , we have a group homomorphism

(2.28) $$ \begin{align} j_\Lambda: T_{P^{\vee}}\rightarrow T_\Lambda,\qquad j_\Lambda(t)^\lambda:=t^{\text{pr}_{E^\prime}(\lambda)}\quad (\lambda\in\Lambda). \end{align} $$

The following proposition is useful for writing co-weights as $\mathbb {Q}$ -linear combination of co-roots. Recall that h is the Coxeter number of W.

Proposition 2.10. For all $y\in E$ , we have

(2.29) $$ \begin{align} \frac{1}{h}\sum_{\alpha\in\Phi_0^+}\alpha(y)\alpha^{\vee}=\text{pr}_{E^\prime}(y). \end{align} $$

In particular, $\text {pr}_{E^\prime }(\Lambda )\subseteq P^{\vee }\subseteq \frac {1}{h}Q^{\vee }$ for all $\Lambda \in \mathcal {L}$ .

Proof. It suffices to prove (2.29) for $y\in E^\prime $ . It then follows by applying Schur’s lemma to the irreducible $W_0$ -module $E^\prime $ and using the fact that $h=\#\Phi _0/r$ ; see [Reference Steinberg40, Thm. 1.1 & Lem. 7.2]. The second statement is immediate.

For $\Lambda \in \mathcal {L}$ , consider the lattice

$$\begin{align*}\widehat{\Lambda}:=\Lambda\oplus\mathbb{Z}K\subset E\oplus\mathbb{R} \end{align*}$$

containing $\widehat {Q}^{\vee }$ . Note that $\widehat {\Lambda }$ is $\mathbb {W}$ -invariant. Set

$$\begin{align*}\mathbf{q}:=x^K\in\mathcal{P}_{\widehat{\Lambda}}. \end{align*}$$

Then $\mathcal {P}_{\widehat {\Lambda }}$ can be alternatively viewed as the group algebra of $\Lambda $ over the Laurent polynomial ring $\mathbf {F}[\mathbf {q}^{\pm 1}]$ ,

$$\begin{align*}\mathcal{P}_{\widehat{\Lambda}}=\bigoplus_{\lambda\in\Lambda}\mathbf{F}[\mathbf{q}^{\pm 1}]x^\lambda. \end{align*}$$

We now immediately obtain the following result.

Proposition 2.11. Let $\Lambda \in \mathcal {L}$ . The $\mathbf {F}$ -linear extension of the $\mathbb {W}$ -action (2.23) and (2.26) on $\widehat {\Lambda }$ to $\mathcal {P}_{\widehat {\Lambda }}$ is $\mathbf {F}[\mathbf {q}^{\pm 1}]$ -linear. Furthermore, for $\lambda \in \Lambda $ , we have

(2.30) $$ \begin{align} \begin{aligned} \ell K(x^\lambda)&:=\mathbf{q}^\ell x^\lambda\,\,\, (\ell\in m^2\mathbb{Z}),\qquad\quad \mu(x^\lambda):=x^{\lambda+\mu}\qquad\,\, (\mu\in Q^{\vee}),\\ v(x^\lambda)&:=x^{v\lambda}\quad (v\in W_0),\qquad \tau(\nu)(x^\lambda):=\mathbf{q}^{-\langle\nu,\lambda\rangle}x^{\lambda}\quad (\nu\in Q^{\vee}), \end{aligned} \end{align} $$

and for $w\in W$ , $\widehat {\lambda }\in \widehat {\Lambda }$ and $a\in \Phi $ , we have

(2.31) $$ \begin{align} w(x^{\widehat{\lambda}})=x^{w\cdot\widehat{\lambda}},\qquad\quad w(x^{a^{\vee}})=x^{(wa)^{\vee}}. \end{align} $$

Without much loss of generality, one may absorb $\mathbf {q}$ into the ground field $\mathbf {F}$ (by replacing $\mathbf {F}$ by $\mathbf {F}(\mathbf {q})$ ). This is convenient for our purposes, since it allows to interpret (2.30) as an action by q-difference reflection operators on the $\mathbf {F}$ -torus $T_\Lambda $ . We discuss it in detail in the following subsection.

2.6 Algebras of q-difference reflection operators

Recall that we fixed a parameter $q\in \mathbf {F}^\times $ from the outset, which is assumed not to be a root of unity (see Subsection 1.7).

Let $\Lambda \in \mathcal {L}$ . The $\mathbb {W}$ -action on $\mathcal {P}_{\widehat {\Lambda }}$ from Proposition 2.11 descends to a $\mathbb {W}$ -action on the polynomial algebra $\mathcal {P}_\Lambda $ via the specialisation map

$$\begin{align*}\mathcal{P}_{\widehat{\Lambda}}\twoheadrightarrow\mathcal{P}_\Lambda, \qquad \sum_{\lambda\in\Lambda}d_\lambda(\mathbf{q})x^\lambda\mapsto \sum_{\lambda\in\Lambda}d_\lambda(q)x^\lambda\qquad\quad (d_\lambda(\mathbf{q})\in\mathbf{F}[\mathbf{q}^{\pm 1}]). \end{align*}$$

The resulting action of the affine Weyl group W on $\mathcal {P}_\Lambda $ is by algebra homomorphisms,

(2.32) $$ \begin{align} v(x^{\lambda}):=x^{v\lambda},\qquad \tau(\mu)(x^{\lambda}):=q^{-\langle\mu,\lambda\rangle}x^\lambda\qquad (v\in W_0,\, \mu\in Q^{\vee}), \end{align} $$

while $\widehat {Q}^{\vee }$ acts by $(\mu +\ell K)(x^\lambda )=q^\ell q^{\lambda +\mu }$ for $\mu \in Q^{\vee }$ and $\ell \in m^2\mathbb {Z}$ . The resulting action of the group algebra $\mathbf {F}[\mathbb {W}]$ of $\mathbb {W}$ over $\mathbf {F}$ on $\mathcal {P}_\Lambda $ descends to an action of the quotient algebra $\mathbf {F}[\mathbb {W}]/(m^2K-q^{m^2})$ .

Denote the smashed product algebra relative to the W-action (2.32) on $\mathcal {P}_\Lambda $ by $W\ltimes \mathcal {P}_\Lambda $ , and write $\mathcal {P}:=\mathcal {P}_{Q^{\vee }}$ . Then

(2.33) $$ \begin{align} \mathbf{F}[\mathbb{W}]/(m^2K-q^{m^2})\overset{\sim}{\longrightarrow}W\ltimes\mathcal{P} \end{align} $$

as $\mathbf {F}$ -algebras by $w\mapsto w$ ( $w\in W$ ) and $\mu \mapsto x^\mu $ ( $\mu \in Q^{\vee }$ ). Hence, $\mathcal {P}_\Lambda $ becomes a $W\ltimes \mathcal {P}$ -module.

Viewing $\mathcal {P}_\Lambda $ as the regular $\mathbf {F}$ -valued functions on $T_\Lambda $ , the action (2.32) can be interpreted as an action by q-difference reflection operators (i.e., it is the contragredient action of a W-action on $T_\Lambda $ by q-dilations and reflections). We will now describe this W-action on $T_\Lambda $ in detail and extend it to an appropriate inverse system of tori.

Let $\Lambda \in \mathcal {L}$ . For $\xi \in Q$ and $z\in \mathbf {F}^\times $ , we denote by $z^\xi \in T_{\Lambda }$ the multiplicative character

(2.34) $$ \begin{align} \lambda\mapsto z^{\xi(\lambda)}\qquad (\lambda\in\Lambda). \end{align} $$

Note that $z^\xi \vert _\Lambda =j_\Lambda (z^\xi \vert _{P^{\vee }})$ since $\xi (\lambda )=\xi (\text {pr}_{E^\prime }(\lambda ))$ for $\xi \in Q$ .

We have the following special case of this construction. Our assumption on the normalisation of the root system $\Phi _0$ allows to view $Q^{\vee }$ as a sub-lattice of Q via the $W_0$ -equivariant isomorphism

$$\begin{align*}E\overset{\sim}{\longrightarrow} E^*,\qquad y\mapsto \langle y,\cdot\rangle. \end{align*}$$

More precisely, since $\|\varphi \|^2=2/m^2$ ( $m\in \mathbb {Z}_{>0}$ ), the co-root lattice $Q^{\vee }$ identifies with a sub-lattice of $mQ$ . Then we obtain the multiplicative characters $z^\mu \in T_{\Lambda }$ ( $\mu \in Q^{\vee }$ ) satisfying $\lambda \mapsto z^{\langle \lambda ,\mu \rangle }$ ( $\lambda \in \Lambda $ ). In particular, for $\mu \in Q^{\vee }$ , we have the multiplicative character $q^\mu \in T_{\Lambda }$ mapping $\lambda \in \Lambda $ to $q^{\langle \mu ,\lambda \rangle }$ . Note that $vz^\xi =z^{v\xi }$ and $vz^\mu =z^{v\mu }$ for $v\in W_0$ , $\xi \in Q$ and $\mu \in Q^{\vee }$ .

Fix $\Lambda \in \mathcal {L}$ . Since $\Lambda $ is $W_0$ -invariant, $W_0$ acts by group automorphisms on $T_\Lambda $ via the contragredient action on $T_\Lambda $ . Extend the $W_0$ -action on $T_\Lambda $ to a W-action by

(2.35) $$ \begin{align} \tau(\mu)t:=q^\mu t\qquad (t\in T_\Lambda,\, \mu\in Q^{\vee}). \end{align} $$

Note that the map (2.28) is W-equivariant, and

$$\begin{align*}(w(p))(t)=p(w^{-1}t)\qquad\quad (w\in W,\, p\in\mathcal{P}_\Lambda,\, t\in T_\Lambda), \end{align*}$$

with the W-action on $\mathcal {P}_\Lambda $ given by (2.32).

Write

$$\begin{align*}q_\alpha:=q^{\frac{2}{\|\alpha\|^2}}\qquad\quad (\alpha\in\Phi_0), \end{align*}$$

which are integral powers of q by the convention that $\|\varphi \|^2=2/m^2$ for some $m\in \mathbb {Z}_{>0}$ . For an affine root $a=(\alpha ,\ell )\in \Phi $ and $\lambda \in \Lambda $ , we have

(2.36) $$ \begin{align} s_a(x^\lambda)=q_\alpha^{-\ell\alpha(\lambda)}x^{s_\alpha \lambda}. \end{align} $$

From now on, we will be working with specialised $\mathbf {q}$ , unless explicitly stated otherwise. In particular,

$$ \begin{align*} x^{\lambda+\ell K}=q^\ell x^\lambda\in\mathcal{P}_\Lambda\qquad\quad (\lambda\in\Lambda,\, m\in\mathbb{Z}), \end{align*} $$

and (2.31) is interpreted accordingly as identity in $\mathcal {P}_\Lambda $ . Note that for $a\in \Phi $ and $t\in T_\Lambda $ , we have

(2.37) $$ \begin{align} s_at=(t^{-a^{\vee}})^{Da}t, \end{align} $$

where $t^{-a^{\vee }}\in \mathbf {F}^\times $ is the evaluation of $x^{-a^{\vee }}\in \mathcal {P}_\Lambda $ at t, and $(t^{-a^{\vee }})^{Da}\in T_\Lambda $ is the resulting multiplicative character of $\Lambda $ .

Let $\mathcal {Q}_\Lambda $ be the quotient field of $\mathcal {P}_\Lambda $ , which identifies with the field of rational functions on $T_\Lambda $ , and set $\mathcal {Q}:=\mathcal {Q}_{Q^{\vee }}$ . We extend the W-action on $\mathcal {P}_\Lambda $ to a W-action on $\mathcal {Q}_\Lambda $ by field automorphisms, and write $W\ltimes \mathcal {Q}_\Lambda $ for the associated smashed product algebra over $\mathbf {F}$ .

2.7 The double affine Hecke algebra

The double affine Hecke algebra $\mathbb {H}=\mathbb {H}(\mathbf {k},q)$ is a flat deformation of $\mathbf {F}[\mathbb {W}]/(m^2K-q^{m^2})\simeq W\ltimes \mathcal {P}$ , with its deformation parameters encoded by a multiplicity function $\mathbf {k}: \Phi \rightarrow \mathbf {F}^\times $ . In this paper, we restrict attention to multiplicity functions that only depend on the length of the gradient of the affine root.

Write ${}^{\text {sh}}\Phi _0$ and ${}^{\text {lg}}\Phi _0$ for the short and long roots in $\Phi _0$ , respectively. Fix

$$\begin{align*}{}^{\text{sh}}k,{}^{\text{lg}}k\in\mathbf{F}^\times, \end{align*}$$

and assume that both are admitting square roots in $\mathbf {F}$ , which we fix once and for all. If all the roots have the same length, then we will view the roots as long roots, and we will denote ${}^{\text {lg}}k$ by k. We write for $a\in \Phi $ ,

$$\begin{align*}k_a:={}^sk \quad \mbox{ if }\quad Da\in {}^s\Phi_0\qquad (s\in\{\text{sh},\text{lg}\}). \end{align*}$$

Then $\mathbf {k}: \Phi \rightarrow \mathbf {F}^\times $ , $a\mapsto k_a$ is W-invariant map (it is in fact invariant for the action of the extended affine Weyl group; see Section 7). We write $k_j:=k_{\alpha _j}$ for $0\leq j\leq r$ . Then $k_j=k_{j^\prime }$ iff $Da_{j^\prime }$ lies in the $W_0$ -orbit of $Da_j$ . In particular, $k_0=k_\varphi $ .

Cherednik’s [Reference Cherednik8] equal lattice double affine Hecke algebra is the following deformation of $W\ltimes \mathcal {P}$ .

Definition 2.12. The double affine Hecke algebra $\mathbb {H}=\mathbb {H}(\mathbf {k},q)$ is the unital associative $\mathbf {F}$ -algebra generated by $T_0,\ldots ,T_r$ and $x^\mu $ ( $\mu \in Q^{\vee }$ ), subject to the following relations:

  • a. The braid relations

    (2.38) $$ \begin{align} T_jT_{j^\prime}T_j\ldots=T_{j^\prime}T_jT_{j^\prime}\cdots \qquad ( m_{jj^\prime} \text{ factors on each side}) \end{align} $$

    for $0\leq j\not =j^\prime \leq r$ , with $m_{jj^\prime }$ the order of $s_js_{j^\prime }$ in W,

    b. The Hecke relations

    (2.39) $$ \begin{align} (T_j-k_j)(T_j+k_j^{-1})=0\qquad\quad (0\leq j\leq r), \end{align} $$

    c.  $x^\lambda x^\mu =x^{\lambda +\mu }$ ( $\lambda ,\mu \in Q^{\vee }$ ) and $x^0=1$ ,

    d.The cross relations

    (2.40) $$ \begin{align} T_jx^\mu-s_j(x^\mu)T_j=(k_j-k_j^{-1})\left(\frac{x^\mu-s_j(x^\mu)}{1-x^{\alpha_j^{\vee}}}\right) \end{align} $$

    for $0\leq j\leq r$ and $\mu \in Q^{\vee }$ , with the W-action (2.32) on $\mathcal {P}$ .

Note that $T_j\in \mathbb {H}^\times $ with inverse $T_j^{-1}=T_j-k_j+k_j^{-1}$ by (2.39). Furthermore, note that

(2.41) $$ \begin{align} \mathbb{H}(\mathbf{1},q)\overset{\sim}{\longrightarrow}W\ltimes\mathcal{P} \end{align} $$

as algebras, with $\mathbf {1}$ the multiplicity function identically equal to $1$ and the isomorphism given by $T_w\mapsto w$ ( $w\in W$ ) and $x^\mu \mapsto x^\mu $ ( $\mu \in Q^{\vee }$ ).

Remark 2.13. The cross relations (2.40) in $\mathbb {H}$ can alternatively be written as

(2.42) $$ \begin{align} \begin{aligned} T_ix^\mu-x^{s_i\mu}T_i&=(k_i-k_i^{-1})\left(\frac{1-x^{-\alpha_i(\mu)\alpha_i^{\vee}}}{1-x^{\alpha_i^{\vee}}}\right)x^\mu\qquad (1\leq i\leq r),\\ T_0x^\mu-q_{\varphi}^{\varphi(\mu)}x^{s_\varphi\mu}T_0&=(k_0-k_0^{-1})\left(\frac{1-(q_\varphi x^{-\varphi^{\vee}})^{\varphi(\mu)}}{1-q_\varphi x^{-\varphi^{\vee}}}\right)x^\mu \end{aligned} \end{align} $$

in view of (2.36).

For $w\in W$ with reduced expression $w=s_{j_1}\cdots s_{j_\ell }$ ( $0\leq j_i\leq \ell $ ), we write

$$ \begin{align*} T_w:=T_{j_1}\cdots T_{j_\ell}\in\mathbb{H}, \end{align*} $$

which is independent of the choice of reduced expression, due to the braid relations (2.38). It follows from (2.13) that there exists a unique group homomorphism

$$\begin{align*}Q^{\vee}\rightarrow \mathbb{H}^\times,\qquad \mu\mapsto Y^\mu \end{align*}$$

such that

(2.43) $$ \begin{align} Y^\mu=T_{\tau(\mu)}\qquad \forall\,\mu\in Q^{\vee}\cap\overline{E}_+. \end{align} $$

Let

$$\begin{align*}\chi_\pm: \Phi_0\rightarrow\{0,1\} \end{align*}$$

be the characteristic function of $\Phi _0^\pm $ . Then we have

(2.44) $$ \begin{align} Y^{v^{-1}\varphi^{\vee}}=T_v^{-1}T_0^{\chi(v^{-1}\varphi)}T_{s_\varphi v}\qquad (v\in W_0), \end{align} $$

where

(2.45) $$ \begin{align} \chi:=\chi_+-\chi_- \end{align} $$

(see, for example, [Reference Macdonald29, (3.3.6)]). In particular,

$$\begin{align*}Y^{\varphi^{\vee}}=T_0T_{s_\varphi}. \end{align*}$$

The Poincaré-Birkhoff-Witt (PBW) theorem [Reference Cherednik8, Thm. 3.2.1] for $\mathbb {H}$ states that

$$\begin{align*}\{x^\mu T_vY^\nu\,\, | \,\, v\in W_0\,\,\&\,\, \mu,\nu\in Q^{\vee}\} \end{align*}$$

as well as $\{x^\mu T_w\,\, | \,\, \mu \in Q^{\vee }, w\in W\}$ are linear $\mathbf {F}$ -linear bases of $\mathbb {H}$ . In particular, the subalgebra of $\mathbb {H}$ spanned by $x^\mu $ ( $\mu \in Q^{\vee }$ ) is isomorphic to $\mathcal {P}$ , as is the subalgebra $\mathcal {P}_Y$ of $\mathbb {H}$ spanned by $Y^\mu $ ( $\mu \in Q^{\vee }$ ). For $p=\sum _{\mu \in Q^{\vee }}d_\mu x^\mu \in \mathcal {P}$ , we will write

$$\begin{align*}p(Y^{\pm 1}):=\sum_{\mu\in Q^{\vee}}d_\mu Y^{\pm\mu}\in\mathcal{P}_Y. \end{align*}$$

The finite Hecke algebra $H_0=H_0(\mathbf {k})$ is the subalgebra of $\mathbb {H}$ generated by $T_1,\ldots ,T_r$ . The finite Hecke algebra $H_0$ has $\{T_v\,\, | \,\, v\in W_0\}$ as a $\mathbf {F}$ -linear basis. The defining relations in terms of $T_1,\ldots ,T_r$ are the Hecke relations (2.39) and the braid relations (2.38) for $1\leq j\not =j^\prime \leq r$ .

The affine Hecke algebra $H=H(\mathbf {k})$ is the subalgebra of $\mathbb {H}$ generated by $T_0,\ldots ,T_r$ . The affine Hecke algebra H has $\{T_w\,\, | \,\, w\in W\}$ as a $\mathbf {F}$ -linear basis. The defining relations in terms of the generators $T_0,\ldots ,T_r$ are the Hecke relations and the braid relations. Another $\mathbf {F}$ -linear basis of the affine Hecke algebra H is $\{T_vY^\mu \,\, | \,\, v\in W_0\,\, \&\,\, \mu \in Q^{\vee }\}$ . The defining relations of H in terms of the subalgebra $H_0$ and $\mathcal {P}_Y$ are the Bernstein-Zelevinsky cross relations

(2.46) $$ \begin{align} Y^\mu T_i-T_iY^{s_i\mu}=(k_i-k_i^{-1})\left(\frac{Y^\mu-Y^{s_i\mu}} {1-Y^{-\alpha_i^{\vee}}}\right) \end{align} $$

for $i=1,\ldots ,r$ and $\mu \in Q^{\vee }$ ; see [Reference Lusztig25].

The following theorem is due to Cherednik; see, for example, [Reference Cherednik8, §3.3.2].

Theorem 2.14. There exists a unique anti-algebra involution $\delta : \mathbb {H}\rightarrow \mathbb {H}$ satisfying $\delta (Y^\mu )=x^{-\mu }$ ( $\mu \in Q^{\vee }$ ) and $\delta (T_i)=T_i$ ( $1\leq i\leq r$ ).

We call $\delta $ the duality anti-involution of $\mathbb {H}$ . Note that

(2.47) $$ \begin{align} \delta(T_0)=T_{s_\varphi}^{-1}x^{-\varphi^{\vee}}. \end{align} $$

It is sometimes convenient to rewrite $\delta (T_0)$ in terms of the element

(2.48) $$ \begin{align} U_0:=q_\varphi^{-1}x^{\varphi^{\vee}}T_0^{-1}=q_\varphi^{-1}x^{\varphi^{\vee}} T_{s_\varphi}Y^{-\varphi^{\vee}}\in\mathbb{H}, \end{align} $$

which satisfies

(2.49) $$ \begin{align} \delta(T_0^{-1})=q_\varphi U_0Y^{\varphi^{\vee}},\qquad \delta(T_0)= Y^{-\varphi^{\vee}}T_0x^{-\varphi^{\vee}} \end{align} $$

and the Hecke relation $(U_0-k_0)(U_0+k_0^{-1})=0$ . The outer automorphism of $\mathbb {H}$ , defined as conjugation by the Gaussian in the polynomial realisation of $\mathbb {H}$ , maps $T_0$ to a scalar multiple of $U_0^{-1}$ (see, for example, [Reference Cherednik9]).

2.8 Intertwiners

We recall in this subsection well-known facts about Cheredniki’s X-intertwiners and Y-intertwiners for the double affine Hecke algebra $\mathbb {H}$ ; see, for example, [Reference Cherednik8, §3.3.3]. For $0\leq j\leq r$ , write

(2.50) $$ \begin{align} S_j^X:=\big(x^{\alpha_j^{\vee}}-1\big)T_j+k_j-k_j^{-1}\in\mathbb{H}. \end{align} $$

The following theorem introduces the X-intertwiners $S_w^X\in \mathbb {H}$ ( $w\in W$ ).

Theorem 2.15. For $w\in W$ , set

$$\begin{align*}S_w^X:=S_{j_1}^X\cdots S_{j_\ell}^X\in\mathbb{H}, \end{align*}$$

where $w=s_{j_1}\cdots s_{j_\ell }$ ( $0\leq j_i\leq r$ ) is a reduced expression. Then $S_w^X$ is well defined and

(2.51) $$ \begin{align} \begin{aligned} S_w^Xp&=w(p)S_w^X,\\ S_{w^{-1}}^XS_{w}^X&=\prod_{a\in\Pi(w)}\big(k_a^{-1}-k_ax^{a^{\vee}}\big)\big(k_a^{-1}-k_ax^{-a^{\vee}}\big) \end{aligned} \end{align} $$

in $\mathbb {H}$ for $p\in \mathcal {P}$ and $w\in W$ .

Let $\mathbb {H}^{X-\text {loc}}$ be the ring of fractions of $\mathbb {H}$ with respect to the multiplicative set $\mathcal {P}^\times $ . The canonical algebra embedding $\mathcal {P}\hookrightarrow \mathbb {H}^{X-\text {loc}}$ extends to an algebra embedding $\mathcal {Q}\hookrightarrow \mathbb {H}^{X-\text {loc}}$ . Multiplication defines a $\mathbf {F}$ -linear isomorphism

(2.52) $$ \begin{align} \mathcal{Q}\otimes_{\mathbf{F}}H\overset{\sim}{\longrightarrow}\mathbb{H}^{X-\text{loc}}. \end{align} $$

The cross relations in $\mathbb {H}^{X-\text {loc}}$ are

(2.53) $$ \begin{align} T_jf-s_j(f)T_j=(k_j-k_j^{-1})\left(\frac{f-s_j(f)}{1-x^{\alpha_j^{\vee}}}\right) \end{align} $$

for $0\leq j\leq r$ and $f\in \mathcal {Q}$ . Set

(2.54) $$ \begin{align} \widetilde{S}_j^X:=\frac{1}{k_jx^{\alpha_j^{\vee}}-k_j^{-1}}S_j^X= \left(\frac{x^{\alpha_j^{\vee}}-1}{k_jx^{\alpha_j^{\vee}}-k_j^{-1}}\right)T_j+ \left(\frac{k_j-k_j^{-1}}{k_jx^{\alpha_j^{\vee}}-k_j^{-1}}\right)\in\mathbb{H}^{X-\text{loc}} \end{align} $$

for $0\leq j\leq r$ . Then $(S_j^X)^2=1$ for $j=0,\ldots ,r$ . The normalised X-intertwiners $\widetilde {S}_w^X\in \mathbb {H}^{X-\text {loc}}$ ( $w\in W$ ) are defined by

$$\begin{align*}\widetilde{S}_w^X:=\widetilde{S}_{j_1}^X\cdots\widetilde{S}_{j_\ell}^X, \end{align*}$$

where $w=s_{j_1}\cdots s_{j_\ell }$ may now be any expression of $w\in W$ as product of simple reflections. Note that

(2.55) $$ \begin{align} \widetilde{S}_w^X=S_w^X\prod_{a\in\Pi(w)}\left(\frac{1}{k_ax^{-a^{\vee}}-k_a^{-1}}\right). \end{align} $$

The following result is from [Reference Stokman41, §3.1].

Theorem 2.16. The assignment

$$\begin{align*}w\otimes f\mapsto\widetilde{S}_w^Xf\qquad (w\in W, f\in\mathcal{Q}) \end{align*}$$

defines an isomorphism $\beth : W\ltimes \mathcal {Q}\overset {\sim }{\longrightarrow }\mathbb {H}^{X-\text {loc}}$ of algebras.

Write $\mathbb {H}^{Y-\text {loc}}$ for the ring of fractions of $\mathbb {H}$ with respect to the multiplicative set $\mathcal {P}_Y^\times $ . The duality anti-involution $\delta $ extends to an algebra anti-isomorphism $\mathbb {H}^{X-\text {loc}}\rightarrow \mathbb {H}^{Y-\text {loc}}$ , which we again denote by $\delta $ . The Y-intertwiners are defined by

$$\begin{align*}S_j^Y:=\delta(S_j^X),\qquad S_{w}^Y:=\delta(S_{w^{-1}}^X), \end{align*}$$

and the normalised versions by $\widetilde {S}_j^Y:=\delta (\widetilde {S}_j^X)$ and $\widetilde {S}_w^Y:=\delta (\widetilde {S}_{w^{-1}}^X)$ for $0\leq j\leq r$ and $w\in W$ . Note that $S_w^Y\in \mathbb {H}$ and $\widetilde {S}_w^Y\in \mathbb {H}^{Y-\text {loc}}$ . Furthermore,

$$\begin{align*}S_w^Y=S_{j_1}^YS_{j_2}^Y\cdots S_{j_\ell}^Y \end{align*}$$

if $w=s_{j_1}\cdots s_{j_\ell }$ is a reduced expression of $w\in W$ . The same holds true for $\widetilde {S}_w^Y$ (in this case, it holds true for any expression $w=s_{j_1}\cdots s_{j_\ell }$ as product of simple reflections). For later reference, we write out explicitly some of the key formulas for the Y-intertwiners.

For $0\leq j\leq r$ , we have

(2.56) $$ \begin{align} S_j^Y=\delta(T_j)\big((Y^{-1})^{\alpha_j^{\vee}}-1\big)+k_j-k_j^{-1}. \end{align} $$

Note here that for $j=0$ , we are evaluating the polynomial $x^{\alpha _0^{\vee }}=q_\varphi x^{-\varphi ^{\vee }}$ in $Y^{-1}$ , so that $(Y^{-1})^{\alpha _0^{\vee }}=q_\varphi Y^{\varphi ^{\vee }}$ (which is not equal to $Y^{-\alpha _0^{\vee }}= q_\varphi ^{-1}Y^{\varphi ^{\vee }}$ ). The Hecke relations (2.39) provide the alternative expression

(2.57) $$ \begin{align} S_j^Y=\delta(T_j^{-1})\big((Y^{-1})^{\alpha_j^{\vee}}-1\big)+(k_j-k_j^{-1})(Y^{-1})^{\alpha_j^{\vee}}. \end{align} $$

An expression for $S_0$ in terms of $U_0$ is

(2.58) $$ \begin{align} S_0^Y=\big(U_0(q_\varphi Y^{\varphi^{\vee}}-1)+k_0-k_0^{-1}\big)q_\varphi Y^{\varphi^{\vee}}. \end{align} $$

For $p\in \mathcal {P}$ and $w\in W$ , we have

(2.59) $$ \begin{align} \begin{aligned} S_w^Yp(Y^{-1})&=w(p)(Y^{-1})S_w^Y,\\ S_{w^{-1}}^YS_{w}^Y&=\prod_{a\in\Pi(w)}\big(k_a^{-1}-k_a(Y^{-1})^{a^{\vee}}\big)\big(k_a^{-1}-k_a(Y^{-1})^{-a^{\vee}}\big) \end{aligned} \end{align} $$

in $\mathbb {H}$ . Finally, we have

(2.60) $$ \begin{align} \widetilde{S}_w^Y=S_w^Y\prod_{a\in\Pi(w)}\left(\frac{1}{k_a(Y^{-1})^{a^{\vee}}-k_a^{-1}}\right) \end{align} $$

for $w\in W$ .

3 Y-parabolically induced cyclic $\mathbb {H}$ -modules

Throughout this section, we fix a lattice $\Lambda \in \mathcal {L}$ . As before (see Subsection 2.6), we will omit $\Lambda $ from the notations when $\Lambda =Q^{\vee }$ .

3.1 Induction parameters

Definition 3.1. For $J\subsetneq [0,r]$ , write

(3.1) $$ \begin{align} \begin{aligned} T_{\Lambda,J}&:=\{\mathfrak{t}\in T_\Lambda \,\, | \,\, \mathfrak{t}^{\alpha_j^{\vee}}=1\quad\,\,\, \forall\, j\in J\},\\ T^{\text{red}}_{\Lambda,J}&:=\{\mathfrak{t}\in T_\Lambda\,\, | \,\, \mathfrak{t}^{D\alpha_j^{\vee}}=1\quad \forall\, j\in J\}. \end{aligned} \end{align} $$

Note that $T^{\text {red}}_{\Lambda ,J}\subseteq T_\Lambda $ is a subtorus. Furthermore, $T_{\Lambda ,I}=T^{\text {red}}_{\Lambda ,I}$ if $I\subseteq [1,r]$ , and $T_{[1,r]}=T_{[1,r]}^{\text {red}}=\{1_T\}$ with $1_T$ the neutral element of $T=T_{Q^{\vee }}$ . If $T_{\Lambda ,J}\not =\emptyset $ , then $T_{\Lambda ,J}$ is a $T^{\text {red}}_{\Lambda ,J}$ -coset since

$$\begin{align*}(tt^\prime)^{a^{\vee}}=t^{a^{\vee}}t^\prime{}^{Da^{\vee}} \end{align*}$$

for $t,t^\prime \in T_\Lambda $ and $a\in \Phi $ . In the following lemma, we provide a natural, mild condition on $q\in \mathbf {F}$ ensuring that $T_J\not =\emptyset $ for all $J\subsetneq [0,r]$ .

Let h be the Coxeter number of $W_0$ – that is, h is the order of $s_1s_2\cdots s_r$ in $W_0$ (see, for example, [Reference Humphreys21, Chpt. 3] and [Reference Steinberg40]).

Lemma 3.2. $T_J\not =\emptyset $ for all $J\subsetneq [0,r]$ if q has a $(2h)^{th}$ root in $\mathbf {F}$ .

Proof. It $I\subseteq [1,r]$ , then $T_I\not =\emptyset $ without restrictions on q since $T_I$ contains $1_T$ .

Fix now a subset $J\subsetneq [0,r]$ containing $0$ . Write $J_0:=[1,r]\cap J$ , which is a proper subset of $[1,r]$ . For $\mathfrak {t}\in T$ , write $c_i:=\mathfrak {t}^{\alpha _i^{\vee }}$ ( $i\in [1,r]$ ). Then $\mathfrak {t}\in T_J$ iff $c_i=1$ for $i\in J_0$ and

(3.2) $$ \begin{align} \prod_{i\in [1,r]\setminus J_0}c_i^{m_i}=q_\varphi, \end{align} $$

where $m_i\in \mathbb {Z}_{>0}$ are the coefficients in the expansion $\varphi ^{\vee }=\sum _{i=1}^rm_i\alpha _i^{\vee }$ of $\varphi ^{\vee }$ in simple co-roots. It thus remains to show that there exists $c_i\in \mathbf {F}$ ( $i\in [1,r]\setminus J_0$ ) satisfying (3.2) when q has a $(2h)^{th}$ root in $\mathbf {F}$ .

It is clear from (3.2) that such $c_i$ exists if $q_\varphi $ has a $m_\ell ^{th}$ root in $\mathbf {F}$ for all $\ell \in [1,r]$ . Since $q_\varphi $ is an integer power of q (due to our choice of normalisation of the root lengths; cf. Subsection 2.1), this is the case when q has a $m(\varphi )^{th}$ root in $\mathbf {F}$ , where

$$\begin{align*}m(\varphi):=\text{lcm}(m_i). \end{align*}$$

By direct inspection using Table 2 of [Reference Humphreys20, §12], we have $m(\varphi )=1$ for types $\text {A}_r$ and $\text {B}_r$ , $m(\varphi )=2$ for types $\text {C}_r$ , $\text {D}_r$ and $\text {G}_2$ , $m(\varphi )=6$ for types $\text {E}_6$ and $\text {F}_4$ , $m(\varphi )=12$ for type $\text {E}_7$ and $m(\varphi )=60$ for type $\text {E}_8$ . Comparing with the Coxeter numbers, which are explicitly listed in [Reference Humphreys21, §3.18, Table 2], one checks by direct inspection that

(3.3) $$ \begin{align} m(\varphi)\vert 2h. \end{align} $$

Hence, $T_J\not =\emptyset $ if q has a $(2h)^{th}$ root in $\mathbf {F}$ .

Let $W^{\text {red}}_J\subseteq W_0$ be the image of the parabolic subgroup $W_J\subseteq W$ under the group epimorphism $D: W\twoheadrightarrow W_0$ . We then have

(3.4) $$ \begin{align} T_{\Lambda,J}\subseteq T_\Lambda^{W_J},\qquad T^{\text{red}}_{\Lambda,J}\subseteq T_\Lambda^{W^{\text{red}}_J}. \end{align} $$

Here, $T^G_\Lambda \subseteq T_\Lambda $ , for a subgroup $G\subseteq W$ , denotes the set of G-fixed elements in $T_\Lambda $ . By (2.37), the inclusions (3.4) are equalities if $D\alpha _j(\Lambda )=\mathbb {Z}$ for all $j\in J$ .

Note that $T_{\Lambda ,J}$ is a $T_{\Lambda ,[1,r]}$ -subset of $T_\Lambda $ for all $J\subsetneq [0,r]$ , since $T_{\Lambda ,[1,r]}=T_{\Lambda ,[1,r]}^{\text {red}}$ . Write

(3.5) $$ \begin{align} \widetilde{T}_\Lambda:=\{\widetilde{s}\in T_{\Lambda} \,\, | \,\, \widetilde{s}\vert_{\Lambda\cap E_{\text{co}}}\equiv 1\},\qquad \widetilde{T}_{\Lambda,J}:=T_{\Lambda,J}\cap\widetilde{T}_\Lambda, \end{align} $$

which are $\widetilde {T}_{\Lambda ,[1,r]}$ -subsets of $T_\Lambda $ . The following lemma, which explains how characters $t\in T_J$ can be lifted to $T_\Lambda $ , will be useful when we discuss the quasi-polynomial representation for extended double affine Hecke algebras in Section 7.

Lemma 3.3. Let $J\subsetneq [0,r]$ .

  1. 1. We have injective maps

    $$\begin{align*}T_{\Lambda,J}/T_{\Lambda,[1,r]}\hookrightarrow T_J,\qquad \widetilde{T}_{\Lambda,J}/\widetilde{T}_{\Lambda,[1,r]}\hookrightarrow T_J \end{align*}$$
    defined by $sT_{\Lambda ,[1,r]}\mapsto s\vert _{Q^{\vee }}$ and $s\widetilde {T}_{\Lambda ,[1,r]}\mapsto s\vert _{Q^{\vee }}$ , respectively.
  2. 2. The maps in (1) are bijective if the restriction map $T_{\text {pr}_{E^\prime }(\Lambda ),J}\rightarrow T_J$ , $s\mapsto s\vert _{Q^{\vee }}$ is surjective.

The condition in (2) is, for instance, met when $\mathbf {F}$ is algebraically closed.

Proof. (1) This is clear.

(2) For $s\in T_{\text {pr}_{E^\prime }(\Lambda )}$ , define $\widetilde {s}\in T_\Lambda $ by

(3.6) $$ \begin{align} \widetilde{s}^{\,\lambda}:=s^{\,\text{pr}_{E^\prime}(\lambda)}\qquad\quad (\lambda\in\Lambda). \end{align} $$

It gives rise to a bijective map

(3.7) $$ \begin{align} T_{\text{pr}_{E^\prime}(\Lambda),J}\overset{\sim}{\longrightarrow}\widetilde{T}_{\Lambda,J},\qquad s\mapsto \widetilde{s} \end{align} $$

satisfying $s\vert _{Q^{\vee }}=\widetilde {s}\vert _{Q^{\vee }}$ for all $s\in T_{\text {pr}_{E^\prime }(\Lambda ),J}$ ; cf. Corollary 2.9. Combined with the assumption that the restriction map $T_{\text {pr}_{E^\prime }(\Lambda ),J}\rightarrow T_J$ , $s\mapsto s\vert _{Q^{\vee }}$ is surjective, we conclude that the maps in (1) are surjective.

The following subsets of $T_\Lambda $ will serve as induction parameters for a class of cyclic double affine Hecke algebra modules, to be defined shortly.

Definition 3.4. For $J\subsetneq [0,r]$ , we set

(3.8) $$ \begin{align} L_{\Lambda,J}:=\{t\in T_\Lambda \,\, | \,\, t^{\alpha_j^{\vee}}=k_j^{-2}\quad\,\,\,\, \forall\, j\in J\}. \end{align} $$

Note that the condition $t^{\alpha _0^{\vee }}=k_0^{-2}$ reduces to $t^{\varphi ^{\vee }}=q_\varphi k_\varphi ^2$ .

We next construct an explicit base element $\mathfrak {s}_J\in T_{P^{\vee }}$ such that

$$\begin{align*}L_{\Lambda,J}=\mathfrak{s}_{J} T_{\Lambda,J} \end{align*}$$

in $T_\Lambda $ , where $\mathfrak {s}_{J}$ is regarded as multiplicative character of $\Lambda $ via Corollary 2.9.

The image $\alpha (C^J)$ of the restriction of the root $\alpha \in \Phi _0^+$ to the face $C^J$ is either $\{0\}$ , $\{1\}$ , or it is contained in the open interval $(0,1)$ (see the proof of Lemma 2.5). We record these three cases by the indicator function

(3.9) $$ \begin{align} \eta_J(\alpha):= \begin{cases} -1\qquad&\mbox{if }\, \alpha(C^J)=\{0\},\\ 0\qquad&\mbox{if }\, \alpha(C^J)\subseteq (0,1),\\ 1\qquad&\mbox{if }\, \alpha(C^J)=\{1\}. \end{cases} \end{align} $$

It can be alternatively described as follows. Set

(3.10) $$ \begin{align} J_0:=[1,r]\cap J, \end{align} $$

and denote by $\Phi _{0,J_0}\subseteq \Phi _0$ the parabolic root subsystem with basis $\{\alpha _i\}_{i\in J_0}$ . Write $\Phi _{0,J_0}^+:=\Phi _0^+\cap \Phi _{0,J_0}$ . For $\alpha \in \Phi _0^+$ , denote by $n_i(\alpha )$ ( $1\leq i\leq r$ ) the nonnegative integers such that $\alpha =\sum _{i=1}^rn_i(\alpha )\alpha _i$ .

Lemma 3.5. For $J\subsetneq [0,r]$ , we have

$$ \begin{align*} \eta_J(\alpha)= \begin{cases} -1\qquad &\mbox{if }\, \alpha\in\Phi_{0,J_0}^+,\\ 0\qquad &\mbox{if }\, \alpha\in\Phi_0^+\setminus (\Phi_0^{+,J}\cup\Phi_{0,J_0}^+),\\ 1\qquad &\mbox{if }\, \alpha\in\Phi_0^{+,J}, \end{cases} \end{align*} $$

where

(3.11) $$ \begin{align} \Phi_0^{+,J}:= \begin{cases} \{\alpha\in\Phi_0^+ \,\, | \,\, n_i(\alpha)=n_i(\varphi)\,\,\,\forall\, i\in [1,r]\setminus J_0\} \qquad &\mbox{ if }\, 0\in J,\\ \emptyset \qquad &\mbox{ if }\, 0\not\in J. \end{cases} \end{align} $$

Proof. This is similar to the proof of Lemma 2.5.

Definition 3.6. For $\Lambda \in \mathcal {L}$ and $J\subsetneq [0,r]$ , define the base-point $\mathfrak {s}_J\in T_{\Lambda }$ by

(3.12) $$ \begin{align} \mathfrak{s}_J:=\prod_{\alpha\in\Phi_0^+}k_\alpha^{\eta_J(\alpha)\alpha} \end{align} $$

(i.e., $\mathfrak {s}_J$ is the multiplicative character mapping $\lambda \in \Lambda $ to $\prod _{\alpha \in \Phi _0^+}k_\alpha ^{\eta _J(\alpha )\alpha (\lambda )}$ ).

Special cases are

$$\begin{align*}\mathfrak{s}_{[1,r]}=\prod_{\alpha\in\Phi_0^+}k_\alpha^{-\alpha},\qquad\quad \mathfrak{s}_{\{0\}}=k_\varphi^{\varphi},\qquad\quad \mathfrak{s}_\emptyset=1_{T_{\Lambda}}, \end{align*}$$

with $1_{T_{\Lambda }}$ the neutral element of $T_{\Lambda }$ . Note that $\mathfrak {s}_J\vert _\Lambda =j_\Lambda (\mathfrak {s}_J\vert _{P^{\vee }})\in T_\Lambda $ since $\mathfrak {s}_J$ maps $\text {pr}_{E_{\text {co}}}(\Lambda )$ to $1$ .

Proposition 3.7. For $J\subsetneq [0,r]$ , we have

$$\begin{align*}\mathfrak{s}_J^{D\alpha_j^{\vee}}=k_j^{-2}\qquad \forall\, j\in J. \end{align*}$$

In particular, $L_{\Lambda ,J}=\mathfrak {s}_{J}T_{\Lambda ,J}$ in $T_\Lambda $ for all $\Lambda \in \mathcal {L}$ .

Proof. Let $i\in J_0$ . Since $s_iC^J=C^J$ , we have $\eta _J(s_i\alpha )=\eta _J(\alpha )$ for all $\alpha \in \Phi _0^+\setminus \{\alpha _i\}$ . Furthermore, $\eta _J(\alpha _i)=-1$ , and hence,

$$\begin{align*}s_i\mathfrak{s}_J=k_i^{-\eta_J(\alpha_i)\alpha_i}\prod_{\alpha\in\Phi_0^+\setminus\{\alpha_i\}}k_\alpha^{\eta_J(\alpha)s_i\alpha}= k_i^{2\alpha_i}\mathfrak{s}_J. \end{align*}$$

By (2.37), we conclude that $\mathfrak {s}_J^{\alpha _i^{\vee }}=k_i^{-2}$ .

Assume now that $0\in J$ . It then remains to show that $\mathfrak {s}_J^{\varphi ^{\vee }}=k_\varphi ^2$ . By (2.18), we have

$$\begin{align*}s_\varphi\mathfrak{s}_J=k_\varphi^{-2\eta_J(\varphi)\varphi}\left(\prod_{\alpha\in\Phi_0^+:\, \alpha(\varphi^{\vee})=1}k_\alpha^{-\eta_J(\alpha)}\right)^\varphi\mathfrak{s}_J. \end{align*}$$

Clearly, $\eta _J(\varphi )=1$ , and the $2$ -group $\langle -s_\varphi \rangle $ acts freely on $\{\alpha \in \Phi _0^+\,\, | \,\, \alpha (\varphi ^{\vee })=1\}$ since $\Phi _0$ is reduced. Let Z be a complete set of representatives of the $\langle -s_\varphi \rangle $ -orbits in $\{\alpha \in \Phi _0^+\,\, | \,\, \alpha (\varphi ^{\vee })=1\}$ . Then we conclude that

$$\begin{align*}s_\varphi\mathfrak{s}_J=k_\varphi^{-2\varphi}\left(\prod_{\alpha\in Z}k_\alpha^{-(\eta_J(\alpha)+\eta_J(-\alpha+\varphi))}\right)^\varphi\mathfrak{s}_J. \end{align*}$$

But for $\alpha \in Z$ , we have $\eta _J(-\alpha +\varphi )=-\eta _J(\alpha )$ (cf. Lemma 3.5); hence, $s_\varphi \mathfrak {s}_J=k_\varphi ^{-2\varphi }\mathfrak {s}_J$ . By (2.37), we then have $\mathfrak {s}_J^{\varphi ^{\vee }}=k_\varphi ^2$ , which completes the proof of the proposition.

3.2 The $\mathbb {H}$ -module $\mathbb {M}^J_t$

We introduce a family of $\mathbb {H}$ -modules, obtained by induction from one-dimensional representations of Y-parabolic subalgebras.

Definition 3.8. Let $J\subsetneq [0,r]$ .

  1. 1. The X-parabolic subalgebra $\mathbb {H}_J^X\subseteq \mathbb {H}$ is the unital subalgebra generated by $T_j$ ( $j\in J$ ) and $\mathcal {P}$ .

  2. 2. The Y-parabolic subalgebra $\mathbb {H}_J^Y\subseteq \mathbb {H}$ is the image of $\mathbb {H}_J^X$ under the duality anti-involution $\delta $ .

Concretely, $\mathbb {H}_{J}^Y\subseteq \mathbb {H}$ is the unital subalgebra generated by $\delta (T_j)$ ( $j\in J$ ) and $\mathcal {P}_Y$ . For $i\in J_0$ , the generator $\delta (T_i)$ is simply $T_i$ . If $0\in J$ , then the algebraic generator $\delta (T_0)=Y^{-\varphi ^{\vee }}T_0x^{-\varphi ^{\vee }}$ of $\mathbb {H}_J^Y$ may be replaced by $U_0$ , since $Y^{\varphi ^{\vee }}\in \mathbb {H}_J^Y$ .

Lemma 3.9. For $t\in L_{J}$ , there exists a unique algebra homomorphism $\chi _{J,t}^X: \mathbb {H}_{J}^X\rightarrow \mathbf {F}$ satisfying $\chi _{J,t}^X(p)=p(t)$ ( $p\in \mathcal {P}$ ) and $\chi _{J,t}^X(T_j)=k_j$ ( $j\in J$ ).

Proof. Let $H_J\subseteq H$ be the parabolic sub-algebra generated by $T_j$ ( $j\in J$ ). A presentation of $\mathbb {H}^X_J$ in terms of $\mathcal {P}$ and $H_J$ is given by Definition 2.12, with the indices j and $j^\prime $ now taken from the subset J.

The assignment $T_j\mapsto k_j$ ( $j\in J$ ) defines the trivial one-dimensional representation of $H_J$ . To show that it extends to a well-defined algebra map $\chi _{J,t}^X: \mathbb {H}_J^X\rightarrow \mathbf {F}$ satisfying $\chi _{J,t}^X(p):=p(t)$ ( $p\in \mathcal {P}$ ), it suffices to show that the cross relations (2.40) for $p\in \mathcal {P}$ and $j\in J$ are respected by $\chi _{J,t}^X$ .

If $j\in J$ and $k_j^2=1$ , then $t\in L_{J}$ implies $t^{\alpha _j^{\vee }}=1$ ; hence, $s_jt=t$ . In this case, the cross relation (2.40) reduces to $T_jp=s_j(p)T_j$ , which is clearly respected by $\chi _{J,t}^X$ .

If $j\in J$ and $k_j^2\not =1$ , then $t^{\alpha _j^{\vee }}=k_j^{-2}\not =1$ , and a direct computation shows that $\chi _{J,t}^X$ respects the cross relation (2.40).

Corollary 3.10. For $t\in L_{J}$ , there exists a unique algebra map $\chi _{J,t}: \mathbb {H}_J^Y\rightarrow \mathbf {F}$ satisfying

(3.13) $$ \begin{align} \begin{aligned} \chi_{J,t}(p(Y))&=p(t^{-1})\qquad\,\,\,\, (p\in\mathcal{P}),\\ \chi_{J,t}(T_i)&=k_i\qquad\qquad\,\,\, (i\in J_0),\\ \chi_{J,t}(T_0x^{-\varphi^{\vee}})&=k_\varphi t^{-\varphi^{\vee}}\qquad \mbox{ if }\,\,\, 0\in J. \end{aligned} \end{align} $$

Proof. Composing the anti-algebra isomorphism $\delta \vert _{\mathbb {H}_J^Y}: \mathbb {H}_J^Y\overset {\sim }{\longrightarrow }\mathbb {H}_J^X$ with $\chi _{J,t}^X: \mathbb {H}_J^X\rightarrow \mathbf {F}$ gives a well-defined algebra map

(3.14) $$ \begin{align} \chi_{J,t}:=\chi_{J,t}^X\circ\delta\vert_{\mathbb{H}_J^Y}: \mathbb{H}_J^Y\rightarrow\mathbf{F}. \end{align} $$

A direct check shows that $\chi _{J,t}$ satisfies (3.13).

Write

$$\begin{align*}\mathbf{F}_{J,t}=\mathbf{F}1_{J,t} \end{align*}$$

for the one-dimensional $\mathbb {H}_{J}^Y$ -module with representation map $\chi _{J,t}$ .

Definition 3.11. We call

$$\begin{align*}\mathbb{M}^J_t:=\mathbb{H}\otimes_{\mathbb{H}_{J}^Y}\mathbf{F}_{J,t} \end{align*}$$

the Y-parabolically induced cyclic $\mathbb {H}$ -module relative to $J\subsetneq [0,r]$ and $t\in L_J$ .

Write

$$\begin{align*}m^J_t:=1\otimes_{\mathbb{H}_{J}^Y}1_{J,t}\in\mathbb{M}^J_t. \end{align*}$$

It is a cyclic vector of $\mathbb {M}^J_t$ , and

$$\begin{align*}hm^J_t=\chi_{J,t}(h)m^J_t\qquad\forall\, h\in\mathbb{H}_J^Y. \end{align*}$$

In particular, $p(Y)m^J_t=p(t^{-1})m^J_t$ for all $p\in \mathcal {P}$ . In Cherednik’s terminology [Reference Cherednik8, §3.6], $\mathbb {M}^J_t$ is a Y-cyclic $\mathbb {H}$ -module, and the $\mathbb {H}$ -module $\mathbb {M}^{\emptyset }_t$ ( $t\in T$ ) is the universal Y-cyclic $\mathbb {H}$ -module $\mathcal {I}_Y[t^{-1}]$ from [Reference Cherednik8, §3.6.1]. See [Reference Vasserot45] for a geometric approach to the representation theory of $\mathbb {H}$ .

Remark 3.12.

  1. 1. For $t\in T$ , write

    (3.15) $$ \begin{align} J(t):=\{j\in [0,r] \,\, | \,\, t^{\alpha_j^{\vee}}=k_j^{-2}\}. \end{align} $$

    Then $L_J=\bigsqcup _{J^\prime \supseteq J}L^{J^\prime }$ (disjoint union) with

    $$\begin{align*}L^{J^\prime}:=\{ t\in T \,\ | \,\, J(t)=J^\prime\}. \end{align*}$$

    If $t^\prime \in L^{J^\prime }$ and $J^\prime \supseteq J$ , then $M^{J^\prime }_{t^\prime }$ is a quotient of $M^J_{t^\prime }$ . In Theorem 6.15, we will show that $M^J_t$ is a simple $\mathbb {H}$ -module for generic $t\in L^J$ .

  2. 2. The $\mathbb {H}$ -representations $\mathbb {M}^I_t$ ( $I\subseteq [1,r]$ & $t\in L^I$ ) are analogues of the so-called standard modules $M(\lambda ,\text {Triv})$ of the rational Cherednik algebra, as defined in [Reference Berest, Etingof and Ginzburg1]. Here, $\lambda \in \mathfrak {h}^*$ with $\mathfrak {h}:=E\otimes _{\mathbb {R}}\mathbb {C}$ , and $\text {Triv}$ is the trivial representation of $W_{0,\lambda }$ , with $W_{0,\lambda }\subseteq W_0$ the stabiliser subgroup of $\lambda $ .

The‘Fourier dual’ of $\mathbb {M}^J_t$ ( $t\in L_J$ ) is the X-parabolically induced cyclic $\mathbb {H}$ -module

$$\begin{align*}{}^X\mathbb{M}^J_t:=\mathbb{H}\otimes_{\mathbb{H}_J^X}\mathbf{F}^X_{J,t}, \end{align*}$$

with $\mathbf {F}^X_{J,t}$ the one-dimensional $\mathbb {H}_J^X$ -module with representation map $\chi _{J,t}^X$ .

Remark 3.13. The following explicit realisations of the Y-parabolically and X-parabolically induced $\mathbb {H}$ -modules are known:

  1. 1. Cherednik’s [Reference Cherednik8] polynomial representation $\pi : \mathbb {H}\rightarrow \text {End}(\mathcal {P})$ , defined in terms of Demazure-Lusztig operators (see (4.8)). It is isomorphic to $\mathbb {M}^{[1,r]}_{t_{\text {sph}}}$ with

    (3.16) $$ \begin{align} t_{\text{sph}}:=\mathfrak{s}_{[1,r]}=\prod_{\alpha\in\Phi_0^+}k_\alpha^{-\alpha}\in L_{[1,r]}. \end{align} $$

    Under suitable generic conditions on q and $\mathbf {k}$ , nonsymmetric Macdonald polynomials provide the simultaneous eigenfunctions of the commuting Cherednik operators $\pi (Y^\mu )$ ( $\mu \in Q^{\vee }$ ).

  2. 2. Analytic realisation of the universal Y-cyclic $\mathbb {H}$ -module $\mathbb {M}_t^\emptyset $ for generic $t\in T$ . In this case, $\mathbf {F}=\mathbb {C}$ and $0<k_a,q<1$ , and $\mathbb {M}_t^\emptyset $ is realised as subrepresentation of the space $\mathcal {M}(T)$ of meromorphic functions on the complex torus T, with the action again defined by Demazure-Lusztig operators. The cyclic vectors $m_t^\emptyset $ are $\mathcal {E}(\cdot ,w_0t)$ with $\mathcal {E}$ Cherednik’s [Reference Cherednik10] global spherical function $\mathcal {E}\in \mathcal {M}(T\times T)$ , also known as the basic hypergeometric function [Reference Stokman43, Thm. 2.13]. The simultaneous eigenfunctions for the action of the commuting Cherednik operators on $\mathcal {M}(T)$ are of the form $\mathcal {E}(\cdot ,t^\prime )$ for appropriate $t^\prime \in T$ .

  3. 3. Cherednik’s realisation of ${}^X\mathbb {M}^J_t$ for generic $t\in L_J$ on the space of finitely supported $\mathbf {F}$ -valued functions on $W/W_J$ , with the action defined in terms of discrete Demazure-Lusztig type operators [Reference Cherednik8, §3.4.2]. The delta-functions are the simultaneous eigenfunctions for the action of the multiplication operators $x^\mu $ ( $\mu \in Q^{\vee }$ ).

In the following section, we realise the Y-parabolically induced $\mathbb {H}$ -modules $\mathbb {M}^J_t$ ( $t\in L_J$ ) on spaces of quasi-polynomials. The resulting simultaneous eigenfunctions for the action of $Y^\mu $ ( $\mu \in Q^{\vee }$ ) become quasi-polynomial generalisations of the nonsymmetric Macdonald polynomials.

Lemma 3.14. For $t\in L_J$ , set

(3.17) $$ \begin{align} m^J_{w;t}:=\delta(T_{w^{-1}})\,m^J_t\in\mathbb{M}^J_t\qquad (w\in W^J). \end{align} $$

Then $\{m^J_{w;t}\,\, | \,\, w\in W^J\}$ is a basis of $\mathbb {M}^J_t$ .

Proof. The Poincaré-Birkhoff-Witt Theorem for $\mathbb {H}$ implies that $\mathbb {H}$ is a free left $\mathbb {H}_{J}^{X}$ -module with basis $\{T_{w^{-1}}\,\, | \,\, w\in W^J\}$ . Now apply the anti-involution $\delta $ and recall that $\mathbb {H}_{J}^Y=\delta (\mathbb {H}_{J}^X)$ .

Lemma 3.15. For $\mu \in Q^{\vee }\cap \overline {E}_-$ and $v\in W_0$ such that $\tau (\mu )v\in W^J$ , we have

$$\begin{align*}m_{\tau(\mu)v;t}^J=x^\mu T_vm_t^J. \end{align*}$$

Proof. By (2.16) and (2.17), we have

$$\begin{align*}m_{\tau(\mu)v;t}^J=\delta(T_{v^{-1}}T_{\tau(-\mu)})m_t^J. \end{align*}$$

But $-\mu \in Q^{\vee }\cap \overline {E}_+$ ; hence, $T_{\tau (-\mu )}=Y^{-\mu }$ by (2.43). Consequently,

$$\begin{align*}m_{\tau(\mu)v;t}^J=\delta(Y^{-\mu})\delta(T_{v^{-1}})m_t^J=x^\mu T_vm_t^J.\\[-24pt] \end{align*}$$

Remark 3.16. For $w\in W$ , write $\mu _w\in Q^{\vee }$ and $v_w\in W_0$ such that $w=\tau (\mu _w)v_w$ . We will prove in Theorem 4.5(3) that $\{x^{\mu _w}T_{v_w}m^J_t\,\,\, | \,\,\, w\in W^J\}$ is also a basis of $\mathbb {M}^J_t$ . The quasi-polynomial realisation of $\mathbb {M}^J_t$ (Theorem 4.5) provides an explicit description of the $\mathbb {H}$ -action on the basis $\{x^{\mu _w}T_{v_w}m^J_t\,\,\, | \,\,\, w\in W^J\}$ of $\mathbb {M}^J_t$ , with the action described in terms of truncated Demazure-Lusztig operators.

4 Quasi-polynomial and quasi-rational representations

Let $c\in E$ be an element lying in the face $C^J$ of $\overline {C}_+$ . In this section, we will give explicit realisations of the $\mathbb {H}$ -modules $\mathbb {M}^{J}_t$ ( $t\in L_{J}$ ) on the space of quasi-polynomials with exponents in the affine Weyl group orbit $\mathcal {O}_c$ . By localisation, we also obtain nontrivial families of W-actions on spaces of quasi-rational functions.

4.1 Spaces of quasi-polynomials

The group algebra $\mathbf {F}[E]$ contains $\mathcal {P}_\Lambda $ as sub-algebra for all $\Lambda \in \mathcal {L}$ . The formula $v(x^y):=x^{vy}$ ( $v\in W_0$ , $y\in E$ ) defines a $W_0$ -action on $\mathbf {F}[E]$ by algebra automorphisms, which is compatible with the $W_0$ -action on $\mathcal {P}_\Lambda $ . We write

$$\begin{align*}x^{(y,\ell)}:=q^\ell x^y\in\mathbf{F}[E]\qquad (y,\ell)\in E\times\mathbb{Z}. \end{align*}$$

For $c\in \overline {C}_+$ , write

$$\begin{align*}\mathcal{P}^{(c)}:=\bigoplus_{y\in\mathcal{O}_c}\mathbf{F}x^y. \end{align*}$$

We call $\mathcal {P}^{(c)}$ the space of quasi-polynomials with exponents in $\mathcal {O}_c$ .

Definition 4.1. Let $\Lambda \in \mathcal {L}$ , $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ,J}$ . For $y\in \mathcal {O}_c$ , define

(4.1) $$ \begin{align} \mathfrak{t}_y:=w_y\mathfrak{t}\in T_\Lambda. \end{align} $$

Note that

$$\begin{align*}\mathfrak{t}_{wy}=w\mathfrak{t}_y \qquad\quad (w\in W,\, y\in\mathcal{O}_{c}) \end{align*}$$

since $W^J=\{w_y\}_{y\in \mathcal {O}_c}$ and $\mathfrak {t}\in T_J\subseteq T^{W_J}$ . In particular,

(4.2) $$ \begin{align} \mathfrak{t}_{y+\mu}=q^\mu\mathfrak{t}_y\qquad (y\in\mathcal{O}_c,\, \mu\in Q^{\vee}). \end{align} $$

In the following lemma, we identify $W\ltimes \mathcal {P}$ with $\mathbb {H}(\mathbf {1},q)$ by the isomorphism (2.41).

Lemma 4.2. For $c\in C^J$ and $\mathfrak {t}\in T_J$ , the formulas $K_{\mathfrak {t}}x^y:=qx^y$ and

(4.3) $$ \begin{align} v_{\mathfrak{t}}x^y:=x^{vy},\quad \tau(\mu)_{\mathfrak{t}}x^y:=\mathfrak{t}_y^{-\mu}x^y,\quad \nu_{\mathfrak{t}}(x^y):=x^{y+\nu} \qquad (v\in W_0,\,\mu,\nu\in Q^{\vee}) \end{align} $$

for $y\in \mathcal {O}_c$ define a $\mathbf {F}$ -linear $\mathbb {W}$ -action on $\mathcal {P}^{(c)}$ . The resulting $\mathbf {F}[\mathbb {W}]$ -action on $\mathcal {P}^{(c)}$ descends to an action of $W\ltimes \mathcal {P}$ via the isomorphism (2.33). Furthermore,

$$\begin{align*}\mathbb{M}_{\mathfrak{t}}^J\overset{\sim}{\longrightarrow}(\mathcal{P}^{(c)},\cdot_{\mathfrak{t}})\quad \mbox{ as modules over }\, \mathbb{H}(\mathbf{1},q)\simeq W\ltimes\mathcal{P} \end{align*}$$

with the isomorphism defined by $(x^\mu v)m_{\mathfrak {t}}^J\mapsto x^{\mu +vc}$ ( $\mu \in Q^{\vee },\, v\in W_0$ ).

Proof. It is a direct check that (4.3) defines a $\mathbb {W}$ -action on $\mathcal {P}^{(c)}$ . For the last statement, note that $\tau (\mu )_{\mathfrak {t}}x^c=\mathfrak {t}^{-\mu }x^c$ for all $\mu \in Q^{\vee }$ , and for $i\in J_0$ (see (3.10)),

$$\begin{align*}\delta(s_i)_{\mathfrak{t}}x^c=s_{i,\mathfrak{t}}x^c=x^{s_ic}=x^c. \end{align*}$$

If $0\in J$ , then $c=s_0c=s_\varphi c+\varphi ^{\vee }$ and $\delta (s_0)=\varphi ^{\vee } s_\varphi \in \mathbb {W}$ ; hence,

$$\begin{align*}\delta(s_0)_{\mathfrak{t}}x^c=(\varphi^{\vee})_{\mathfrak{t}}x^{s_\varphi c}=x^{s_\varphi c+\varphi^{\vee}}=x^c. \end{align*}$$

Hence, there exists a unique morphism $\mathbb {M}_{\mathfrak {t}}^J\rightarrow (\mathcal {P}^{(c)},\cdot _{\mathfrak {t}})$ of $W\ltimes \mathcal {P}$ -modules satisfying $m_{\mathfrak {t}}^J\mapsto x^c$ . Then $(x^\mu v)m_{\mathfrak {t}}^J$ maps to $(\mu v)_{\mathfrak {t}}x^c=x^{\mu +vc}$ for $\mu \in Q^{\vee }$ and $v\in W_0$ .

In the special case $c=0$ , we necessarily have $\mathfrak {t}=1_T$ . In this case, the $W\ltimes \mathcal {P}$ -action on $\mathcal {P}^{(0)}=\mathcal {P}$ reduces to the action introduced in Subsection 2.6,

$$\begin{align*}w_{1_T}x^\mu=w(x^\mu)\qquad\quad (w\in W,\, \mu\in Q^{\vee}). \end{align*}$$

Note that formula (2.36) generalises to

(4.4) $$ \begin{align} s_{a,\mathfrak{t}}x^y=\mathfrak{t}_y^{-\ell\alpha^{\vee}}x^{s_\alpha y}\qquad (a=(\alpha,\ell)\in\Phi,\,\, y\in\mathcal{O}_c). \end{align} $$

In particular, $s_{0,\mathfrak {t}}x^y=\mathfrak {t}_y^{\varphi ^{\vee }}x^{s_\varphi y}$ for $y\in \mathcal {O}_c$ .

The natural extension of formula (2.31) to the present context is as follows.

Lemma 4.3. Let $c\in C^J$ , $\mathfrak {t}\in T_J$ , $y\in \mathcal {O}_c$ and $a\in \Phi $ such that $a(y)\in \mathbb {Z}$ . Then

  1. 1. $\mathfrak {t}_y^{Da^{\vee }}=q_{Da}^{Da(y)}$ ,

  2. 2. $ s_{a,\mathfrak {t}}x^y=x^{y-Da(y)a^{\vee }}$ .

Proof. Write $\alpha :=Da\in \Phi _0$ and $n:=\alpha (y)\in \mathbb {Z}$ . The affine root $b:=w_y^{-1}(\alpha ,-n)\in \Phi $ satisfies $b(c_y)=0$ . Since $c_y\in C^J$ , we conclude that $b\in \Phi _J$ . Note that $\{\alpha _0^{\vee },\ldots ,\alpha _r^{\vee }\}$ is a basis for the affine coroot system $\Phi ^{\vee }$ ; hence, the coroot $b^{\vee }$ lies in $\bigoplus _{j\in J}\mathbb {Z}\alpha _j^{\vee }$ , and hence, $\mathfrak {t}^{b^{\vee }}=1$ . Then

$$\begin{align*}1=\mathfrak{t}^{b^{\vee}}=(w_y\mathfrak{t})^{(\alpha,-n)^{\vee}}=q_\alpha^{-n}\mathfrak{t}_y^{\alpha^{\vee}}, \end{align*}$$

where the second equality follows from (2.31). Hence, $\mathfrak {t}_y^{\alpha ^{\vee }}=q_\alpha ^{\alpha (y)}$ , which proves (1).

Set $\ell :=a(0)\in \mathbb {Z}$ . Then $a=(\alpha ,\ell )$ and

$$\begin{align*}s_{a,\mathfrak{t}}x^y=\mathfrak{t}_y^{-\ell\alpha^{\vee}}x^{s_\alpha y}=q_\alpha^{-\ell\alpha(y)}x^{s_\alpha y}=x^{y-Da(y)a^{\vee}}, \end{align*}$$

where we have used (4.4) for the first equality, (1) for the second equality, and (2.36) and (2.31) for the last equality. This proves (2).

Let $T_{\mathbb {R}}^{\vee }$ be the space of $Q^{\vee }$ -orbits in E, with $Q^{\vee }$ acting on E by translations. It is a real torus admitting a natural $W_0$ -action. For $\gamma \in T^{\vee }_{\mathbb {R}}$ , set

$$\begin{align*}\mathcal{P}_\gamma:=\bigoplus_{y\in\gamma}\mathbf{F}x^y, \end{align*}$$

which is a cyclic $\mathcal {P}$ -submodule of $\mathbf {F}[E]$ . It equals $\mathcal {P}$ when $\gamma $ is the neutral element of $T_{\mathbb {R}}^{\vee }$ . Writing $[c]:=c+Q^{\vee }\in T_{\mathbb {R}}^{\vee }$ , we have

$$\begin{align*}\mathcal{P}^{(c)}=\bigoplus_{\gamma\in W_0[c]}\mathcal{P}_\gamma \end{align*}$$

since $W_0[c]\simeq Wc/Q^{\vee }$ , which allows one to think of $\mathcal {P}^{(c)}$ as vector-valued polynomial functions on T.

The spaces $\mathcal {P}_\gamma $ ( $\gamma \in T_{\mathbb {R}}^{\vee }$ ) naturally arise in the Bethe ansatz for the trigonometric Gaudin model [Reference Mukhin and Varchenko31, Reference Mukhin, Tarasov and Varchenko32]. They also serve as the ambient spaces of the multivariable Baker-Akhiezer functions; see, for example, [Reference Chalykh4, Reference Chalykh and Etingof5, Reference Stokman42]. Multivariable Baker-Akhiezer functions are terminating power series solutions of the spectral problem for the Macdonald operators, which exist for special values of $k_j$ (for instance, when $k_j$ a nonpositive integral power of $q_{\alpha _j}$ ). In [Reference Chalykh4, Reference Mukhin and Varchenko31, Reference Mukhin, Tarasov and Varchenko32], elements in $\mathcal {P}_\gamma $ are called quasi-polynomials or quasi-exponentials.

For rational weights $c\in \mathbb {Q}\otimes _{\mathbb {Z}}Q^{\vee }$ , we may view $\mathcal {P}^{(c)}$ as subspace of the space of polynomials on the torus $\text {Hom}(\Lambda ,\mathbf {F}^\times )$ for an appropriate lattice $Q^{\vee }\subset \Lambda \subset E^\prime $ . This is a natural viewpoint in the context of metaplectic representation theory; see Subsection 9.2.

In Subsection 4.3, we deform the $\mathfrak {t}$ -dependent $W\ltimes \mathcal {P}$ -action on $\mathcal {P}^{(c)}$ to an action of the double affine Hecke algebra $\mathbb {H}$ , with the $T_j$ ’s acting by truncated Demazure-Lusztig operators.

4.2 Truncated divided difference operators

For $a\in \Phi $ , let $\nabla _a$ be the linear operator on $\mathbf {F}[E]$ defined by

(4.5) $$ \begin{align} \nabla_a(x^y):=\left(\frac{1-x^{-\lfloor Da(y)\rfloor a^{\vee}}}{1-x^{a^{\vee}}}\right)x^y\qquad (y\in E), \end{align} $$

where $\lfloor z\rfloor \in \mathbb {Z}$ is the floor of $z\in \mathbb {R}$ . Note that $\nabla _a$ is well defined, since the numerator is divisible by $1-x^{a^{\vee }}$ in $\mathbf {F}[E]$ . The subspaces $\mathcal {P}^{(c)}$ ( $c\in \overline {C}_+$ ) are $\nabla _a$ -stable.

We will write $\nabla _j:=\nabla _{\alpha _j}$ , so that

$$\begin{align*}\nabla_0(x^y)=\left(\frac{1-(q_\varphi^{-1}x^{\varphi^{\vee}})^{\lfloor -\varphi(y)\rfloor}} {1-q_\varphi x^{-\varphi^{\vee}}}\right)x^y,\qquad \nabla_i(x^y)= \left(\frac{1-x^{-\lfloor \alpha_i(y)\rfloor\alpha_i^{\vee}}}{1-x^{\alpha_i^{\vee}}}\right)x^y \end{align*}$$

for $1\leq i\leq r$ and $y\in E$ .

Lemma 4.4. For $a\in \Phi $ and $y,y^\prime \in E$ such that $a(y^\prime )\in \mathbb {Z}$ , we have

(4.6) $$ \begin{align} \begin{aligned} \nabla_a(x^{y^\prime})&=\left(\frac{1-x^{-Da(y^\prime)a^{\vee}}}{1-x^{a^{\vee}}}\right)x^{y^\prime},\\ \nabla_a(x^{y+y^\prime})&-x^{y^\prime-Da(y^\prime)a^{\vee}}\nabla_a(x^{y})=\nabla_a(x^{y^\prime})x^{y}. \end{aligned} \end{align} $$

Proof. This is a straightforward verification which we leave to the reader.

It follows from (2.31) and the first formula in (4.6) that $\nabla _a$ restricts to the usual divided difference operator on $\mathcal {P}_\Lambda $ ( $\Lambda \in \mathcal {L}$ ),

(4.7) $$ \begin{align} \nabla_a(x^\mu)=\frac{x^\mu-s_a(x^\mu)}{1-x^{a^{\vee}}}=\frac{x^\mu-x^{s_a\cdot\mu}}{1-x^{a^{\vee}}} \qquad (\mu\in\Lambda). \end{align} $$

Cherednik’s [Reference Cherednik8, Thm. 3.2.1] polynomial representation $\pi : \mathbb {H}\rightarrow \text {End}(\mathcal {P})$ is then explicitly defined by

(4.8) $$ \begin{align} \begin{aligned} \pi(T_j)x^\mu&:=k_js_j(x^\mu)+(k_j-k_j^{-1})\nabla_j(x^\mu),\\ \pi(x^\lambda)x^\mu&:=x^{\lambda+\mu} \end{aligned} \end{align} $$

for $0\leq j\leq r$ and $\lambda ,\mu \in Q^{\vee }$ .

In the next subsection, we introduce families of quasi-polynomial representations $\pi _{c,\mathfrak {t}}: \mathbb {H}\rightarrow \text {End}(\mathcal {P}^{(c)})$ ( $c\in C^J,\,\mathfrak {t}\in T_J$ ). The action of $\mathcal {P}$ will still be by multiplication operators. The action of $T_j$ will be reminiscent to formula (4.8) for $\pi (T_j)$ ; its first term will involve the $\mathfrak {t}$ -twisted W-action on $\mathcal {P}^{(c)}$ , and its second term the truncated divided difference operator $\nabla _j\vert _{\mathcal {P}^{(c)}}$ .

4.3 The quasi-polynomial realisation of $\mathbb {M}_t^J$

Denote by $\chi _B: \mathbb {R}\rightarrow \{0,1\}$ the characteristic function of the subset $B\subseteq \mathbb {R}$ . Define $\eta : \mathbb {R}\rightarrow \{-1,0,1\}$ by

(4.9) $$ \begin{align} \eta=\chi_{\mathbb{Z}_{>0}}-\chi_{\mathbb{Z}_{\leq 0}}, \end{align} $$

and set

(4.10) $$ \begin{align} \kappa_v(y):=\prod_{\alpha\in\Pi(v)}k_\alpha^{-\eta(\alpha(y))}\qquad (v\in W_0,\, y\in E). \end{align} $$

Note that $\kappa _v\in \mathcal {F}_\Sigma (E,\mathbf {F}^\times )$ by Lemma 2.5.

Theorem 4.5. Let $J\subsetneq [0,r]$ , $c\in C^J$ and $\mathfrak {t}\in T_J$ .

  1. (1) The formulas

    (4.11) $$ \begin{align} \begin{aligned} \pi_{c,\mathfrak{t}}(T_j)x^y&:=k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}s_{j,\mathfrak{t}}x^y +(k_j-k_j^{-1})\nabla_j(x^y),\\ \pi_{c,\mathfrak{t}}(x^\mu)x^y&:=x^{y+\mu} \end{aligned} \end{align} $$
    for $0\leq j\leq r$ , $\mu \in Q^{\vee }$ and $y\in \mathcal {O}_c$ turn $\mathcal {P}^{(c)}$ in a $\mathbb {H}$ -module, which we denote by $\mathcal {P}_{\mathfrak {t}}^{(c)}$ .
  2. (2) There exists a unique isomorphism

    $$\begin{align*}\phi_{c,\mathfrak{t}}: \mathbb{M}^{J}_{\mathfrak{s}_{J}\mathfrak{t}}\overset{\sim}{\longrightarrow} \mathcal{P}^{(c)}_{\mathfrak{t}} \end{align*}$$
    of $\mathbb {H}$ -modules satisfying $\phi _{c,\mathfrak {t}}\big (m^{J}_{\mathfrak {s}_{J}\mathfrak {t}}\big )=x^c$ .
  3. (3) We have

    (4.12) $$ \begin{align} \phi_{c,\mathfrak{t}}\big(x^\mu T_vm^{J}_{\mathfrak{s}_{J}\mathfrak{t}}\big)= \kappa_v(c)x^{\mu+vc}\qquad (\mu\in Q^{\vee},\, v\in W_0). \end{align} $$

    In particular, $\{x^{\mu _w}T_{v_w}m^{J}_{\mathfrak {s}_{J}\mathfrak {t}}\,\, | \,\, w\in W^J\}$ is a basis of $\mathbb {M}^{J}_{\mathfrak {s}_{J}\mathfrak {t}}$ .

Note that the $\mathbb {H}$ -action on $\mathcal {P}^{(c)}_{\mathfrak {t}}$ reduces to the $\mathfrak {t}$ -dependent $W\ltimes \mathcal {P}$ -action on $\mathcal {P}^{(c)}$ from Lemma 4.2 when $k_a=1$ for all $a\in \Phi $ .

Furthermore, for $c=0$ and $\mathfrak {t}=1_T$ , $\pi _{0,1_T}: \mathbb {H}\rightarrow \text {End}(\mathcal {P})$ is Cherednik’s polynomial representation $\pi $ (see (4.8)). In this case, the theorem is due to Cherednik; see [Reference Cherednik7, Thm. 2.3] and [Reference Cherednik9, Thm. 3.1]. We call $\pi _{c,\mathfrak {t}}$ the quasi-polynomial representation of $\mathbb {H}$ .

The proof of Theorem 4.5 is relegated to Section 5. The proof of part (1), which will be given in Subsection 5.2, will follow closely the proof of [Reference Sahi, Stokman and Venkateswaran38, Thm. 3.7] introducing the metaplectic affine Hecke algebra action on Laurent polynomials (in Section 9, we will show how [Reference Sahi, Stokman and Venkateswaran38, Thm. 3.7] follows from Theorem 4.5). Similarly as in the context of Cherednik’s basic representation and its associated Macdonald polynomials (the special case $c=0$ and $\mathfrak {t}=1_T$ ), the proof of parts (2) and (3) in Subsections 5.45.5, as well as the existence of the quasi-polynomial eigenfunctions of $\pi _{c,\mathfrak {t}}(Y^\lambda )$ ( $\lambda \in Q^{\vee }$ ) in Section 6, will require a detailed analysis of the triangularity properties of the truncated Demazure-Lusztig operators $\pi _{c,\mathfrak {t}}(T_j)$ ( $j\in [0,r]$ ). For a detailed discussion of the triangularity properties of Demazure-Lusztig operators in the context of Cherednik’s polynomial representation, see, for example, [Reference Macdonald29, §4.6].

In the following two subsections, we give some direct consequences of Theorem 4.5.

4.4 The face dependence of the quasi-polynomial representation

The quasi-polynomial representations $\pi _{c,\mathfrak {t}}$ are equivalent for different values of $c\in C^J$ by Theorem 4.5(2). For $\mathfrak {t}\in T_J$ and $c,c^\prime \in C^J$ , the isomorphism is explicitly given by

(4.13) $$ \begin{align} \mathcal{P}_{\mathfrak{t}}^{(c)}\overset{\sim}{\longrightarrow}\mathcal{P}_{\mathfrak{t}}^{(c^\prime)},\qquad x^{wc}\mapsto x^{wc^\prime}\quad (w\in W^J). \end{align} $$

As a direct consequence of (4.13), we obtain the following result.

Corollary 4.6. Let $J\subsetneq [0,r]$ , $\mathfrak {t}\in T_{J}$ , $w\in W^J$ and $h\in \mathbb {H}$ . Then

$$\begin{align*}\pi_{c,\mathfrak{t}}(h)x^{wc}= \sum_{w^\prime\in W^J}e_{w,w^\prime;\mathfrak{t}}^{J,h} x^{w^\prime c}\qquad \forall\,c\in C^J \end{align*}$$

with coefficients $e_{w,w^\prime ;\mathfrak {t}}^{J,h}\in \mathbf {F}$ ( $w^\prime \in W^J$ ) independent of the choice of $c\in C^J$ .

The closure relations between the faces $C^J$ of $\overline {C}_+$ are translated to quotient relations for the quasi-polynomial representations $\pi _{c,\mathfrak {t}}$ (see Remark 3.12(1)).

Corollary 4.7. Let $J\subseteq J^\prime \subsetneq [0,r]$ , $c\in C^J$ , $c^\prime \in C^{J^\prime }$ and $\mathfrak {t}^\prime \in T_{J^\prime }$ .

  1. 1. There exists a unique epimorphism

    $$\begin{align*}\text{pr}_{c,c^\prime}^{\mathfrak{t}^\prime}:\mathcal{P}^{(c)}_{\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime} \twoheadrightarrow\mathcal{P}^{(c^\prime)}_{\mathfrak{t}^\prime} \end{align*}$$
    of $\mathbb {H}$ -modules mapping $x^c$ to $x^{c^\prime }$ .
  2. 2. We have

    $$\begin{align*}\text{pr}_{c,c^\prime}^{\mathfrak{t}^\prime}(x^{\mu+vc})=\frac{\kappa_v(c^\prime)}{\kappa_v(c)}x^{\mu+vc^\prime}\qquad (\mu\in Q^{\vee},\, v\in W_0). \end{align*}$$

Proof. Write $t^\prime :=\mathfrak {s}_{J^\prime }\mathfrak {t}^\prime \in L_{J^\prime }\subseteq L_J=\mathfrak {s}_JT_J$ . Recall from Remark 3.12(1) that we have an epimorphism $\psi : \mathbb {M}^J_{t^\prime }\twoheadrightarrow \mathbb {M}^{J^\prime }_{t^\prime }$ of $\mathbb {H}$ -modules mapping $m^J_{t^\prime }$ to $m^{J^\prime }_{t^\prime }$ . Then

$$\begin{align*}\text{pr}^{\mathfrak{t}^\prime}_{c,c^\prime}:=\phi_{c^\prime,\mathfrak{t}^\prime}\circ\psi\circ\phi_{c,\mathfrak{s}_J^{-1}t^\prime}^{-1}: \mathcal{P}^{(c)}_{\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime} \twoheadrightarrow\mathcal{P}^{(c^\prime)}_{\mathfrak{t}^\prime} \end{align*}$$

has the desired properties (use Theorem 4.5 (2)&(3)).

Remark 4.8. Let $J\subseteq [1,r]$ and $c\in C^J$ . Then Corollary 4.7 gives an epimorphism $\text {pr}_{c,0}^{1_{T}}: \mathcal {P}^{(c)}_{\mathfrak {s}_J^{-1}\mathfrak {s}_0} \twoheadrightarrow \mathcal {P}^{(0)}_{1_T}$ of $\mathbb {H}$ -modules. The co-domain is Cherednik’s basic representation, and

$$\begin{align*}\text{pr}_{c,0}^{1_T}\big(x^{\mu+vc}\big)=\Big(\prod_{\alpha\in\Pi(v)}k_\alpha^{1+\eta(\alpha(c))}\Big)x^\mu\qquad (\mu\in Q^{\vee},\, v\in W_0). \end{align*}$$

4.5 Affine Weyl group actions on quasi-rational functions

We derive an affine Weyl group action on quasi-rational functions by localizing the quasi-polynomial representation $\pi _{c,\mathfrak {t}}$ (Theorem 4.5(1)). In Section 10, we will show that this action gives rise to an affine analog of Chinta and Gunnells’ $W_0$ -action on rational functions [Reference Chinta and Gunnells13, Reference Chinta and Gunnells14]. The approach in this subsection follows the paper [Reference Sahi, Stokman and Venkateswaran38], where the Chinta-Gunnells’ $W_0$ -action was obtained by localization of an appropriate metaplectic affine Hecke algebra representation.

Consider the $\mathcal {Q}$ -algebra

$$\begin{align*}\mathbf{F}_{\mathcal{Q}}[E]:=\mathcal{Q}\otimes_{\mathcal{P}}\mathbf{F}[E], \end{align*}$$

which is isomorphic to the localization of $\mathbf {F}[E]$ by $\mathcal {P}^\times $ . The algebra $W_0\ltimes \mathcal {Q}$ naturally acts on $\mathbf {F}_{\mathcal {Q}}[E]$ .

For $\gamma \in T_{\mathbb {R}}^{\vee }$ , denote by $\mathcal {Q}_\gamma $ the $\mathcal {Q}$ -submodule of $\mathbf {F}_{\mathcal {Q}}[E]$ generated by $\mathcal {P}_\gamma $ . Then

$$\begin{align*}\mathcal{Q}_\gamma=\mathcal{Q}x^y\qquad \forall\, y\in\gamma. \end{align*}$$

Write $\mathcal {Q}^{(c)}$ for the $\mathcal {Q}$ -submodule of $\mathbf {F}_{\mathcal {Q}}[E]$ generated by $\mathcal {P}^{(c)}$ ( $c\in \overline {C}_+$ ), so that

$$\begin{align*}\mathcal{Q}^{(c)}=\bigoplus_{\gamma\in W_0[c]}\mathcal{Q}_\gamma. \end{align*}$$

Note that $\mathcal {Q}^{(c)}$ is a cyclic $W_0\ltimes \mathcal {Q}$ -submodule of $\mathbf {F}_{\mathcal {Q}}[E]$ .

Fix $c\in C^J$ and $\mathfrak {t}\in T_J$ . The $W\ltimes \mathcal {P}$ -action on $\mathcal {P}^{(c)}$ from Lemma 4.2 extends to a $W\ltimes \mathcal {Q}$ -action on $\mathcal {Q}^{(c)}$ . The following theorem provides a $\mathbf {k}$ -deformation of this action.

Theorem 4.9. Let $J\subsetneq [0,r]$ , $c\in C^J$ and $\mathfrak {t}\in T_J$ . The formulas

(4.14) $$ \begin{align} \begin{aligned} \sigma_{c,\mathfrak{t}}(s_j)(x^y)&:= \frac{k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}(x^{\alpha_j^{\vee}}-1)}{(k_jx^{\alpha_j^{\vee}}-k_j^{-1})} s_{j,\mathfrak{t}}x^y+\frac{(k_j-k_j^{-1})}{(k_jx^{\alpha_j^{\vee}}-k_j^{-1})}x^{y-\lfloor D\alpha_j(y) \rfloor\alpha_j^{\vee}},\\ \sigma_{c,\mathfrak{t}}(f)(x^y)&:=fx^y \end{aligned} \end{align} $$

for $0\leq j\leq r$ , $y\in \mathcal {O}_c$ and $f\in \mathcal {Q}$ turn $\mathcal {Q}^{(c)}$ into a $W\ltimes \mathcal {Q}$ -module, which we denote by $\mathcal {Q}_{\mathfrak {t}}^{(c)}$ .

Proof. Let $\check {s}_j$ be the linear operator on $\mathcal {P}^{(c)}$ defined by

(4.15) $$ \begin{align} \check{s}_j(x^y):=x^{y-\lfloor D\alpha_j(y)\rfloor\alpha_j^{\vee}}\qquad (y\in\mathcal{O}_c). \end{align} $$

Note that

(4.16) $$ \begin{align} \check{s}_j(px^y)=s_j(p)\check{s}_j(x^y) \end{align} $$

for $0\leq j\leq r$ , $y\in \mathcal {O}_c$ and $p\in \mathcal {P}$ . We extend $\check {s}_j$ to a linear operator on $\mathcal {Q}^{(c)}$ by requiring (4.16) to hold true for all $p\in \mathcal {Q}$ .

Transporting the $\mathbb {H}^{X-\text {loc}}$ -action of the induced $\mathbb {H}^{X-\text {loc}}$ -module $\mathbb {H}^{X-\text {loc}}\otimes _{\mathbb {H}}\mathcal {P}^{(c)}_{\mathfrak {t}}$ to $\mathcal {Q}^{(c)}$ through the linear isomorphism

$$\begin{align*}\mathcal{Q}^{(c)}\overset{\sim}{\longrightarrow}\mathbb{H}^{X-\text{loc}}\otimes_{\mathbb{H}}\mathcal{P}^{(c)}_{\mathfrak{t}},\qquad fx^y\mapsto f\otimes_{\mathbb{H}}x^y\qquad (f\in\mathcal{Q},\, y\in\mathcal{O}_c) \end{align*}$$

(see (2.52)) gives a representation

$$\begin{align*}\pi_{c,\mathfrak{t}}^{X-\text{loc}}: \mathbb{H}^{X-\text{loc}}\rightarrow \text{End}(\mathcal{Q}^{(c)}). \end{align*}$$

By (2.53) and (4.16), the representation map $\pi _{c,\mathfrak {t}}^{X-\text {loc}}$ is explicitly given by

(4.17) $$ \begin{align} \begin{aligned} \pi_{c,\mathfrak{t}}^{X-\text{loc}}(T_j)(fx^y)&=k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}s_{j,\mathfrak{t}}(fx^y)+(k_j-k_j^{-1})\left(\frac{fx^y-\check{s}_j(fx^y)}{1-x^{\alpha_j^{\vee}}}\right),\\ \pi_{c,\mathfrak{t}}^{X-\text{loc}}(f^\prime)(fx^y)&=f^\prime fx^y \end{aligned} \end{align} $$

for $0\leq j\leq r$ , $y\in \mathcal {O}_c$ and $f,f^\prime \in \mathcal {Q}$ . By Theorem 2.16, we obtain a representation $\sigma _{c,\mathfrak {t}}: W\ltimes \mathcal {Q}\rightarrow \text {End}(\mathcal {Q}^{(c)})$ , defined by

(4.18) $$ \begin{align} \sigma_{c,\mathfrak{t}}(w\otimes f):=\pi_{c,\mathfrak{t}}^{X-\text{loc}}(\widetilde{S}_w^Xf) \qquad (w\in W, f\in\mathcal{Q}). \end{align} $$

Then $\sigma _{c,\mathfrak {t}}(f^\prime )(fx^y)=ff^\prime x^y$ for $f,f^\prime \in \mathcal {Q}$ and $y\in \mathcal {O}_c$ , and

$$ \begin{align*} \begin{aligned} \sigma_{c,\mathfrak{t}}(s_j)&=\pi_{c,\mathfrak{t}}^{X-\text{loc}}(\widetilde{S}_j^X)\\ &=\pi_{c,\mathfrak{t}}^{X-\text{loc}}\left( \left(\frac{x^{\alpha_j^{\vee}}-1}{k_jx^{\alpha_j^{\vee}}-k_j^{-1}}\right)T_j+ \left(\frac{k_j-k_j^{-1}}{k_jx^{\alpha_j^{\vee}}-k_j^{-1}}\right)\right)\\ &=\frac{k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}(x^{\alpha_j^{\vee}}-1)}{(k_jx^{\alpha_j^{\vee}}-k_j^{-1})} s_{j,\mathfrak{t}}+\frac{(k_j-k_j^{-1})}{(k_jx^{\alpha_j^{\vee}}-k_j^{-1})}\check{s}_j \end{aligned} \end{align*} $$

as linear operators on $\mathcal {Q}^{(c)}$ .

Remark 4.10.

  1. 1. Note that for $0\leq j\leq r$ , $f\in \mathcal {Q}$ and $y\in \mathcal {O}_c$ , we have

    $$\begin{align*}s_{j,\mathfrak{t}}(fx^y)= \check{s}_j(fx^y)\,\,\mbox{ if }\, \alpha_j(y)\in\mathbb{Z} \end{align*}$$
    by Lemma 4.3(2), and hence, $\sigma _{c,\mathfrak {t}}(s_j)(fx^y)=\check {s}_j(fx^y)$ if $\alpha _j(y)\in \mathbb {Z}$ . In particular, $\sigma _{0,1_T}$ reduces to the standard $W\ltimes \mathcal {Q}$ -action on $\mathcal {Q}^{(0)}=\mathcal {Q}$ as described in Subsection 2.6.
  2. 2. We have

    (4.19) $$ \begin{align} \sigma_{c,\mathfrak{t}}=\pi_{c,\mathfrak{t}}^{X-\text{loc}}\circ\beth \end{align} $$
    by Theorem 2.16 and (4.18).

Remark 4.11. Theorem 4.9 allows to rewrite the truncated Demazure-Lusztig type operators $\pi _{c,\mathfrak {t}}^{X-\text {loc}}(T_j)$ (see (4.11)) as standard Demazure-Lusztig operators involving the W-action $\sigma _{c,\mathfrak {t}}$ on $\mathcal {Q}^{(c)}$ ,

$$ \begin{align*} \pi_{c,\mathfrak{t}}^{X-\text{loc}}(T_j)f=k_jf+k_j^{-1}\left(\frac{1-k_j^2x^{\alpha_j^{\vee}}}{1-x^{\alpha_j^{\vee}}}\right)\big(\sigma_{c,\mathfrak{t}}(s_j)f-f\big) \qquad (f\in\mathcal{Q}^{(c)}). \end{align*} $$

This follows from (4.19) and the fact that

$$\begin{align*}\beth^{-1}(T_j)=k_j+k_j^{-1}\left(\frac{1-k_j^2x^{\alpha_j^{\vee}}}{1-x^{\alpha_j^{\vee}}}\right)(s_j-1) \end{align*}$$

in $W\ltimes \mathcal {Q}$ . This formula shows that $\pi _{c,\mathfrak {t}}^{X-\text {loc}}(T_j)$ is a so-called metaplectic Demazure-Lusztig operator (see [Reference Chinta, Gunnells and Puskas15, Reference Patnaik and Puskas34, Reference Patnaik and Puskas35]) in the context of metaplectic representation theory of reductive groups over non-archimedean local fields; see Section 10 for further details.

Remark 4.12. In the limit case where $\mathbf {k}=\mathbf {0}$ (meaning $k_{\alpha } = 0$ for all $\alpha \in \Phi _0$ ), the W-action from Theorem 4.9 does not depend on $\mathfrak {t}$ and satisfies

$$\begin{align*}\sigma_{c,\mathfrak{t}}^{\mathbf{k}=\mathbf{0}}(s_j)(x^y)=\check{s}_j(x^y)\qquad (0\leq j\leq r,\, y\in\mathcal{O}_c), \end{align*}$$

with $\check {s}_j$ defined by (4.15). For $y \in \mathcal {O}_c$ , let $\xi _y\in y+P^{\vee }$ be the unique element satisfying $0 \leq \alpha _i(\xi _y) < 1$ for all $1 \leq i \leq r$ . Then $\check {s}_i(x^y)=x^{\xi _y+s_i(y-\xi _y)}$ for $i\in [1,r]$ ; hence,

$$ \begin{align*} \sigma_{c, \mathfrak{t}}^{\mathbf{k}=\mathbf{0}}(v)(x^y) = x^{\xi_y + v (y-\xi_y)} \end{align*} $$

for all $v \in W_0$ . In particular, this is a shift of the standard $W_0$ -action.

5 The proof of the main theorem

This section is dedicated to the proof of Theorem 4.5. As a byproduct, we obtain triangularity properties of the commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) which will play an important role in the construction of the quasi-polynomial generalisations of the Macdonald polynomials in Section 6.

5.1 Properties of the base point $\mathfrak {s}_J$

Recall the step function $\eta $ defined by (4.9). It satisfies the identities

(5.1) $$ \begin{align} \eta(x)+\eta(-x)=-2\chi_{\{0\}}(x),\qquad \eta(x)+\eta(1-x)=0 \end{align} $$

for all $x\in \mathbb {R}$ .

Definition 5.1. Let $\Lambda \in \mathcal {L}$ . For $y\in E$ , set

(5.2) $$ \begin{align} \mathfrak{s}_y:=\prod_{\alpha\in\Phi^+_0}k_\alpha^{\eta(\alpha(y))\alpha}\in T_{\Lambda}. \end{align} $$

Note that $\mathfrak {s}_y\vert _\Lambda =j_\Lambda (\mathfrak {s}_y\vert _{P^{\vee }})$ since $\mathfrak {s}_y$ maps $\text {pr}_{E_{\text {co}}}(\Lambda )$ to $1$ .

For $J\subsetneq [0,r]$ , the element $\mathfrak {s}_J\in T_{P^{\vee }}$ (see (3.12)) can be recovered from $\mathfrak {s}_y$ as follows.

Lemma 5.2. Let $\Lambda \in \mathcal {L}$ . The function $y\mapsto \mathfrak {s}_y$ lies in $\mathcal {F}_\Sigma (E,T_{\Lambda })$ . Furthermore,

(5.3) $$ \begin{align} \mathfrak{s}_c=\mathfrak{s}_{J}\,\, \mbox{ if }\,\, c\in C^J. \end{align} $$

Proof. The first statement follows from Lemma 2.5 applied to $g=\eta $ . The second statement follows from (3.9) and (3.12).

We now analyse how the affine Weyl group W acts on $\mathfrak {s}_y\in T_{\Lambda }$ .

Lemma 5.3. Let $y\in E$ and $j\in [0,r]$ .

  1. 1. If $s_j\in W_y$ , then $\mathfrak {s}_y^{D\alpha _j^{\vee }}=k_j^{-2}$ and $s_{D\alpha _j}\mathfrak {s}_y=k_j^{2D\alpha _j}\mathfrak {s}_y$ ,

  2. 2. If $s_j\in W\setminus W_y$ , then $s_{D\alpha _j}\mathfrak {s}_y=\mathfrak {s}_{s_jy}$ .

Proof. Let $1\leq i\leq r$ . Then $\Pi (s_i)=\{\alpha _i\}$ ; hence,

$$\begin{align*}s_i\mathfrak{s}_y=\prod_{\alpha\in s_i\Phi_0^+}k_\alpha^{\eta(\alpha(s_iy))\alpha}= \mathfrak{s}_{s_iy}k_i^{-(\eta(\alpha_i(y))+\eta(-\alpha_i(y)))\alpha_i}. \end{align*}$$

Suppose that $s_iy=y$ . Then $\alpha _i(y)=0$ ; hence, $\eta (\alpha _i(y))+\eta (-\alpha _i(y))=-2$ , so $s_i\mathfrak {s}_y=\mathfrak {s}_yk_i^{2\alpha _i}$ in $T_{P^{\vee }}$ . However, $s_{i}\mathfrak {s}_y=\mathfrak {s}_y\big (\mathfrak {s}_y^{-\alpha _i^{\vee }}\big )^{\alpha _i}$ by (2.37), so $\big (\mathfrak {s}_y^{-\alpha _i^{\vee }}\big )^{\alpha _i}=k_i^{2\alpha _i}$ in $T_{\Lambda }$ . This in particular holds true for $\Lambda =P^{\vee }$ . Hence, $\mathfrak {s}_y^{\alpha _i^{\vee }}=k_i^{-2}$ .

Suppose that $s_iy\not =y$ . Then $\alpha _i(y)\not =0$ ; hence, $\eta (\alpha _i(y))+\eta (-\alpha _i(y))=0$ by the first equality of (5.1). It follows that $s_i\mathfrak {s}_y=\mathfrak {s}_{s_iy}$ .

The proof for $j=0$ is a bit more involved. First note that

(5.4) $$ \begin{align} s_\varphi\mathfrak{s}_y=\mathfrak{s}_{s_0y}\prod_{\alpha\in\Phi_0^+\setminus\Pi(s_\varphi)} k_\alpha^{(\eta(\alpha(s_\varphi y))-\eta(\alpha(s_0y)))\alpha} \prod_{\beta\in\Pi(s_\varphi)}k_\beta^{-(\eta(-\beta(s_\varphi y))+\eta(\beta(s_0y)))\beta} \end{align} $$

since $k_{s_\varphi \gamma }=k_\gamma $ for all $\gamma \in \Phi _0$ . Now $\alpha \in \Phi _0^+\setminus \Pi (s_\varphi )$ implies $\alpha (\varphi ^{\vee })=0$ by (2.18); hence,

$$\begin{align*}\eta(\alpha(s_\varphi y))-\eta(\alpha(s_0y))=\eta(\alpha(y))-\eta(\alpha(y))=0. \end{align*}$$

Hence, the product over $\alpha \in \Phi _0^+\setminus \Pi (s_\varphi )$ in (5.4) is equal to $1$ . By the explicit description (2.18) of $\Pi (s_\varphi )$ , we then have

$$\begin{align*}s_\varphi\mathfrak{s}_y=\mathfrak{s}_{s_0y}k_\varphi^{-(\eta(2-\varphi(y))+\eta(\varphi(y)))\varphi} \prod_{\beta\in\Phi_0^+:\, \beta(\varphi^{\vee})=1} k_\beta^{-(\eta(-\beta(s_\varphi y))+\eta(\beta(s_\varphi y)+1))\beta}. \end{align*}$$

The product over $\beta $ is equal to $1$ by (5.1), so we conclude that

$$\begin{align*}s_\varphi\mathfrak{s}_y=\mathfrak{s}_{s_0y}k_\varphi^{-(\eta(2-\varphi(y))+\eta(\varphi(y)))\varphi}. \end{align*}$$

Suppose that $s_0y=y$ . Then $\varphi (y)=1$ ; hence, $\eta (2-\varphi (y))+\eta (\varphi (y))=2\eta (1)=2$ and $s_\varphi \mathfrak {s}_y=\mathfrak {s}_yk_\varphi ^{-2\varphi }$ in $T_{\Lambda }$ . Similarly as before, this implies that $\mathfrak {s}_y^{-\varphi ^{\vee }}=k_\varphi ^{-2}$ .

Suppose that $s_0y\not =y$ . Then $\varphi (y)\not =1$ and $\eta (2-\varphi (y))+\eta (\varphi (y))=0$ (indeed, it is $0+0$ if $\varphi (y)\in \mathbb {R}\setminus \mathbb {Z}$ , it is $-1+1$ if $\varphi (y)\in \mathbb {Z}_{>1}$ and it is $1-1$ if $\varphi (y)\in \mathbb {Z}_{<1}$ ). Hence, $s_\varphi \mathfrak {s}_y=\mathfrak {s}_{s_0y}$ .

Proposition 5.4. Let $\Lambda \in \mathcal {L}$ . We have

$$\begin{align*}\mathfrak{s}_y=(Dw_y)\mathfrak{s}_{c_y} \end{align*}$$

in $T_{\Lambda }$ for all $y\in E$ .

Proof. Write $c:=c_y\in \overline {C}_+$ and $J:=\mathbf {J}(c)$ . Fix a reduced expression $w_y=s_{j_1}\cdots s_{j_\ell }$ for $w_y\in W^J$ and write $y_m:=s_{j_m}\cdots s_{j_\ell }c$ ( $1\leq m\leq \ell $ ) and $y_{\ell +1}:=c$ . Then $s_{j_{m}}y_{m+1}\not =y_m$ for $m=1\ldots ,\ell $ . Repeated application of Lemma 5.3(2) then gives $\mathfrak {s}_y=(Dw_y)\mathfrak {s}_c$ .

Recall from Proposition 3.7 that $L_{\Lambda ,J}=\mathfrak {s}_JT_{\Lambda ,J}$ for $\Lambda \in \mathcal {L}$ . Furthermore, for $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ,J}$ , we defined $\mathfrak {t}_y\in T_\Lambda $ ( $y\in \mathcal {O}_c$ ) by $\mathfrak {t}_y:=w_y\mathfrak {t}$ (see Definition 4.1).

Corollary 5.5. Let $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ,J}$ . For $y\in \mathcal {O}_c$ , we have

$$\begin{align*}\mathfrak{s}_y\mathfrak{t}_y=w_y(\mathfrak{s}_J\mathfrak{t}) \end{align*}$$

in $T_\Lambda $ .

Proof. This follows from

$$\begin{align*}w_y(\mathfrak{s}_J\mathfrak{t})=((Dw_y)\mathfrak{s}_J)(w_y\mathfrak{t})= \mathfrak{s}_y\mathfrak{t}_y \end{align*}$$

for $y\in \mathcal {O}_c$ , where the second equality follows from Proposition 5.4.

We will also need analogous results for scalar-valued functions on E, in particular for $\kappa _v: E\rightarrow \mathbf {F}^\times $ (see (4.10)). Recall the characteristic function $\chi _{\pm }: \Phi _0\rightarrow \{0,1\}$ of $\Phi _0^{\pm }$ .

Lemma 5.6.

  1. 1. For $u,v\in W_0$ and $y\in E$ ,

    (5.5) $$ \begin{align} \kappa_{uv}(y)=\kappa_{v}(y)\prod_{\alpha\in v^{-1}\Pi(u)}k_\alpha^{-\chi(\alpha) \eta(\chi(\alpha)\alpha(y))}, \end{align} $$
    where $\chi :=\chi _+-\chi _-$ .
  2. 2. If $y\in E^{\text {reg}}$ , then $\kappa _{uv}(y)=\kappa _u(vy)\kappa _v(y)$ for all $u,v\in W_0$ .

Proof. (1) Formula (5.5) follows from a direct computation, using that $\Pi (uv)$ is the disjoint union of the subsets $\Pi (v)\setminus (\Phi _0^+\cap (-v^{-1}\Pi (u)))$ and $\Phi _0^+\cap v^{-1}\Pi (u)$ .

(2) By (5.1), we have $\eta (-x)=-\eta (x)$ for $x\in \mathbb {R}^\times $ . Hence, for $y\in E^{\text {reg}}$ , formula (5.5) reduces to

$$\begin{align*}\kappa_{uv}(y)=\kappa_{v}(y)\prod_{\alpha\in v^{-1}\Pi(u)}k_\alpha^{-\eta(\alpha(y))}= \kappa_v(y)\kappa_u(vy). \end{align*}$$

We next introduce an affine version of $\kappa _v(y)$ . Set

(5.6) $$ \begin{align} k(y):=\prod_{\alpha\in\Phi_0^+}k_\alpha^{\frac{\eta(\alpha(y))}{2}}\qquad (y\in E), \end{align} $$

and define

(5.7) $$ \begin{align} k_w(y):=\frac{k(wy)}{k(y)}\qquad (y\in E, w\in W). \end{align} $$

The $k_w(y)$ satisfy the cocycle condition $k_{ww^\prime }(y)=k_w(w^\prime y)k_{w^\prime }(y)$ for $w,w^\prime \in W$ and $y\in E$ . Note, furthermore, that

$$\begin{align*}k_{\tau(\mu)}(y)=\frac{k(y+\mu)}{k(y)}=\prod_{\alpha\in\Phi_0^+}k_\alpha^{(\eta(\alpha(y)+\alpha(\mu))- \eta(\alpha(y)))/2} \end{align*}$$

for $y\in E$ and $\mu \in Q^{\vee }$ .

By Lemma 2.5, we have $k,k_w\in \mathcal {F}_\Sigma (E,\mathbf {F}^\times )$ for all $w\in W$ .

Lemma 5.7. For $0\leq j\leq r$ and $y\in E$ ,

(5.8) $$ \begin{align} k_{s_j}(y)=\kappa_{Ds_j}(y)\,\, \mbox{ if }\, \alpha_j(y)\not=0. \end{align} $$

Proof. We need to show that

(5.9) $$ \begin{align} \begin{aligned} k_{s_i}(y)&=k_i^{-\eta(\alpha_i(y))}\qquad\qquad\quad \mbox{ if }\,\, \alpha_i(y)\not=0,\\ k_{s_0}(y)&=\prod_{\alpha\in\Pi(s_\varphi)}k_\alpha^{-\eta(\alpha(y))}\qquad\,\, \mbox{ if }\,\, \alpha_0(y)\not=0 \end{aligned} \end{align} $$

for $1\leq i\leq r$ . The first line follows directly from Lemma 2.1 and (5.1). For the second line one computes, using (2.18),

$$\begin{align*}k(s_0y)=k_{\varphi}^{\eta(-\varphi(y)+2)/2}\prod_{\alpha\in\Pi(s_\varphi)\setminus\{\varphi\}} k_\alpha^{\eta(\alpha(s_\varphi y)+1)/2} \prod_{\beta\in\Phi_0^+\setminus\Pi(s_\varphi)}k_\beta^{\eta(\beta(y))/2}. \end{align*}$$

By (5.1), one then gets

$$\begin{align*}k_{s_0}(y)=k_{\varphi}^{(\eta(-\varphi(y)+2)+\eta(\varphi(y)))/2}\prod_{\alpha\in\Pi(s_\varphi)}k_\alpha^{-\eta(\alpha(y))}. \end{align*}$$

Since $\eta (2-x)+\eta (x)=0$ if $x\not =1$ (see the proof of Lemma 5.3), this reduces to the second formula in (5.9).

5.2 The $\mathbb {H}$ -action on $\mathcal {P}^{(c)}$

In this subsection, we give the proof of Theorem 4.5(1). Fix $c\in C^J$ and $\mathfrak {t}\in T_J$ . An important role in the proof is played by the following map.

Definition 5.8. Define the surjective linear map $\psi _{c,\mathfrak {t}}: \mathbb {H}\twoheadrightarrow \mathcal {P}^{(c)}$ by

(5.10) $$ \begin{align} \psi_{c,\mathfrak{t}}(x^\mu T_v Y^\nu):=\kappa_v(c)(\mathfrak{s}_c\mathfrak{t})^{-\nu} x^{\mu+vc} \end{align} $$

for $\mu ,\nu \in Q^{\vee }$ and $v\in W_0$ .

Note that $\psi _{c,\mathfrak {t}}(1)=x^c$ .

Lemma 5.9. Let $h^\prime \in \mathbb {H}$ . Then

$$ \begin{align*} \psi_{c,\mathfrak{t}}(hh^\prime)=\pi_{c,\mathfrak{t}}(h)\psi_{c,\mathfrak{t}}(h^\prime) \end{align*} $$

for $h=T_j$ ( $0\leq j\leq r$ ) and $h=x^\mu $ ( $\mu \in Q^{\vee }$ ).

Proof. For the duration of the proof, we write

(5.11) $$ \begin{align} t:=\mathfrak{s}_c \mathfrak{t}\in L_{J}. \end{align} $$

Recall that $\chi _{\pm }: \Phi _0\rightarrow \{0,1\}$ is the characteristic function of $\Phi _0^{\pm }$ , and $\chi :=\chi _+-\chi _-$ . Note that for $v\in W_0$ , we have $\chi (v^{-1}\alpha _i)=\pm 1$ iff $\ell (s_iv)=\ell (v)\pm 1$ by Lemma 2.1.

The lemma is immediate for $h=x^\nu $ ( $\nu \in Q^{\vee }$ ). We split the proof for $h=T_j$ in two cases.

Case 1: $h=T_i$ with $1\leq i\leq r$ .

This case is similar to the proof of [Reference Sahi, Stokman and Venkateswaran38, Thm. 3.7]; see, in particular, [Reference Sahi, Stokman and Venkateswaran38, Lemma 3.2]. We shortly discuss it here for completeness. Fix $v\in W_0$ and $\mu ,\nu \in Q^{\vee }$ . By the cross relation (2.40) in $\mathbb {H}$ and the fact that $T_iT_v=\chi _-(v^{-1}\alpha _i)(k_i-k_i^{-1})T_v+T_{s_iv}$ , we have for $v\in W_0$ and $\mu ,\nu \in Q^{\vee }$ ,

(5.12) $$ \begin{align}\begin{aligned} \psi_{c,\mathfrak{t}}(T_ix^\mu T_vY^\nu)&= \kappa_{s_iv}(c)t^{-\nu}x^{s_i\mu+s_ivc}\\ &\quad+\kappa_v(c)t^{-\nu}(k_i-k_i^{-1})\big(\nabla_i(x^\mu)x^{vc}+\chi_-(v^{-1}\alpha_i)x^{s_i\mu+vc}\big). \end{aligned}\end{align} $$

By Lemma 2.1, we have $0\leq \chi (v^{-1}\alpha _i)\alpha _i(vc)\leq 1$ since $c\in \overline {C}_+$ , and

$$\begin{align*}\kappa_{s_iv}(c)=\kappa_v(c)k_i^{-\chi(v^{-1}\alpha_i)\eta(\chi(v^{-1}\alpha_i)\alpha_i(vc))} \end{align*}$$

by (5.5). So (5.12) reduces to

(5.13) $$ \begin{align} \begin{aligned} \psi_{c,\mathfrak{t}}(T_ix^\mu T_vY^\nu)=\kappa_v(c)t^{-\nu}\Big(&k_i^{-\chi(v^{-1}\alpha_i)\eta(\chi(v^{-1}\alpha_i)\alpha_i(vc))} x^{s_i\mu+s_ivc}\\ &+ (k_i-k_i^{-1})\big(\nabla_i(x^\mu)x^{vc}+\chi_-(v^{-1}\alpha_i)x^{s_i\mu+vc}\big)\Big). \end{aligned} \end{align} $$

However,

(5.14) $$ \begin{align} \pi_{c,\mathfrak{t}}(T_i)\psi_{c,\mathfrak{t}}(x^\mu T_vY^\nu)= \kappa_v(c)t^{-\nu}\Big(k_i^{\chi_{\mathbb{Z}}(\alpha_i(vc))}x^{s_i\mu+s_ivc} +(k_i-k_i^{-1})\nabla_i(x^{\mu+vc})\Big), \end{align} $$

and

(5.15) $$ \begin{align} \nabla_i(x^{\mu+vc})= \begin{cases} \nabla_i(x^\mu)x^{vc}+x^{s_i\mu+vc}\quad &\mbox{ if }\,\, -1\leq \alpha_i(vc)<0,\\ \nabla_i(x^\mu)x^{vc}\quad &\mbox{ if }\,\, 0\leq\alpha_i(vc)<1,\\ \nabla_i(x^\mu)x^{vc}-x^{s_i\mu+s_ivc} \quad &\mbox{ if }\,\, \alpha_i(vc)=1. \end{cases} \end{align} $$

To show that (5.13) is equal to (5.14), we distinguish subcases depending on whether $vc$ lies on one of the three affine root hyperplanes $H_{i,m}:=\{z\in E\,\, | \,\, \alpha _i(z)=m\}$ ( $m\in \{-1,0,1\}$ ) or that it lies between $H_{i,m}$ and $H_{i,m+1}$ for $m=-1$ and $m=0$ . The cases $vc\in H_{i,-1}$ , and $vc$ lying between $H_{i,-1}$ and $H_{i,0}$ , are taken together as case 1a.

Case 1a: $-1\leq \alpha _i(vc)<0$ . Then $v^{-1}\alpha _i\in \Phi _0^-$ , $\eta (-\alpha _i(vc))= \chi _{\mathbb {Z}}(\alpha _i(vc))$ and $\chi _-(v^{-1}\alpha _i)=1$ . Taking (5.15) into account, it follows now directly that (5.13) is equal to (5.14).

Case 1b: $\alpha _i(vc)=0$ . Then $s_ivc=vc$ and $\eta (\chi (v^{-1}\alpha _i)\alpha _i(vc))=-1$ , but it is undetermined whether $v^{-1}\alpha _i$ is a positive or a negative root. The expression (5.13) simplifies to

$$ \begin{align*} \begin{aligned} \psi_{c,\mathfrak{t}}(&T_ix^\mu T_vY^\nu)\\ &=\kappa_v(c)t^{-\nu} \Big(k_i^{\chi(v^{-1}\alpha_i)}x^{s_i\mu+vc}+(k_i-k_i^{-1})\big(\nabla_i(x^\mu)x^{vc}+\chi_-(v^{-1}\alpha_i)x^{s_i\mu+vc}\big)\Big)\\ &=\kappa_v(c)t^{-\nu}\Big(k_ix^{s_i\mu+vc}+(k_i-k_i^{-1})\nabla_i(x^\mu)x^{vc}\Big). \end{aligned} \end{align*} $$

Note here that for $v^{-1}\alpha _i\in \Phi _0^-$ , the last equality follows by combining the first and third term, which has the effect that it flips the prefactor $k_i^{-1}$ of the first term to $k_i$ . Formula (5.15) now reduces this expression to (5.14).

Case 1c: $0<\alpha _i(vc)<1$ . Then $v^{-1}\alpha _i\in \Phi _0^+$ so $\chi _-(v^{-1}\alpha _i)=0$ , and furthermore, $\eta (\alpha _i(vc))=0=\chi _{\mathbb {Z}}(\alpha _i(vc))$ . Taking also (5.15) into account, it follows that (5.13) is equal to (5.14).

Case 1d: $\alpha _i(vc)=1$ . Then $\chi _-(v^{-1}\alpha _i)=0$ and $\eta (\alpha _i(vc))=1$ , so

(5.16) $$ \begin{align} \psi_{c,\mathfrak{t}}(T_ix^\mu T_vY^\nu)= \kappa_v(c)t^{-\nu}\Big(k_i^{-1}x^{s_i\mu+s_ivc}+(k_i-k_i^{-1})\nabla_i(x^\mu)x^{vc}\Big), \end{align} $$

while (5.14) by (5.15) becomes

$$\begin{align*}\pi_{c,\mathfrak{t}}(T_i)\psi_{c,\mathfrak{t}}(x^\mu T_vY^\nu)= \kappa_v(c)t^{-\nu}\Big(k_ix^{s_i\mu+s_ivc}+ (k_i-k_i^{-1})(\nabla_i(x^\mu)x^{vc}-x^{s_i\mu+s_ivc})\Big). \end{align*}$$

Moving the correction term of the divided difference operator to the front is altering the prefactor $k_i$ of $x^{s_i\mu +s_ivc}$ to $k_i^{-1}$ . This reduces the last expression to (5.16).

Case 2: $h=T_0$ . Fix $v\in W_0$ and $\mu ,\nu \in Q^{\vee }$ . By (2.44), we have

$$\begin{align*}T_0T_v=T_{s_\varphi v}Y^{-v^{-1}\varphi^{\vee}}+\chi_+(v^{-1}\varphi)(k_\varphi-k_\varphi^{-1})T_v \end{align*}$$

for $v\in W_0$ (recall here that $k_0=k_\varphi $ ). Combined with the cross relation (2.42) in $\mathbb {H}$ and the definition of $\psi _{c,\mathfrak {t}}$ , we get

$$ \begin{align*} \begin{aligned} \psi_{c,\mathfrak{t}}(T_0x^\mu T_vY^\nu)&= \kappa_{s_\varphi v}(c)t^{-\nu+v^{-1}\varphi^{\vee}}q_\varphi^{\varphi(\mu)} x^{s_\varphi\mu+s_\varphi vc}\\ &+ \kappa_v(c)t^{-\nu}(k_\varphi-k_\varphi^{-1}) \big(\nabla_0(x^\mu)x^{vc}+\chi_+(v^{-1}\varphi)q_\varphi^{\varphi(\mu)}x^{s_\varphi \mu+vc} \big). \end{aligned} \end{align*} $$

Now apply (5.5) to $\kappa _{s_\varphi v}(c)$ and use (2.18) to arrive at

(5.17) $$ \begin{align} \kappa_{s_\varphi v}(c)=\kappa_v(c)k_{\varphi}^{\chi(v^{-1}\varphi) \eta(\chi(v^{-1}\varphi)\varphi(vc))}\prod_{\alpha\in\Phi_0^+}k_\alpha^{\epsilon_v(\alpha)} \end{align} $$

with $\epsilon _v(\alpha ):=-\chi (v^{-1}\alpha )\eta (\chi (v^{-1}\alpha )\alpha (vc))\alpha (\varphi ^{\vee })$ for $\alpha \in \Phi _0$ . Since $\epsilon _v(-\alpha )=\epsilon _v(\alpha )$ for all $\alpha \in \Phi _0$ , the product over $\alpha \in \Phi _0^+$ in (5.17) is unchanged when taking a different choice of positive roots for $\Phi _0^+$ . Choosing $v\Phi _0^+$ as set of positive roots, we get

$$ \begin{align*} \begin{aligned} \kappa_{s_\varphi v}(c)&=\kappa_v(c)k_{\varphi}^{\chi(v^{-1}\varphi) \eta(\chi(v^{-1}\varphi)\varphi(vc))}\prod_{\alpha\in\Phi_0^+}k_\alpha^{-\eta(\alpha(c))\alpha(v^{-1}\varphi^{\vee})}\\ &=\kappa_v(c)k_{\varphi}^{\chi(v^{-1}\varphi) \eta(\chi(v^{-1}\varphi)\varphi(vc))}\mathfrak{s}_c^{-v^{-1}\varphi^{\vee}}, \end{aligned} \end{align*} $$

where the last equality follows from the explicit expression of $\mathfrak {s}_y$ (see Definition 5.1). By (5.11), we conclude that

(5.18) $$ \begin{align} \begin{aligned} \psi_{c,\mathfrak{t}}(T_0x^\mu T_vY^\nu)=\kappa_v(c)&t^{-\nu} \Big(\mathfrak{t}^{v^{-1}\varphi^{\vee}}k_\varphi^{\chi(v^{-1}\varphi)\eta(\chi(v^{-1}\varphi) \varphi(vc))}q_\varphi^{\varphi(\mu)}x^{s_\varphi\mu+s_\varphi vc}\\ &\,\,+ (k_\varphi-k_\varphi^{-1}) \big(\nabla_0(x^\mu)x^{vc}+\chi_+(v^{-1}\varphi)q_\varphi^{\varphi(\mu)}x^{s_\varphi \mu+vc} \big)\Big). \end{aligned} \end{align} $$

However,

(5.19) $$ \begin{align} \begin{aligned} \pi_{c,\mathfrak{t}}(T_0)&\psi_{c,\mathfrak{t}}(x^\mu T_vY^\nu)=\\ &=\kappa_v(c)t^{-\nu}\Big(\mathfrak{t}^{v^{-1}\varphi^{\vee}}k_\varphi^{\chi_{\mathbb{Z}}(\varphi(vc))}q_\varphi^{\varphi(\mu)}x^{s_\varphi \mu+s_\varphi vc} +(k_\varphi-k_\varphi^{-1})\nabla_0(x^{\mu+vc})\Big) \end{aligned} \end{align} $$

since $\mathfrak {t}_{\mu +vc}^{\varphi ^{\vee }}=q_\varphi ^{\varphi (\mu )}\mathfrak {t}^{v^{-1}\varphi ^{\vee }}$ by Definition 4.1. Furthermore,

(5.20) $$ \begin{align} \nabla_0(x^{\mu+vc})= \begin{cases} \nabla_0(x^\mu)x^{vc}-q_{\varphi}^{\varphi(\mu+vc)}x^{s_\varphi \mu+s_\varphi vc} \quad &\mbox{ if }\,\,\,\, \varphi(vc)=-1,\\ \nabla_0(x^\mu)x^{vc}\quad &\mbox{ if }\,\,\,\, -1<\varphi(vc)\leq 0,\\ \nabla_0(x^\mu)x^{vc}+q_{\varphi}^{\varphi(\mu)}x^{s_\varphi\mu+vc}\quad &\mbox{ if }\,\,\,\, 0<\varphi(vc)\leq 1. \end{cases} \end{align} $$

Now as in case 1, we prove that the expression (5.18) is equal to (5.19) by distinguishing various cases related to the position of $vc$ relative to the affine root hyperplanes $\{z\in E \,\, | \,\, \varphi (z)=m\}$ ( $m=-1,0,1$ ). Since there are various additional subtleties, we discuss the cases in detail.

Case 2a: $0<\varphi (vc)\leq 1$ . Then $v^{-1}\varphi \in \Phi _0^+$ ; hence, $\chi (v^{-1}\varphi )=1$ , $\eta (\varphi (vc))=\chi _{\mathbb {Z}}(\varphi (vc))$ and $\chi _+(v^{-1}\varphi )=1$ . The desired result now follows immediately from (5.20).

Case 2b: $\varphi (vc)=0$ . Then $\eta (\varphi (vc))=-1$ , $s_\varphi vc=vc$ , and $\mathfrak {t}^{v^{-1}\varphi ^{\vee }}=1$ by Lemma 4.3. Substitution into (5.18) gives

$$ \begin{align*} \begin{aligned} \psi_{c,\mathfrak{t}}(T_0x^\mu T_vY^\nu)= \kappa_v(c)t^{-\nu}&\Big(k_\varphi^{-\chi(v^{-1}\varphi)}q_{\varphi}^{\varphi(\mu)} x^{s_\varphi\mu+vc}\\ &\,\,+(k_\varphi-k_\varphi^{-1})\big(\nabla_0(x^\mu)x^{vc}+\chi_+(v^{-1}\varphi) q_\varphi^{\varphi(\mu)}x^{s_\varphi\mu+vc}\big)\Big)\\ =\kappa_v(c)t^{-\nu}&\Big(k_\varphi q_\varphi^{\varphi(\mu)}x^{s_\varphi\mu+vc} +(k_\varphi-k_\varphi^{-1})\nabla_0(x^\mu)x^{vc}\Big). \end{aligned} \end{align*} $$

Note here that for $v^{-1}\varphi \in \Phi _0^+$ , the second equality follows by combining the first and third term. By (5.20), this expression reduces to (5.19) (using once more $\mathfrak {t}^{v^{-1}\varphi ^{\vee }}=1$ ).

Case 2c: $-1<\varphi (vc)<0$ . Then $v^{-1}\varphi \in \Phi _0^-$ , $\eta (-\varphi (vc))=0$ and $\chi _+(v^{-1}\varphi )=0$ . Using (5.20), it follows that (5.18) equals (5.19).

Case 2d: $\varphi (vc)=-1$ . Then $v^{-1}\varphi \in \Phi _0^-$ , $\eta (-\varphi (vc))=1$ , $\chi _+(v^{-1}\varphi )=0$ , and $\mathfrak {t}^{v^{-1}\varphi ^{\vee }}=q_\varphi ^{-1}$ by Lemma 4.3. Hence,

(5.21) $$ \begin{align} \psi_{c,\mathfrak{t}}(T_0x^\mu T_vY^\nu)= \kappa_v(c)t^{-\nu}\Big(k_\varphi^{-1}q_\varphi^{\varphi(\mu+vc)} x^{s_\varphi\mu+s_\varphi vc}+(k_\varphi-k_\varphi^{-1})\nabla_0(x^\mu)x^{vc}\Big). \end{align} $$

By (5.20), the expression (5.19) becomes

(5.22) $$ \begin{align} \begin{aligned} \pi_{c,\mathfrak{t}}(T_0)\psi_{c,\mathfrak{t}}(x^\mu T_vY^\nu)= \kappa_v(c)&t^{-\nu}\Big(k_\varphi q^{\varphi(\mu+vc)}_\varphi x^{s_\varphi\mu+s_\varphi vc}\\ &+(k_\varphi-k_\varphi^{-1})\big(\nabla_0(x^\mu)x^{vc}-q_\varphi^{\varphi(\mu+vc)}x^{s_\varphi\mu+s_\varphi vc}\big)\Big). \end{aligned} \end{align} $$

Then (5.22) reduces to (5.21) by combining the first and third term in (5.22) into a single term.

Lemma 5.9 implies that $\text {ker}(\psi _{c,\mathfrak {t}})\subseteq \mathbb {H}$ is a left ideal; hence, $\mathbb {H}/\text {ker}(\psi _{c,\mathfrak {t}})$ is canonically a left $\mathbb {H}$ -module. Now consider $\mathcal {P}^{(c)}$ as left $\mathbb {H}$ -module by transporting the left $\mathbb {H}$ -action on $\mathbb {H}/\text {ker}(\psi _{c,\mathfrak {t}})$ through the linear isomorphism $\mathbb {H}/\text {ker}(\psi _{c,\mathfrak {t}})\overset {\sim }{\longrightarrow }\mathcal {P}^{(c)}$ induced by $\psi _{c,\mathfrak {t}}$ . By Lemma 5.9 again, its representation map is characterised by (4.11). This completes the proof of Theorem 4.5(1).

Corollary 5.10. View $\mathbb {H}$ as left $\mathbb {H}$ -module with respect to the left-regular $\mathbb {H}$ -action.

  1. 1. The surjective linear map $\psi _{c,\mathfrak {t}}:\mathbb {H}\twoheadrightarrow \mathcal {P}^{(c)}_{\mathfrak {t}}$ , defined by (5.10), is $\mathbb {H}$ -linear.

  2. 2. We have

    $$\begin{align*}\psi_{c,\mathfrak{t}}(h)=\pi_{c,\mathfrak{t}}(h)x^c\qquad \forall\, h\in\mathbb{H}. \end{align*}$$

    In particular, $\pi _{c,\mathfrak {t}}(T_v)x^c=\kappa _v(c)x^{vc}$ for all $v\in W_0$ .

Proof. (1) This is immediate from the paragraph preceding the corollary.

(2) By (1), we have

$$\begin{align*}\psi_{c,\mathfrak{t}}(h)=\pi_{c,\mathfrak{t}}(h)\psi_{c,\mathfrak{t}}(1)= \pi_{c,\mathfrak{t}}(h)x^c\qquad \forall\, h\in\mathbb{H}. \end{align*}$$

The second statement then follows from (5.10).

5.3 The $\mathbb {H}$ -module $\mathcal {P}^{(c)}_{\mathfrak {t}}$ as quotient of $\mathbb {M}^J_{\mathfrak {s}_c\mathfrak {t}}$

Take $c\in C^J$ and $\mathfrak {t}\in T_{J}$ . Then $\mathfrak {s}_c\mathfrak {t}\in L_J$ , so we have the one-dimensional representation $\chi _{J,\mathfrak {s}_c\mathfrak {t}}: \mathbb {H}_J^Y\rightarrow \mathbf {F}$ from (3.14).

Lemma 5.11. We have $\pi _{c,\mathfrak {t}}(h)x^c=\chi _{J,\mathfrak {s}_{c}\mathfrak {t}}(h)x^c$ for all $h\in \mathbb {H}_{J}^Y$ .

Proof. Write $t:=\mathfrak {s}_c\mathfrak {t}\in L_{J}$ for the duration of this proof.

For $\mu \in Q^{\vee }$ , we have by Corollary 5.10,

$$\begin{align*}\pi_{c,\mathfrak{t}}(Y^\mu)x^c=\psi_{c,\mathfrak{t}}(Y^\mu)=t^{-\mu}x^c=\chi_{J,t}(Y^\mu)x^c. \end{align*}$$

For $i\in J_0$ , we have $\kappa _{s_i}(c)=k_i$ and $s_ic=c$ since $\alpha _i(c)=0$ ; hence,

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_i))x^c=\pi_{c,\mathfrak{t}}(T_i)x^c= k_ix^c=\chi_{J,t}(\delta(T_i))x^c, \end{align*}$$

where we used Corollary 5.10(2) for the second equality.

Suppose that $0\in J$ , so that $s_0c=c$ and $\varphi (c)=1$ . We need to show that $\pi _{c,\mathfrak {t}}(\delta (T_0))x^c= k_{\varphi }x^c$ . Using $\delta (T_0^{-1})=x^{\varphi ^{\vee }}T_0^{-1}Y^{\varphi ^{\vee }}$ , we compute

$$ \begin{align*} \pi_{c,\mathfrak{t}}(\delta(T_0^{-1}))x^c=t^{-\varphi^{\vee}}\pi_{c,\mathfrak{t}}(x^{\varphi^{\vee}}T_0^{-1})x^c=q_\varphi k_\varphi t^{-\varphi^{\vee}}x^c, \end{align*} $$

where the second equality follows from a direct computation using that

$$\begin{align*}\pi_{c,\mathfrak{t}}(T_j)x^y=k_jx^y+\left(\frac{k_jx^{\alpha_j^{\vee}}-k_j^{-1}}{x^{\alpha_j^{\vee}}-1}\right) (x^{y-D\alpha_j(y)\alpha_j^{\vee}}-x^y)\,\,\, \mbox{ if }\, y\in\mathcal{O}_c\, \mbox{ and }\, \alpha_j(y)\in\mathbb{Z}. \end{align*}$$

Now $\varphi (c)=1$ and $\mathfrak {t}\in T_{J}$ implies that $\mathfrak {t}^{-\varphi ^{\vee }}=q_\varphi ^{-1}$ by Lemma 4.3, and $\mathfrak {s}_c^{-\varphi ^{\vee }}=k_\varphi ^{-2}$ by Lemma 5.3(1). Hence, $\pi _{c,\mathfrak {t}}(\delta (T_0^{-1}))x^c=k_\varphi ^{-1}x^c$ , as desired.

Corollary 5.12. There exists a unique epimorphism $\phi _{c,\mathfrak {t}}: \mathbb {M}^{J}_{\mathfrak {s}_c\mathfrak {t}}\twoheadrightarrow \mathcal {P}^{(c)}_{\mathfrak {t}}$ of $\mathbb {H}$ -modules such that

$$\begin{align*}\psi_{c,\mathfrak{t}}(h)=\phi_{c,\mathfrak{t}}\big(hm^{J}_{\mathfrak{s}_c\mathfrak{t}}\big)\qquad \forall\, h\in\mathbb{H}. \end{align*}$$

Furthermore,

(5.23) $$ \begin{align} \phi_{c,\mathfrak{t}}\big(x^\mu T_vm^{J}_{\mathfrak{s}_c\mathfrak{t}}\big)= \kappa_v(c)x^{\mu+vc} \qquad (\mu\in Q^{\vee},\, v\in W_0). \end{align} $$

Proof. The first part is an immediate consequence of Corollary 5.10(1) and Lemma 5.11. For (5.23), note that

$$\begin{align*}\phi_{c,\mathfrak{t}}\big(x^\mu T_vm^{J}_{\mathfrak{s}_c\mathfrak{t}}\big)= \psi_{c,\mathfrak{t}}(x^\mu T_v)=\kappa_v(c)x^{\mu+vc} \end{align*}$$

for $\mu \in Q^{\vee }$ and $v\in W_0$ , where the second equality follows from (5.10).

5.4 The parabolic Bruhat order on E

The restrictions of the Bruhat order on W to the subsets $W^J$ ( $J\subsetneq [0,r]$ ) of minimal coset representatives can be glued together to a partial order on E, which we call the parabolic Bruhat order on E. This subsection is devoted to a detailed study of its properties. In the next subsection, we will continue with the proof of Theorem 4.5 by showing that the action of $\delta (T_j)$ on $\mathcal {P}^{(c)}_{\mathfrak {t}}$ is triangular with respect to the basis $\{x^y\}_{y\in \mathcal {O}_c}$ of monomials, partially ordered by the restriction of the parabolic Bruhat order to the set $\mathcal {O}_c$ of quasi-exponents.

Let $\leq _B$ be the Bruhat order on W with respect to the simple reflections $s_0,\ldots ,s_r$ (see [Reference Humphreys21, Chpt. 5] for details). We write $w<_Bw^\prime $ if $w\leq _Bw^\prime $ and $w\not =w^\prime $ . The following two properties of $\leq _B$ are well known.

Lemma 5.13.

  1. 1. Let $w,w^\prime \in W$ and $0\leq j\leq r$ , satisfying $w<_Bw^\prime <_Bs_jw^\prime $ . Then we have $s_jw<_Bs_jw^\prime $ .

  2. 2. For $a\in \Phi $ and $w\in W$ , we have

    $$\begin{align*}ws_a<_Bw\,\,\, \Leftrightarrow\,\,\, a\in\Pi(w)\cup -\Pi(w)\,\,\,\Leftrightarrow\,\,\, \ell(ws_a)<\ell(w). \end{align*}$$

Proof. (1) See, for example, [Reference Humphreys21, Prop. 5.9].

(2) See [Reference Macdonald29, (2.3.3)].

Definition 5.14. For $y,z\in E$ , we write $y\leq z$ if $y\in Wz$ and $w_y\leq _Bw_z$ .

In other words, elements in E from different W-orbits are incomparable, and the restriction of $\leq $ to a W-orbit $\mathcal {O}_c$ ( $c\in C^J$ ) corresponds to the Bruhat order on $W^J$ through the natural bijection $\mathcal {O}_c\overset {\sim }{\longrightarrow } W^J$ , $y\mapsto w_y$ (which, in turn, can be regarded as the natural partial order on the coset space $W/W_c\simeq W^J$ induced by $\leq _B$ ). In particular, $\leq $ defines a partial order on E.

We will now establish an alternative description of the partial order $\leq $ .

Definition 5.15. Let $\alpha \in \Phi ^+_0$ and $y,z\in E$ . Write $y\prec _\alpha z$ if the following two conditions hold true:

  1. 1. $y=s_{a}z$ for some $a\in \Phi $ with $Da=\alpha $ ,

  2. 2. either $|\alpha (y)|<|\alpha (z)|$ or $\alpha (y)=-\alpha (z)>0$ .

Let $\prec $ be the transitive closure of the relations $\prec _\alpha $ ( $\alpha \in \Phi ^+_0$ ) on E.

Remark 5.16.

  1. 1. For $\alpha \in \Phi ^+_0$ , we have $y\prec _\alpha s_\alpha y$ iff $\alpha (y)>0$ .

  2. 2. Elements from different W-orbits in E are incomparable with respect to $\prec $ .

  3. 3. The relation $\prec $ on $Q^{\vee }$ coincides with the relation $\prec _{\Phi _0^+}$ introduced by Haiman in [Reference Haiman19, (5.14)].

Let $\text {sgn}: \mathbb {Z}\rightarrow \{-1,1\}$ be the sign function, so $\text {sgn}(m)=\frac {m}{|m|}$ for $m\not =0$ and $\text {sgn}(0)=1$ .

Proposition 5.17. Let $c\in \overline {C}_+$ , $w\in W$ and $a=(\alpha ,\ell )\in \Phi $ with $\alpha \in \Phi _0^+$ . The following three statements are equivalent.

  1. 1. $s_aw<_Bw$ and $a(wc)\not =0$ .

  2. 2. $\text {sgn}(\ell )a(wc)<0$ .

  3. 3. $s_awc\prec _\alpha wc$ .

Proof. (1) $\Leftrightarrow $ (2): By Lemma 5.13(2), the identity $w^{-1}s_aw=s_{w^{-1}a}$ , and the description (2.5) of $\Phi ^+$ , we have

$$\begin{align*}s_aw<_Bw\,\,\,\Leftrightarrow\,\,\,\text{sgn}(\ell)w^{-1}a\in\Phi^-. \end{align*}$$

Furthermore, for $b\in \Phi $ with $b(c)\not =0$ , we have $b\in \Phi ^-$ iff $b(c)<0$ , since $c\in \overline {C}_+$ . The result follows from these observations and the fact that $(w^{-1}a)(c)=a(wc)$ .

(2) $\Rightarrow $ (3): If $\ell =0$ , then $\alpha (wc)<0$ ; hence, $s_awc=s_\alpha wc\prec _\alpha wc$ by Remark 5.16(1). If $\ell \not =0$ , then $0<|\ell |<-\text {sgn}(\ell )\alpha (wc)$ , which implies that

$$\begin{align*}|\alpha(wc)|=-\text{sgn}(\ell)\alpha(wc)>\pm\text{sgn}(\ell)(\alpha(wc)+2\ell). \end{align*}$$

Here, the inequality for the minus-sign follows from $|\ell |>0$ , and the inequality for the plus-sign follows from $|\ell |<-\text {sgn}(\ell )\alpha (wc)$ . Hence, $|\alpha (wc)+2\ell |<|\alpha (wc)|$ , and we conclude that

$$\begin{align*}|\alpha(s_awc)|=|\alpha(wc)+2\ell|< |\alpha(wc)|. \end{align*}$$

Hence, $s_awc\prec _\alpha wc$ .

(3) $\Rightarrow $ (2): We have either $|\alpha (wc)+2\ell |<|\alpha (wc)|$ or $\alpha (wc)+2\ell =\alpha (wc)<0$ . In the second case, we have $\ell =0$ , and hence, $\text {sgn}(\ell )a(wc)=\alpha (wc)<0$ , as desired. Now suppose that $|\alpha (wc)+2\ell |<|\alpha (wc)|$ . Then $\alpha (wc)\not =0$ . If $\alpha (wc)>0$ , then $\ell \in \mathbb {Z}_{<0}$ and $-\alpha (wc)-2\ell <\alpha (wc)$ (i.e., $-a(wc)<0$ ). If $\alpha (wc)<0$ , then $\ell \in \mathbb {Z}_{>0}$ and $\alpha (wc)+2\ell <-\alpha (wc)$ (i.e., $a(wc)<0$ ). So again, we have $\text {sgn}(\ell )a(wc)<0$ , as desired.

Corollary 5.18. Let $w,w^\prime \in W$ with $w\leq _B w^\prime $ . Then $wc\preceq w^\prime c$ for all $c\in \overline {C}_+$ .

Proof. Let $c\in \overline {C}_+$ and suppose that $w<_Bw^\prime $ . Then there exists a chain

$$\begin{align*}w=:w_0^\prime<_Bw_1^\prime<_B\cdots <_Bw_{m-1}^\prime<_Bw_m^\prime:=w^\prime \end{align*}$$

with $w_j^\prime \in W$ satisfying $w_{j-1}^\prime =s_{b_j}w_j^\prime $ ( $1\leq j\leq m$ ) for affine roots $b_j\in \Phi $ with $\beta _j:=Db_j\in \Phi _0^+$ . If $b_j(w_{j-1}^\prime c)=0$ , then $w_{j-1}^\prime c=w_j^\prime c$ . If $b_j(w_{j-1}^\prime c)\not =0$ , then $w_{j-1}^\prime c\prec _{\beta _j}w_j^\prime c$ by Proposition 5.17. Hence, $wc\preceq w^\prime c$ .

Proposition 5.19. The relation $\preceq $ on E coincides with the partial order $\leq $ on E.

Proof. Let $y,z\in E$ . If $y\leq z$ , then $w_y\leq _Bw_z$ and $c_y=c_z$ . Write $c:=c_y$ . Then we get $y=w_yc\preceq w_zc=z$ by Corollary 5.18.

Suppose that $y\prec z$ . Then y and z lie in the same W-orbit by Remark 5.16(2); hence, $c_y=c_z$ . Write $c:=c_y$ . There exists a chain

$$\begin{align*}y=:y_0\prec_{\beta_1}y_1\prec_{\beta_2}\cdots\prec_{\beta_{m-1}}y_{m-1}\prec_{\beta_m}y_m:=z \end{align*}$$

with $\beta _1,\ldots ,\beta _m\in \Phi _0^+$ and $y_j\in \mathcal {O}_c$ satisfying $y_{j-1}=s_{b_j}y_j$ for some $b_j\in \Phi $ with $Db_j=\beta _j$ . Then $s_{b_j}w_{y_j}<_Bw_{y_j}$ ( $1\leq j\leq m$ ) by Proposition 5.17. Since $w_{y_{j-1}}c=y_{j-1}=s_{b_j}w_{y_j}c$ and $w_{y_{j-1}}$ is a minimal coset representative of $w_{y_{j-1}} W_c$ , we have $w_{y_{j-1}}\leq _B s_{b_j}w_{y_j}$ ; hence, $w_{y_{j-1}}<_{B}w_{y_j}$ ( $1\leq j\leq m$ ). This implies $w_y<_{B}w_z$ ; hence, $y<z$ .

Corollary 5.20. Let $y,z\in E$ and $0\leq j\leq r$ , and suppose that $y<z<s_jz$ . Then $s_jy<s_jz$ .

Proof. Let $c\in \overline {C}_+$ such that $y,z\in \mathcal {O}_c$ . Since $w_{s_jz}$ is a minimal coset representative of the coset $s_jw_z W_c$ , we have $w_{s_jz}\leq _B s_jw_z$ . By the assumption $z<s_jz$ , we have $w_z<_Bw_{s_jz}$ . Combining the two relations, we get $w_z<_Bs_jw_z$ . Since furthermore $y<z$ , we conclude that

$$\begin{align*}w_y<_Bw_z<_Bs_jw_z. \end{align*}$$

Hence, $s_jw_y<_Bs_jw_z$ by Lemma 5.13(1). Then $s_jy<s_jz$ by Corollary 5.18 and Proposition 5.19.

Proposition 5.21. Let $y\in E$ and $a\in \Phi ^+$ . Then $y<s_ay$ iff $a(y)>0$ .

Proof. Suppose that $y<s_ay$ . Then $a(y)\not =0$ and $w_y<_B w_{s_ay}$ . Writing $s_a w_y=w_{s_ay}w^\prime $ with $w^\prime \in W_{c_y}$ , we have

$$\begin{align*}w_y<_Bw_{s_ay}\leq_B w_{s_ay}w^\prime=s_a w_y \end{align*}$$

since $\ell (w_{s_ay}w^\prime )= \ell (w_{s_ay})+\ell (w^\prime )$ . So $w_y<_Bs_a w_y$ , which implies that $w_y^{-1}a\in \Phi ^+$ by [Reference Humphreys21, Prop. 5.7]. It follows that $a(y)=(w_y^{-1}a)(c_y)\geq 0$ . Since $a(y)\not =0$ , this forces $a(y)>0$ .

For the converse implication, let $a=(\alpha ,\ell )\in \Phi ^+$ such that $a(y)=\ell +\alpha (y)>0$ . The condition that a is a positive root implies that $\ell \in \mathbb {Z}_{\geq 0}$ if $\alpha \in \Phi _0^+$ , and $\ell \in \mathbb {Z}_{>0}$ if $\alpha \in \Phi _0^-$ . It follows that

$$\begin{align*}\alpha(s_ay)=-\alpha(y)-2\ell<0. \end{align*}$$

Hence, $|\alpha (s_ay)|=\alpha (y)+2\ell $ . Since $\ell \in \mathbb {Z}_{\geq 0}$ , we conclude that $\alpha (y)\leq |\alpha (s_ay)|$ . If $\alpha (y)=|\alpha (s_ay)|$ , then $\ell =0$ , so $a(y)=\alpha (y)>0$ . Then $\alpha (y)=-\alpha (s_\alpha y)=-\alpha (s_ay)>0$ . We conclude that $y\prec s_ay$ , and hence, $y<s_ay$ by Proposition 5.19. If $\alpha (y)<|\alpha (s_ay)|$ , then $\ell \in \mathbb {Z}_{>0}$ and

$$\begin{align*}-\alpha(y)=-(\alpha(y)+\ell)+\ell< (\alpha(y)+\ell)+\ell=|\alpha(s_ay)|. \end{align*}$$

Hence, $|\alpha (y)|<|\alpha (s_ay)|$ . Consequently, $y\prec s_ay$ , so $y<s_ay$ by Proposition 5.19.

Corollary 5.22. Let $y\in E$ and $a\in \Phi ^+$ . Then $s_ay<y$ iff $a\in \Pi (w_y^{-1})$ iff $a(y)<0$ .

Proof. Combine Proposition 2.4 and Proposition 5.21.

Corollary 5.23. Let $c\in C^J$ , $y\in \mathcal {O}_c$ and $j\in [0,r]$ .

  1. 1. If $\alpha _j(y)\not =0$ , then $w_{s_jy}=s_jw_y$ and

    $$\begin{align*}\ell(s_jw_y)=\ell(w_y)\pm 1\quad \Leftrightarrow\quad \alpha_j(y)\gtrless 0. \end{align*}$$
  2. 2. If $\alpha _j(y)=0$ , then $w_{s_jy}=w_y$ and $\ell (s_jw_y)=\ell (w_y)+1$ .

Proof. (1) If $\alpha _j(y)\not =0$ , then $s_jy\not =y$ , and so, $s_jw_y\not \in w_yW_{c}$ . Hence, $s_jw_y\in W^{J}$ by Lemma 2.2(2). Furthermore, $s_jw_yc=s_jy=w_{s_jy}c$ ; hence, $s_jw_y=w_{s_jy}$ . If $\alpha _j(y)>0$ , then $w_y<_Bs_jw_y$ by Proposition 5.21; hence, $\ell (s_jw_y)=\ell (w_y)+1$ by Lemma 5.13(2). Similarly, if $\alpha _j(y)<0$ , then $w_{s_jy}<_Bw_y$ by Corollary 5.22; hence, $\ell (s_jw_y)=\ell (w_y)-1$ by Lemma 5.13(2) again.

(2) If $\alpha _j(y)=0$ , then $s_jy=y$ , so the first equality is trivial. Then $s_jw_y\in w_yW_{c}$ ; hence, the second part of the statement follows from Lemma 2.2(2).

We conclude this subsection with the following lemma, which explicitly describes root strings with respect to $\prec _\alpha $ .

Lemma 5.24. Let $\alpha \in \Phi _0^+$ and $y\in \mathcal {O}_c$ . Then we have the following:

  1. 1. If $\alpha (y)\in \mathbb {R}_{<0}$ , then $s_\alpha y-\ell \alpha ^{\vee }\prec _\alpha y$ for $0\leq \ell \leq -1-\lfloor \alpha (y)\rfloor $ ,

  2. 2. If $\alpha (y)\in \mathbb {Z}_{>1}$ , then $y-\ell \alpha ^{\vee }\prec _\alpha y$ for $1\leq \ell \leq \alpha (y)-1$ ,

  3. 3. If $\alpha (y)\in \mathbb {R}_{>1}\setminus \mathbb {Z}_{>1}$ , then $s_\alpha y+\ell \alpha ^{\vee }\prec _\alpha y$ for $1\leq \ell \leq \lfloor \alpha (y)\rfloor $ .

Proof. (1) Write $y_\ell :=s_\alpha y-\ell \alpha ^{\vee }$ with $0\leq \ell \leq -1-\lfloor \alpha (y)\rfloor $ . Then $y_\ell =s_ay$ with $a:=(\alpha ,\ell )\in \Phi $ . If $\ell =0$ , then $\alpha (y_0)=-\alpha (y)>0$ , and so, $y_0\prec _\alpha y$ . If $1\leq \ell \leq -1-\lfloor \alpha (y)\rfloor $ , then

$$\begin{align*}\alpha(y)<2+2\lfloor\alpha(y)\rfloor-\alpha(y)\leq -2\ell-\alpha(y)=\alpha(y_\ell)<-\alpha(y). \end{align*}$$

Hence, $|\alpha (y_\ell )|<|\alpha (y)|$ , and so $y_\ell \prec _\alpha y$ .

(2) Write $y_\ell :=y-\ell \alpha ^{\vee }$ with $1\leq \ell \leq \alpha (y)-1$ . Then we have $y_\ell =s_ay$ with $a=(\alpha ,\ell -\alpha (y))\in \Phi $ since $\alpha (y)\in \mathbb {Z}$ , and $-\alpha (y)+2\leq \alpha (y)-2\ell =\alpha (y_\ell )\leq \alpha (y)-2$ . Hence, $|\alpha (y_\ell )|<|\alpha (y)|$ , and so, $y_\ell \prec _\alpha y$ .

(3) Write $y_\ell :=s_\alpha y+\ell \alpha ^{\vee }$ with $1\leq \ell \leq \lfloor \alpha (y)\rfloor $ . Then $y_\ell =s_ay$ with $a=(\alpha ,-\ell )\in \Phi $ , and

$$\begin{align*}-\alpha(y)<2-\alpha(y)\leq 2\ell-\alpha(y)=\alpha(y_\ell)\leq 2\lfloor \alpha(y)\rfloor-\alpha(y)<\alpha(y), \end{align*}$$

where the last inequality is a consequence of the fact that $\alpha (y)\not \in \mathbb {Z}$ . Hence, $|\alpha (y_\ell )|<|\alpha (y)|$ , and so, $y_\ell \prec _\alpha y$ .

5.5 Triangularity properties

Throughout this subsection, we fix $c\in C^J$ and $\mathfrak {t}\in T_{J}$ . We will analyse the triangularity properties of the action of $Y^\mu $ ( $\mu \in Q^{\vee }$ ) and $\delta (T_j)$ ( $0\leq j\leq r$ ) on $\mathcal {P}^{(c)}_{\mathfrak {t}}$ with respect to the basis $\{x^y\}_{y\in \mathcal {O}_c}$ of quasi-monomials, partially ordered by the parabolic Bruhat order on $\mathcal {O}_c$ .

Definition 5.25. For $f\in \mathbf {F}[E]$ and $y\in E$ , we write

(5.24) $$ \begin{align} f=dx^y+\text{l.o.t.} \end{align} $$

if f is of the form

$$\begin{align*}f=dx^y+\sum_{z<y}d_{z}x^z\qquad (d,d_z\in\mathbf{F}). \end{align*}$$

If $d\not =0$ , then y is called the leading exponent of f, and d its leading coefficient.

Remark 5.26. If $f\in \mathbf {F}[E]$ is of the form (5.24), then f is a quasi-polynomial. In fact, $z<y$ forces $z\in Wy$ , so that $f\in \mathcal {P}^{(c_y)}$ . Furthermore, for $y\in \mathcal {O}_c$ ( $c\in C^J$ ), we have

$$\begin{align*}\#\{z\in E \,\, | \,\, z<y\}<\infty \end{align*}$$

since $\#\{w\in W^J\,\, | \,\, w<_Bw_y\}<\infty $ .

We now first show that the linear operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) are triangular. For an affine root $a\in \Phi $ , define $G_{\mathfrak {t}}(a)\in \text {End}(\mathcal {P}^{(c)})$ by

$$\begin{align*}G_{\mathfrak{t}}(a)x^y:=k_a^{\chi_{\mathbb{Z}}(a(y))}x^y+ (k_a-k_a^{-1})\left(\frac{1-x^{\lfloor Da(y)\rfloor a^{\vee}}}{1-x^{-a^{\vee}}}\right)s_{a,\mathfrak{t}}x^y\qquad (y\in\mathcal{O}_{c}). \end{align*}$$

Using Lemma 4.2 and the fact that $w_{\mathfrak {t}}x^y$ for $w\in W$ is a scalar multiple of $x^{(Dw)y}$ , one shows that

(5.25) $$ \begin{align} G_{\mathfrak{t}}(\alpha_j)=s_{j,\mathfrak{t}}\pi_{c,\mathfrak{t}}(T_j)\qquad (j=0,\ldots,r) \end{align} $$

and

(5.26) $$ \begin{align} G_{\mathfrak{t}}(wa)=w_{\mathfrak{t}}G_{\mathfrak{t}}(a)w_{\mathfrak{t}}^{-1}\qquad (w\in W,\,\, a\in\Phi) \end{align} $$

as linear operators on $\mathcal {P}^{(c)}$ .

Lemma 5.27. Fix $a=(\alpha ,\ell )\in \Phi $ with $\alpha \in \Phi _0^+$ and $\ell \in \mathbb {Z}$ . Then

$$\begin{align*}G_{\mathfrak{t}}(a)x^y=k_\alpha^{-\eta(\alpha(y))}x^y+\text{l.o.t.} \end{align*}$$

for all $y\in \mathcal {O}_{c}$ .

Proof. Note that

$$ \begin{align*} k_\alpha^{-\eta(\alpha(y))}= \begin{cases} k_a^{-1}\quad &\mbox{ if }\alpha(y)\in\mathbb{Z}_{>0},\\ k_a^{\chi_{\mathbb{Z}}(a(y))}\quad &\text{ if } \alpha(y)\in\mathbb{R}\setminus\mathbb{Z}_{>0}. \end{cases} \end{align*} $$

Keeping in mind Lemma 4.3, one then obtains by direct computations,

$$ \begin{align*} \begin{aligned} G_{\mathfrak{t}}(a)x^y&-k_\alpha^{-\eta(\alpha(y))}x^y\\ &= \begin{cases} (k_a-k_a^{-1})(1+x^{-a^{\vee}}+\cdots+x^{(1+\lfloor\alpha(y)\rfloor)a^{\vee}})s_{a,\mathfrak{t}}x^{y} &\mbox{ if } \alpha(y)\in\mathbb{R}_{<0},\\ 0 &\mbox{ if } 0\leq \alpha(y)\leq 1,\\ (k_a^{-1}-k_a)(x^{-a^{\vee}}+x^{-2a^{\vee}}+\cdots+x^{(1-\alpha(y))a^{\vee}})x^y &\mbox{ if } \alpha(y)\in\mathbb{Z}_{>1},\\ (k_a^{-1}-k_a)(x^{a^{\vee}}+x^{2a^{\vee}}+\cdots+x^{\lfloor \alpha(y)\rfloor a^{\vee}})s_{a,\mathfrak{t}}x^{y} \,\,&\mbox{ if } \alpha(y)\in \mathbb{R}_{>1}\setminus\mathbb{Z}_{>1}. \end{cases} \end{aligned} \end{align*} $$

Since $x^{a^{\vee }}=q_\alpha ^\ell x^{\alpha ^{\vee }}$ and $s_{a,\mathfrak {t}}x^y=\mathfrak {t}_y^{-\ell \alpha ^{\vee }}x^{y-\alpha (y)\alpha ^{\vee }}$ by (4.4), the result now follows from Lemma 5.24.

Proposition 5.28. For $\mu \in Q^{\vee }$ and $y\in \mathcal {O}_{c}$ , we have

$$\begin{align*}\pi_{c,\mathfrak{t}}(Y^\mu)x^y=(\mathfrak{s}_y\mathfrak{t}_y)^{-\mu}x^y+\text{l.o.t.} \end{align*}$$

in $\mathcal {P}^{(c)}$ .

Proof. It suffices to prove the proposition for $\mu \in Q^{\vee }\cap \overline {E}_+$ .

If $\tau (\mu )=s_{j_1}\cdots s_{j_m}$ is a reduced expression and $b_1,\ldots ,b_m$ are the associated positive affine roots in $\Pi (\tau (\mu ))$ (see (2.12)), then

$$\begin{align*}\pi_{c,\mathfrak{t}}(Y^\mu)=\tau(\mu)_{\mathfrak{t}}G_{\mathfrak{t}}(b_1)\cdots G_{\mathfrak{t}}(b_m) \end{align*}$$

as linear operators on $\mathcal {P}^{(c)}$ by the equivariance properties (5.25) and (5.26) of $G_{\mathfrak {t}}(a)$ . Furthermore, since $\mu \in Q^{\vee }\cap \overline {E}_+$ , we have

$$ \begin{align*} \Pi(\tau(\mu))= \bigsqcup_{\alpha\in\Phi^+_0} \{(\alpha,\ell)\,\, | \,\, 0\leq \ell<\alpha(\mu)\}. \end{align*} $$

Combined with Lemma 5.27, we get for $y\in \mathcal {O}_{c}$ ,

$$ \begin{align*} \begin{aligned} \pi_{c,\mathfrak{t}}(Y^\mu)x^y&= \left( \prod_{a\in\Pi(\tau(\mu))}k_{Da}^{-\eta(Da(y))}\right)\tau(\mu)_{\mathfrak{t}}x^y+\text{ l.o.t.}\\ &= \left(\prod_{\alpha\in\Phi_0^+}k_\alpha^{-\eta(\alpha(y))\alpha(\mu)}\right)\mathfrak{t}_y^{-\mu}x^y+\text{l.o.t.}\\ &=(\mathfrak{s}_y\mathfrak{t}_y)^{-\mu}x^y+\text{ l.o.t.} \end{aligned} \end{align*} $$

in $\mathcal {P}^{(c)}$ , as desired.

Recall $k_w\in \mathcal {F}_\Sigma (E,\mathbf {F}^\times )$ ( $w\in W$ ), defined by (5.7).

Proposition 5.29. For $c\in C^J$ and $w\in W^J$ , we have

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_{w^{-1}}))x^c=k_w(c) x^{wc}+\text{l.o.t.} \end{align*}$$

(here $wc$ is the action of w on $c\in E$ by translations and reflections).

Proof. Let $w=s_{j_1}\cdots s_{j_m}$ be a reduced expression for $w\in W^{J}$ , so that

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_{w^{-1}}))x^c= \pi_{c,\mathfrak{t}}(\delta(T_{j_1}))\cdots\pi_{c,\mathfrak{t}}(\delta(T_{j_m}))x^c. \end{align*}$$

Write $\Pi (w)=\{b_1,\ldots ,b_m\}$ with $b_i:=s_{j_m}\cdots s_{j_{i+1}}\alpha _{j_i}$ ( $1\leq i<m$ ) and $b_m:=\alpha _{j_m}$ . Then

$$\begin{align*}\alpha_{j_i}(s_{{j_{i+1}}}\cdots s_{j_{m}}c)=b_i(c)>0\qquad (i=1,\ldots,m) \end{align*}$$

since $w\in W^{J}$ . It thus suffices to show for $j=0,\ldots ,r$ that

(5.27) $$ \begin{align} \pi_{c,\mathfrak{t}}(\delta(T_j))x^y=k_{s_j}(y)x^{s_jy}+\text{l.o.t.}\qquad \mbox{ if }\,\,\, y\in\mathcal{O}_c\, \mbox{ and }\, \alpha_j(y)>0. \end{align} $$

For $y\in \mathcal {O}_c$ and $1\leq i\leq r$ , we have

(5.28) $$ \begin{align} \pi_{c,\mathfrak{t}}(T_i^{-1})x^y= \begin{cases} k_i^{-1}x^{s_iy}+\text{ l.o.t.} \qquad &\mbox{ if }\,\, \alpha_i(y)\in \mathbb{Z}_{\geq 0},\\ k_i^{\chi_{\mathbb{Z}}(\alpha_i(y))}x^{s_iy}+\text{ l.o.t.}\qquad &\mbox{ otherwise}. \end{cases} \end{align} $$

This follows in a similar way as Lemma 5.27, using as extra ingredients Proposition 5.21 and the fact that

(5.29) $$ \begin{align} \lfloor z\rfloor + \lfloor -z\rfloor= \begin{cases} -1\qquad &\mbox{ if } z\in\mathbb{R}\setminus\mathbb{Z},\\ 0\qquad &\mbox{ if } z\in\mathbb{Z} \end{cases} \end{align} $$

(below we give the similar, but more involved, proof of the variant (5.32) of formula (5.28) for the simple reflection $s_0$ ). As an easy consequence of Proposition 5.21 and the Hecke relations (2.39) for $T_i$ , we conclude that for $1\leq i\leq r$ and $y\in \mathcal {O}_c$ with $\alpha _i(y)>0$ ,

(5.30) $$ \begin{align} \pi_{c,\mathfrak{t}}(T_i^{\pm 1})x^y=k_i^{-\eta(\alpha_i(y))}x^{s_iy} +\text{ l.o.t.} \end{align} $$

Then (5.27) for $j>0$ follows from (5.8), (5.30) and the fact that $\delta (T_i)=T_i$ for $1\leq i\leq r$ .

For (5.27) with $j=0$ , we first note that for $U_0=q_\varphi ^{-1}x^{\varphi ^{\vee }}T_0^{-1}\in \mathbb {H}$ (see (2.48)),

(5.31) $$ \begin{align} \pi_{c,\mathfrak{t}}(U_0^{-1})x^y=k_\varphi^{\chi_{\mathbb{Z}}(\alpha_0(y))}\mathfrak{t}_y^{\varphi^{\vee}} x^{s_{\varphi}y-\alpha_0^{\vee}}+ (k_\varphi-k_\varphi^{-1})\,\left(\frac{1-x^{-\lfloor \alpha_0(y)+1\rfloor \alpha_0^{\vee}}}{1-x^{\alpha_0^{\vee}}} \right)\,x^{y+\alpha_0^{\vee}} \end{align} $$

for $y\in \mathcal {O}_{c}$ . Here, we used that $\mathfrak {t}_{y-\varphi ^{\vee }}^{\varphi ^{\vee }}=q_\varphi ^{-2}\mathfrak {t}_y^{\varphi ^{\vee }}$ (see Definition 4.1), and the fact that $q_\varphi x^{-\varphi ^{\vee }}=x^{\alpha _0^{\vee }}$ . We will now first show that for $y\in \mathcal {O}_{c}$ ,

(5.32) $$ \begin{align} \pi_{c,\mathfrak{t}}(U_0^{-1})x^y= \begin{cases} k_\varphi^{-1}q_\varphi^{-\alpha_0(y)}x^{s_0y}+\text{ l.o.t.} \qquad &\mbox{ if } \varphi(y)\in\mathbb{Z}_{\leq 1},\\ k_\varphi^{\chi_{\mathbb{Z}}(\varphi(y))}\mathfrak{t}_y^{\varphi^{\vee}}q_\varphi^{-1}x^{s_{0}y}+ \text{ l.o.t.}\qquad &\mbox{ otherwise} \end{cases} \end{align} $$

(note here that $x^{s_0y}=x^{\tau (\varphi ^{\vee })s_\varphi y}=x^{s_\varphi y+\varphi ^{\vee }}=q_\varphi x^{s_\varphi y-\alpha _0^{\vee }}$ ). If $\varphi (y)\in \mathbb {Z}_{\leq 1}$ , then $\alpha _0(y)\geq 0$ , and it follows from (5.31) and Lemma 4.3 that

$$ \begin{align*} \begin{aligned} \pi_{c,\mathfrak{t}}(U_0^{-1})x^y&= k_{\varphi}x^{y-\alpha_0(y)\alpha_0^{\vee}}+(k_\varphi^{-1}-k_\varphi)\big( 1+x^{\alpha_0^{\vee}}+\cdots+x^{\alpha_0(y)\alpha_0^{\vee}}\big)x^{y-\alpha_0(y)\alpha_0^{\vee}}\\ &=k_\varphi^{-1}x^{y-\alpha_0(y)\alpha_0^{\vee}}+(k_\varphi^{-1}-k_\varphi)\big(1+x^{-\alpha_0^{\vee}}+\cdots+ x^{(1-\alpha_0(y))\alpha_0^{\vee}}\big)x^y, \end{aligned} \end{align*} $$

with the second equation interpreted as $k_\varphi ^{-1}x^{y}$ when $\alpha _0(y)=0$ .

When $\varphi (y)=1$ (i.e., $\alpha _0(y)=0$ ), it is now clear that (5.32) is correct. If $\varphi (y)\in \mathbb {Z}_{<1}$ , then $\alpha _0(y)\in \mathbb {Z}_{>0}$ , and hence, $y<s_0y$ by Proposition 5.21. In particular, for $\varphi (y)=0$ , we get

$$\begin{align*}\pi_{c,\mathfrak{t}}(U_0^{-1})x^y=k_\varphi^{-1}x^{y-\alpha_0(y)\alpha_0^{\vee}}+ (k_\varphi^{-1}-k_\varphi)x^y=k_\varphi^{-1}q_\varphi^{-\alpha_0(y)}x^{s_0y}+\text{l.o.t.}, \end{align*}$$

as desired. If $\varphi (y)\in \mathbb {Z}_{<0}$ , then Lemma 5.24(1) and Proposition 5.19 give $y+\ell \varphi ^{\vee }<y$ for $1\leq \ell \leq -\varphi (y)$ . In addition, $y<s_0y$ by Proposition 5.21, and so, (5.32) is correct. Suppose that $\varphi (y)\in \mathbb {R}_{<1}\setminus \mathbb {Z}_{<1}$ . Then (5.31) gives

$$\begin{align*}\pi_{c,\mathfrak{t}}(U_0^{-1})x^y=\mathfrak{t}_y^{\varphi^{\vee}} x^{s_{\varphi}y-\alpha_0^{\vee}}+(k_\varphi^{-1}-k_\varphi)\big(1+x^{-\alpha_0^{\vee}}+\cdots +x^{-\lfloor\alpha_0(y)\rfloor\alpha_0^{\vee}}\big)x^y. \end{align*}$$

We have $\varphi (s_0y)=\alpha _0(y)+1=2-\varphi (y)\in \mathbb {R}_{>1}\setminus \mathbb {Z}_{>1}$ , and hence, $y+\ell \varphi ^{\vee }<s_0y$ for $0\leq \ell \leq -\lfloor \varphi (y)\rfloor $ by Lemma 5.24(3) and Proposition 5.19. Furthermore, $\lfloor \alpha _0(y)\rfloor =-\lfloor \varphi (y)\rfloor $ by (5.29), so (5.32) is correct. If $1<\varphi (y)\leq 2$ , then $\lfloor \alpha _0(y)+1\rfloor =0$ , and (5.32) immediately follows from (5.31). Finally, we consider the case $\varphi (y)\in \mathbb {R}_{>2}$ . Then $1+\alpha _0(y)\in \mathbb {R}_{<0}$ , and so (5.31) gives

$$\begin{align*}\pi_{c,\mathfrak{t}}(U_0^{-1})x^y= k_\varphi^{\chi_{\mathbb{Z}}(\alpha_0(y))}\mathfrak{t}_y^{\varphi^{\vee}}x^{s_{\varphi}y-\alpha_0^{\vee}} +(k_\varphi-k_\varphi^{-1})\big(x^{\alpha_0^{\vee}}+x^{2\alpha_0^{\vee}}+ \cdots+x^{(-2-\lfloor-\varphi(y)\rfloor)\alpha_0^{\vee}}\big)x^y. \end{align*}$$

Now $\varphi (s_0y)\in \mathbb {R}_{<0}$ , so Lemma 5.24(1) and Proposition 5.19 give $y-\ell \varphi ^{\vee }<s_0y$ for $1\leq \ell \leq -2-\lfloor -\varphi (y)\rfloor $ . This completes the proof of (5.32).

Fix $y\in \mathcal {O}_c$ with $\alpha _0(y)>0$ . Formula (5.32), the Hecke relation for $U_0$ , Lemma 4.3 and Proposition 5.21 imply that

(5.33) $$ \begin{align} \pi_{c,\mathfrak{t}}(U_0^{\pm 1})x^y=k_\varphi^{-\eta(\alpha_0(y))}\mathfrak{t}_y^{\varphi^{\vee}}q_\varphi^{-1}x^{s_0y} +\text{ l.o.t.} \end{align} $$

By the formula $\delta (T_0)=q_\varphi ^{-1}Y^{-\varphi ^{\vee }}U_0^{-1}$ , (5.33) and Proposition 5.28, we conclude that

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_0))x^y=k_\varphi^{-\eta(\alpha_0(y))} (\mathfrak{s}_{s_0y}\mathfrak{t}_{s_0y})^{\varphi^{\vee}}\mathfrak{t}_y^{\varphi^{\vee}}q_\varphi^{-2}x^{s_0y}+\text{ l.o.t.} \end{align*}$$

But $\alpha _0(y)>0$ implies $s_0y\not =y$ ; hence, Lemma 5.3(2) and Corollary 5.5 give

$$\begin{align*}(\mathfrak{s}_{s_0y}\mathfrak{t}_{s_0y})^{\varphi^{\vee}}=(s_\varphi\mathfrak{s}_y)^{\varphi^{\vee}}(s_0\mathfrak{t}_{y})^{\varphi^{\vee}} =q_\varphi^2\mathfrak{s}_y^{-\varphi^{\vee}}\mathfrak{t}_y^{-\varphi^{\vee}}. \end{align*}$$

So we obtain

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_0))x^y=\mathfrak{s}_y^{-\varphi^{\vee}}k_\varphi^{\eta(\varphi(y))} x^{s_0y}+\text{ l.o.t.}, \end{align*}$$

where we have also used (5.1) to rewrite the $k_\varphi $ -factor. Now observe that

$$\begin{align*}\mathfrak{s}_y^{-\varphi^{\vee}}=k_\varphi^{-\eta(\varphi(y))} \prod_{\alpha\in\Pi(s_\varphi)}k_\alpha^{-\eta(\alpha(y))} \end{align*}$$

by Definition 5.1 and (2.18). We conclude that

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_0))x^y=\left(\prod_{\alpha\in\Pi(s_\varphi)}k_\alpha^{-\eta(\alpha(y))} \right)x^{s_0y}+\text{l.o.t.} \end{align*}$$

The proof of (5.27) for $j=0$ now follows from (5.9).

5.6 Completion of the proof

Fix $c\in C^J$ and $\mathfrak {t}\in T_{J}$ . To complete the proof of Theorem 4.5 it remains to show that the epimorphism $\phi _{c,\mathfrak {t}}: \mathbb {M}^{J}_{\mathfrak {s}_c\mathfrak {t}}\twoheadrightarrow \mathcal {P}^{(c)}_{\mathfrak {t}}$ of $\mathbb {H}$ -modules, defined in Corollary 5.12, is an isomorphism.

Note that $\phi _{c,\mathfrak {t}}$ maps the standard basis element $m_{w,\mathfrak {s}_c\mathfrak {t}}^{J}=\delta (T_{w^{-1}}) m^{J}_{\mathfrak {s}_c\mathfrak {t}}$ ( $w\in W^{J}$ ) of $\mathbb {M}^{J}_{\mathfrak {s}_c\mathfrak {t}}$ to

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_{w^{-1}}))x^c=k_w(c)x^{wc}+\text{l.o.t.} \end{align*}$$

It follows that $\{\phi _{c,\mathfrak {t}}(m_{w,\mathfrak {s}_c\mathfrak {t}}^{J}\big )\,\, | \,\, w\in W^{J}\}$ is linear independent in $\mathcal {P}^{(c)}$ ; hence, indeed, $\phi _{c,\mathfrak {t}}$ is an isomorphism of $\mathbb {H}$ -modules.

6 Quasi-polynomial generalisations of Macdonald polynomials

In this section, J will be a fixed proper subset of $[0,r]$ . We show that for $c\in C^J$ , the corresponding commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) are simultaneously diagonalisable under suitable restrictions on the character $\mathfrak {t}\in T_J$ . We study their simultaneous eigenfunctions, which are quasi-polynomial generalisations of the Macdonald polynomials.

6.1 Simple spectrum conditions

Let $c\in C^J$ . We will construct in this section simultaneous eigenfunctions for the commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) when the character $\mathfrak {t}$ lies in the following subset of $T_J$ .

Definition 6.1. Denote by $T_J^\prime \subseteq T_J$ the set of elements $\mathfrak {t}\in T_J$ for which the map

(6.1) $$ \begin{align} W^J\rightarrow T,\qquad w\mapsto w(\mathfrak{s}_J\mathfrak{t}) \end{align} $$

is injective.

If $c\in C^J$ , then $W^J=\{w_y\}_{y\in \mathcal {O}_c}$ , and the condition can be rephrased as the requirement that

$$\begin{align*}\mathcal{O}_c\rightarrow T,\qquad y\mapsto \mathfrak{s}_y\mathfrak{t}_y=\mathfrak{t}_y\prod_{\alpha\in\Phi_0^+}k_\alpha^{\eta(\alpha(y))\alpha} \end{align*}$$

is injective, where $\mathfrak {t}_y=w_y\mathfrak {t}$ (see Corollary 5.5). We will now show that for $J=I\subseteq [1,r]$ , we have $T_I^\prime \not =\emptyset $ under generic conditions on $k_a$ , and that for $0\in J\subsetneq [0,r]$ , we have $T_J^\prime \not =\emptyset $ if we in addition assume that $\mathbf {F}$ contains a $(2h)^{th}$ root of q.

In the classical context of Cherednik’s basic representation, we have $J=[1,r]$ , $T_{[1,r]}=\{1_T\}$ , $C^{[1,r]}=\{0\}$ and $\mathcal {O}_0=Q^{\vee }$ . The requirement that $1_T\in T_{[1,r]}^\prime $ then amounts to generic conditions on $k_a$ , since q is not a root of unity and

$$\begin{align*}\mathfrak{s}_\mu(1_T)_\mu=q^\mu\prod_{\alpha\in\Phi_0^+}k_\alpha^{\eta(\alpha(\mu))\alpha}\qquad (\mu\in Q^{\vee}) \end{align*}$$

(here, we use that $(1_T)_\mu =q^\mu $ , in view of (4.2)).

If $J=I\subseteq [1,r]$ , then $W_I$ is already a parabolic subgroup of $W_0$ , and we denote by $W_0^I$ the resulting minimal coset representatives of $W_0/W_I$ . For $c\in C^I$ , each $y\in \mathcal {O}_c$ can be uniquely written as $y=\mu +vc$ with $\mu \in Q^{\vee }$ and $v\in W_0^I$ , and

$$\begin{align*}\mathfrak{s}_y\mathfrak{t}_{y}=q^\mu(v\mathfrak{t})\prod_{\alpha\in\Phi_0^+}k_\alpha^{\eta(\alpha(\mu+vc))\alpha}. \end{align*}$$

For generic $k_a$ , the maps $Q^{\vee }\rightarrow T$ , $\mu \mapsto q^\mu \prod _{\alpha \in \Phi _0^+}k_\alpha ^{\eta (\alpha (\mu +vc))\alpha }$ ( $v\in W_0^I$ ) are injective, and the requirement $\mathfrak {t}\in T_I^\prime $ results in generic conditions on $\mathfrak {t}\in T_I$ .

Finally, consider the general case $J\subsetneq [0,r]$ when $\mathbf {F}$ contains a $(2h)^{th}$ root $q^{\frac {1}{2h}}$ of q. Recall that this assumption on q ensures that $T_J\not =\emptyset $ by Lemma 3.2. In fact, one can construct elements in $T_J$ as follows. For $y\in \frac {1}{2h}P^{\vee }$ , let $q^{y}\in T$ be the character which takes the value

$$\begin{align*}q^{\langle y,\alpha^{\vee}\rangle}=q_\alpha^{\alpha(y)} \end{align*}$$

at the coroot $\alpha ^{\vee }\in \Phi _0^{\vee }$ . Using the table [Reference Humphreys20, Table 2, Chpt. 3] containing the explicit values of the expansion coefficients $n_i(\varphi )$ of $\varphi $ as positive integral linear combination of simple roots, as well as the corresponding explicit values $h=1+\sum _{i=1}^rn_i(\varphi )$ for the Coxeter number, it follows by a straightforward case-by-case check that $C^J\cap \frac {1}{2h}P^{\vee }\not =\emptyset $ (cf. the proof Lemma 3.2). Then for

(6.2) $$ \begin{align} \lambda_J\in C^J\cap\tfrac{1}{2h}P^{\vee}, \end{align} $$

we have $q^{\lambda _J}\in T_J$ and $w_yq^{\lambda _J}=q^y$ ( $y\in \mathcal {O}_{\lambda _J}$ ). Since the values of $\mathfrak {s}_y=v_y^{-1}\mathfrak {s}_{J}$ ( $y\in \mathcal {O}_{\lambda _J}$ ) are monomials in ${}^{\text {sh}}k$ and ${}^{\text {lg}}k$ (with possibly negative exponents), we have $q^{\lambda _J}\in T_J^\prime $ for generic $k_a$ .

Remark 6.2. Let $P\subseteq E^*$ be the weight lattice of $\Phi _0$ . Initially it is actually a full lattice in $(E^\prime )^*$ , which we view as a sublattice of $E^*$ by declaring the weights to be zero on $E_{\text {co}}$ . Let $\{\varpi _i\}_{i=1}^r\subset E^*$ be the fundamental weights with respect to $\Delta _0$ . Write $I^{\text {co}}:=[1,r]\setminus I$ for a subset $I\subseteq [1,r]$ . For a subset $J\subsetneq [0,r]$ and $\mathbf {z}:=(z_i)_{i\in J_0^{\text {co}}}\in (\mathbf {F}^\times )^{\#J_0^{\text {co}}}$ , set

$$\begin{align*}\mathfrak{t}(\mathbf{z}):=\prod_{j\in J_0^{\text{co}}}z_j^{\varpi_j}. \end{align*}$$

Then

$$ \begin{align*} T_J^{\text{red}}= \begin{cases} \{\,\mathfrak{t}(\mathbf{z})\,\, | \,\, \mathbf{z}\in\mathbf{F}^{\#J_0^{\text{co}}}\,\}\quad &\mbox{ if }\,\, J_0=J,\\ \{\,\mathfrak{t}(\mathbf{z})\,\, | \,\, \mathbf{z}\in\mathbf{F}^{\#J_0^{\text{co}}}\,\, \&\,\, \prod_{i\in J_0^{\text{co}}}z_i^{m_i}=1\,\}\quad &\mbox{ if }\,\, J_0\not=J, \end{cases} \end{align*} $$

where $\varphi ^{\vee }=\sum _{i=1}^rm_i\alpha _i^{\vee }$ ( $m_i\in \mathbb {Z}_{>0}$ ) is the expansion of $\varphi ^{\vee }\in \Phi _0^{\vee }$ in simple coroots.

6.2 The monic Y-eigenbasis of $\mathcal {P}^{(c)}_{\mathfrak {t}}$

For $c\in C^J$ and $\mathfrak {t}\in T_J$ , we have encountered so far two natural bases of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ , the basis $\{x^y\}_{y\in \mathcal {O}_c}$ of quasi-monomials and the basis $\{m_y^{J}(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ with

(6.3) $$ \begin{align} m_y^{J}(x;\mathfrak{t}):=\pi_{c,\mathfrak{t}}\big(\delta(T_{w_y^{-1}})\big)x^c\qquad (y\in\mathcal{O}_c). \end{align} $$

The latter basis corresponds to the basis $\{m_{w;\mathfrak {s}_J\mathfrak {t}}^J\}_{w\in W^J}$ of $\mathbb {M}_{\mathfrak {s}_J\mathfrak {t}}^J$ through the isomorphism $\phi _{c,\mathfrak {t}}$ from Theorem 4.5(2). The change of basis matrix between these two bases is triangular with respect to the partially ordered set $(\mathcal {O}_c,\leq )$ ,

(6.4) $$ \begin{align} m_y^{J}(x;\mathfrak{t})=k_{w_y}(c)x^{y}+\sum_{y^\prime<y}d_{y,y^\prime;\mathfrak{t}}^{J}x^{y^\prime}\qquad (y\in\mathcal{O}_c) \end{align} $$

( $d_{y,y^\prime ;\mathfrak {t}}^{J}\in \mathbf {F}$ ) by Proposition 5.29. Note that for $u,w\in W^J$ satisfying $u<_Bw$ , the coefficient

$$\begin{align*}d_{w,u;\mathfrak{t}}^J:=d_{wc,uc;\mathfrak{t}}^{J} \end{align*}$$

of $x^{uc}$ in the expansion (6.4) of $m_{wc}^{J}(x;\mathfrak {t})$ does not depend on $c\in C^J$ by Corollary 4.6. We now construct for $\mathfrak {t}\in T_J^\prime $ a third basis of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ consisting of simultaneous eigenfunctions of $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ), which is always triangular with respect to the basis of quasi-monomials.

We start with introducing some general terminology. For a left $\mathbb {H}$ -module M and $s\in T$ , write

$$\begin{align*}M[s]:=\{m\in M \,\,\, | \,\,\, p(Y)m=p(s^{-1})m\quad \forall\, p\in\mathcal{P} \}. \end{align*}$$

We call $s\in T$ a Y-weight of M if $M[s]\not =0$ . A nonzero vector in $M[s]$ is called a Y-weight vector of weight s. The set of all Y-weights of M is called the Y-spectrum of M and will be denoted by $\mathcal {S}(M)$ . Following Cherednik [Reference Cherednik8, §3.6], we say that M is Y-semisimple if

$$\begin{align*}M=\bigoplus_{s\in\mathcal{S}(M)}M[s]. \end{align*}$$

The Y-spectrum of a Y-semisimple $\mathbb {H}$ -module M is called simple if $\dim _{\mathbf {F}}(M[s])=1$ for all $s\in \mathcal {S}(M)$ . A basis of M consisting of Y-weight vectors is called a Y-eigenbasis of M.

Theorem 6.3. Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ .

  1. 1. For $y\in \mathcal {O}_c$ , there exists a unique simultaneous eigenfunction of the commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )\in \text {End}(\mathcal {P}^{(c)})$ ( $\mu \in Q^{\vee }$ ) of the form

    $$\begin{align*}E_y^J(x;\mathfrak{t})=x^y+\sum_{y^\prime<y}e_{y,y^\prime;\mathfrak{t}}^{J}x^{y^\prime}\qquad (e_{y,y^\prime;\mathfrak{t}}^{J}\in\mathbf{F}). \end{align*}$$
  2. 2. $E_y^J(x;\mathfrak {t})\in \mathcal {P}^{(c)}_{\mathfrak {t}}[\mathfrak {s}_y\mathfrak {t}_y]$ for $y\in \mathcal {O}_c$ .

  3. 3. $\{E_y^J(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ is a Y-eigenbasis of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ . In particular, $\mathcal {P}^{(c)}_{\mathfrak {t}}\simeq \mathbb {M}_{\mathfrak {s}_c\mathfrak {t}}^J$ is Y-semisimple, with simple Y-spectrum $\{\mathfrak {s}_y\mathfrak {t}_y\}_{y\in \mathcal {O}_c}$ .

  4. 4. For $u,w\in W^J$ such that $u<_Bw$ , the coefficient

    $$\begin{align*}e_{w,u;\mathfrak{t}}^J:=e_{wc,uc;\mathfrak{t}}^{J} \end{align*}$$
    of $x^{uc}$ in the monomial expansion of $E_{wc}^J(x;\mathfrak {t})$ does not depend on $c\in C^J$ .

Proof. Parts (1)&(2) are due to Proposition 5.28, part (3) is a direct consequence of (1),(2) and Theorem 4.5(2), and part (4) follows from (4.13).

Remark 6.4.

  1. 1. If $1_T\in T_{[1,r]}^\prime $ and $\mu \in \mathcal {O}_0=Q^{\vee }$ , then

    $$\begin{align*}E_\mu^{[1,r]}(x;1_T)\in\mathcal{P}^{(0)}_{1_T} \end{align*}$$
    is the monic non-symmetric Macdonald polynomial of degree $\mu $ (see, for example, [Reference Cherednik8, Reference Macdonald29]).
  2. 2. For $c\in C^J$ , $\mathfrak {t}\in T_J$ and $y\in \mathcal {O}_c$ , we have by (6.4) and Theorem 6.3(1),

    $$\begin{align*}E_y^J(x;\mathfrak{t})=k_{w_y}(c)^{-1}m_y^{J}(x;\mathfrak{t})+ \cdots \end{align*}$$
    with $\cdots $ meaning a linear combination of the $m_{y^\prime }^{J}(x;\mathfrak {t})$ with $y^\prime <y$ .

The following corollary is immediate.

Corollary 6.5. Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ .

  1. 1. The transition matrices between the three bases $\{x^y\}_{y\in \mathcal {O}_c}$ , $\{m_y^{J}(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ and $\{E_y^{J}(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ are triangular with respect to $\leq $ .

  2. 2. Identifying $(\mathcal {O}_c,\leq )$ with $(W^J,\leq _B)$ as partially ordered sets by the map

    $$\begin{align*}W^J\overset{\sim}{\longrightarrow}\mathcal{O}_c,\qquad w\mapsto wc, \end{align*}$$
    the transition matrices $A=(a_{u,w})_{u,w\in W^J}$ between the bases in (1) do not depend on $c\in C^J$ .

In the following lemma, we give two elementary properties of the monic quasi-polynomials $E_y^J(x;\mathfrak {t})$ .

Lemma 6.6. For $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ , we have

  1. 1. $E_{c}^J(x,\mathfrak {t})=x^{c}$ .

  2. 2. $E_{y+z}^J(x;\mathfrak {t})= x^zE_y^J(x;\mathfrak {t})$ when $y\in \mathcal {O}_c$ and $z\in E_{\text {co}}$ .

Proof. (1) By Theorem 4.5(2), we have $x^c\in \mathcal {P}^{(c)}_{\mathfrak {t}}[\mathfrak {s}_J\mathfrak {t}]$ . Now apply Theorem 6.3.

(2) This follows from the fact that $c+z\in C^J$ , $\mathcal {O}_{c+z}=\mathcal {O}_c+z$ and

$$\begin{align*}\pi_{c,\mathfrak{t}}(h)x^{y+z}=x^z\pi_{c,\mathfrak{t}}(h)x^{y} \end{align*}$$

for all $h\in \mathbb {H}$ .

6.3 Creation operators and irreducibility conditions

The Y-semisimple $\mathbb {H}$ -module $\mathcal {P}^{(c)}_{\mathfrak {t}}$ ( $c\in C^J$ , $\mathfrak {t}\in T_J^\prime $ ) has simple Y-spectrum $\{w(\mathfrak {s}_J\mathfrak {t})\}_{w\in W^J}$ by Theorem 6.3. In this subsection, we will use the action of Y-intertwiners $S_w^Y$ ( $w\in W^J$ ) to create the quasi-polynomial $E_{wc}^J(x;\mathfrak {t})\in \mathcal {P}^{(c)}_{\mathfrak {t}}[w(\mathfrak {s}_J\mathfrak {t})]$ from the quasi-monomial $x^c\in \mathcal {P}^{(c)}_{\mathfrak {t}}[\mathfrak {s}_J\mathfrak {t}]$ .

Lemma 6.7. Let M be a $\mathbb {H}$ -module, $s\in \mathcal {S}(M)$ , and $w\in W$ . Then $S_w^Y$ maps $M[s]$ to $M[ws]$ .

Proof. For a polynomial $p=\sum _{\mu \in Q^{\vee }}a_\mu x^\mu \in \mathcal {P}$ and $w\in W$ , write

$$\begin{align*}p_w^\iota:=\sum_{\mu\in Q^{\vee}}a_\mu w(x^{-\mu})=\sum_{\mu\in Q^{\vee}}a_\mu x^{-w\cdot\mu}\in\mathcal{P}. \end{align*}$$

Then we have $p(Y)S_w^Y =S_w^Yp_w^\iota (Y^{-1})$ by (2.59). For $m\in M[s]$ , we then have

$$ \begin{align*} p(Y)S_w^Ym=S_{w}^Yp^\iota_w(Y^{-1})m =p^\iota_w(s)S_w^Ym=p((ws)^{-1})S_w^Ym. \end{align*} $$

Hence, $S_w^Ym\in M[ws]$ .

Applying the lemma to $\mathcal {P}^{(c)}_{\mathfrak {t}}$ ( $c\in C^J$ , $\mathfrak {t}\in T_J$ ), we conclude that

(6.5) $$ \begin{align} \pi_{c,\mathfrak{t}}(S_{w}^Y)x^c\in\mathcal{P}_{\mathfrak{t}}^{(c)}[w(\mathfrak{s}_J\mathfrak{t})]\qquad (w\in W), \end{align} $$

since $x^c\in \mathcal {P}^{(c)}_{\mathfrak {t}}[\mathfrak {s}_J\mathfrak {t}]$ .

For $w\in W$ , set

(6.6) $$ \begin{align} d_{w}(x):=\prod_{a\in\Pi(w)}(x^{a^{\vee}}-1)\in\mathcal{P}. \end{align} $$

Proposition 6.8. Let $c\in C^J$ and $\mathfrak {t}\in T_J$ .

  1. 1. $\pi _{c,\mathfrak {t}}(S_w^Y)x^c=0$ if $w\in W\setminus W^J$ .

  2. 2. There exist $\epsilon _{w,u}\in \mathbf {F}$ ( $u,w\in W^J$ : $u<_Bw$ ) such that

    (6.7) $$ \begin{align} \pi_{c,\mathfrak{t}}(S_w^Y)x^c=d_{w}(\mathfrak{s}_J\mathfrak{t})m_{wc}^{J}(x;\mathfrak{t})+\sum_{u\in W^J: u<_Bw}\epsilon_{w,u}m_{uc}^{J}(x;\mathfrak{t}) \end{align} $$
    for all $w\in W^J$ .

Proof. For the duration of the proof, we write $t:=\mathfrak {s}_J\mathfrak {t}\in L_J$ .

(1) Let $j\in J$ . It suffices to show that $\pi _{c,\mathfrak {t}}(S_j^Y)x^c=0$ . Since $x^c\in \mathcal {P}^{(c)}_{\mathfrak {t}}[t]$ and $\pi _{c,\mathfrak {t}}(\delta (T_j))x^c=k_jx^c$ by Theorem 4.5, substitution of the explicit expression (2.56) for $S_j^Y$ gives

$$ \begin{align*} \pi_{c,\mathfrak{t}}(S_j^Y)x^c=\big(k_j(t^{\alpha_j^{\vee}}-1)+k_j-k_j^{-1}\big)x^c. \end{align*} $$

This vanishes since $t\in L_J$ .

(2) The proof is by induction to $\ell (w)$ . Fix $w\in W^{J}$ with $\ell (w)>0$ . Write $w=s_{j}u$ with $j\in [0,r]$ and $\ell (s_ju)=\ell (u)+1$ . Then $u\in W^{J}$ by Lemma 2.2, and $u<_Bw$ . Furthermore, $S_w^Y=S_j^YS_u^Y$ and

(6.8) $$ \begin{align} d_{w}(x)=\big(x^{(u^{-1}\alpha_j)^{\vee}}-1\big)d_{u}(x) \end{align} $$

by Lemma 2.1. By the induction hypothesis, the explicit expression (2.56) for $S_j^Y$ and (6.5), we then have for $c\in C^J$ and $\mathfrak {t}\in T_J$ ,

$$ \begin{align*} \begin{aligned} &\pi_{c,\mathfrak{t}}(S_w^Y)x^c= \big((t^{(u^{-1}\alpha_j)^{\vee}}-1)\pi_{c,\mathfrak{t}}(\delta(T_j))+k_j-k_j^{-1}\big)\pi_{c,\mathfrak{t}}(S_u^Y)x^c\\ &\,\,\,\,\,\,\,\in d_{w}(t)m_{wc}^{J}(x;\mathfrak{t})+\sum_{w^\prime\in W^{J}:\, w^\prime<_Bu} \big(t^{(u^{-1}\alpha_j)^{\vee}}-1\big)\epsilon_{u,w^\prime}\pi_{c,\mathfrak{t}}(\delta(T_j))m_{w^\prime c}^{J}(x;\mathfrak{t})+\mathcal{P}^{(c)}_{<w} \end{aligned} \end{align*} $$

for some $\epsilon _{u,w^\prime }\in \mathbf {F}$ , with

$$\begin{align*}\mathcal{P}^{(c)}_{<w}:=\bigoplus_{w^{\prime\prime}\in W^J:\, w^{\prime\prime}<_Bw}\mathbf{F}m_{w^{\prime\prime}c}^J(x;\mathfrak{t}). \end{align*}$$

It thus remains to show that $\pi _{c,\mathfrak {t}}(\delta (T_j))m_{w^\prime c}^{J}(x;\mathfrak {t})\in \mathcal {P}^{(c)}_{<w}$ when $w^\prime \in W^{J}$ and $w^\prime <_Bu$ . There are three cases to consider.

Case 1: If $\ell (s_jw^\prime )=\ell (w^\prime )+1$ and $s_jw^\prime \not \in W^{J}$ , then $s_jw^\prime =w^\prime s_{j^\prime }$ with $j^\prime \in J$ by Lemma 2.2, and hence, $\pi _{c,\mathfrak {t}}(\delta (T_j))m_{w^\prime c}^{J}(x;\mathfrak {t})=k_jm_{w^\prime c}^{J}(x;\mathfrak {t})\in \mathcal {P}^{(c)}_{<w}$ since $w^\prime <_Bu<_Bw$ .

Case 2: If $\ell (s_jw^\prime )=\ell (w^\prime )+1$ and $s_jw^\prime \in W^{J}$ , then $\pi _{c,\mathfrak {t}}(\delta (T_j))m_{w^\prime c}^{J}(x;\mathfrak {t})=m_{s_jw^\prime c}^{J}(x;\mathfrak {t})$ lies in $\mathcal {P}^{(c)}_{<w}$ since $s_jw^\prime <_Bs_ju=w$ by Lemma 5.13.

Case 3: If $\ell (s_jw^\prime )=\ell (w^\prime )-1$ , then $s_jw^\prime <_Bw^\prime <w$ and $s_jw^\prime \in W^{J}$ by Lemma 2.2; hence,

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_j))m_{w^\prime c}^{J}(x;\mathfrak{t})= m_{s_jw^\prime c}^{J}(x;\mathfrak{t})+(k_j-k_j^{-1})m_{w^\prime c}^{J}(x;\mathfrak{t}) \end{align*}$$

lies in $\mathcal {P}_{<w}^{(c)}$ .

The basis elements $m_{wc}^J(x;\mathfrak {t})$ ( $w\in W^J$ ) of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ can now be expanded in terms of $\pi _{c,\mathfrak {t}}(S_u^Y)x^c$ ( $u\in W^J$ ) when the leading coefficients $d_w(\mathfrak {s}_J\mathfrak {t})$ ( $w\in W^J$ ) in (6.7) are nonzero. This will be ensured by the following generic condition on $\mathfrak {t}\in T_J$ .

Definition 6.9. We say that $\mathfrak {t}\in T_J$ is J-regular if $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}\not =1$ for all $a\in \Phi \setminus \Phi _{J}$ .

We now have the following consequence of Proposition 6.8(2).

Corollary 6.10. Let $c\in C^J$ and $\mathfrak {t}\in T_J$ . Assume that $\mathfrak {t}$ is J-regular.

  1. 1. For $w\in W^J$ , we have

    $$\begin{align*}m_{wc}^{J}(x;\mathfrak{t})=\frac{1}{d_{w}(\mathfrak{s}_J\mathfrak{t})}\pi_{c,\mathfrak{t}}(S_w^Y)x^c+\sum_{u\in W^{J}:\, u<_B w}\epsilon^\prime_{w,u}\pi_{c,\mathfrak{t}}(S_u^Y)x^c \end{align*}$$
    for some $\epsilon _{w,u}^\prime \in \mathbf {F}$ .
  2. 2. $\{\pi _{c,\mathfrak {t}}(S_w^Y)x^c\,\, | \,\, w\in W^J\}$ is a Y-eigenbasis of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ .

Proof. For $w\in W^{J}$ , we have $\Pi (w)\subseteq \Phi ^+\setminus \Phi ^+_{J}$ by (2.14); hence, $d_{w}\in \mathcal {P}$ does not vanish at $\mathfrak {s}_J\mathfrak {t}$ for J-regular $\mathfrak {t}\in T_J$ . The corollary now follows from Lemma 3.14 and Proposition 6.8.

In Theorem 6.3(3), we described the Y-spectrum of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ when $\mathfrak {t}\in T_J^\prime $ , in which case the Y-spectrum is simple. When the condition $\mathfrak {t}\in T_J^\prime $ is replaced by the assumption that $\mathfrak {t}\in T_J$ is J-regular, we have the following result.

Corollary 6.11. Let $c\in C^J$ and suppose that $\mathfrak {t}\in T_J$ is J-regular. Then

  1. 1. $\mathcal {P}^{(c)}_{\mathfrak {t}}$ is Y-semisimple.

  2. 2. For $s\in T$ , we have

    $$\begin{align*}\mathrm{Dim}_{\mathbf{F}}\big(\mathcal{P}^{(c)}_{\mathfrak{t}}[s]\big)=\#\{w\in W^J\,\, | \,\, s=w(\mathfrak{s}_J\mathfrak{t})\}. \end{align*}$$

Remark 6.12. Suppose that $\mathfrak {t}\in T_J$ is J-regular. Then $\mathcal {P}^{(c)}_{\mathfrak {t}}$ ( $c\in C^J$ ) is Y-semisimple with finite dimensional Y-weight spaces, since q is not a root of unity. In the terminology of Cherednik [Reference Cherednik8, §3.6.1], $\mathcal {P}^{(c)}_{\mathfrak {t}}\simeq \mathbb {M}^{J}_{\mathfrak {s}_J\mathfrak {t}}$ is a Y-cyclic, Y-semisimple $\mathbb {H}$ -module in category $\mathcal {O}_Y$ .

In the following theorem, we relate $\pi _{c,\mathfrak {t}}(S_w^Y)x^c$ for $c\in C^J$ , $\mathfrak {t}\in T_J$ and $w\in W^J$ to the monic quasi-polynomial simultaneous eigenfunction $E_{wc}^J(x;\mathfrak {t})$ of the commuting operators $\pi _{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in Q^{\vee }$ ) from Theorem 6.3. This requires $\mathfrak {t}\in T_J^\prime $ , but $\mathfrak {t}$ does not have to be J-regular.

Theorem 6.13. Let $c\in C^J$ , $\mathfrak {t}\in T_J^\prime $ and $w\in W^J$ . Then

(6.9) $$ \begin{align} \pi_{c,\mathfrak{t}}(S_w^Y)x^c= d_w(\mathfrak{s}_J\mathfrak{t})k_w(c)E_{wc}^J(x; \mathfrak{t}) \end{align} $$

with $k_w(y)\in \mathbf {F}$ and $d_w\in \mathcal {P}$ defined by (5.7) and (6.6), respectively.

Proof. Let $w\in W^J$ . We have $\pi _{c,\mathfrak {t}}(S_w^Y)x^c=\epsilon _wE_{wc}^J(x;\mathfrak {t})$ for some $\epsilon _w\in \mathbf {F}$ because the Y-weight space $\mathcal {P}^{(c)}_{\mathfrak {t}}[w(\mathfrak {s}_J\mathfrak {t})]$ is one-dimensional. By (6.7) and (6.4), we have

$$\begin{align*}\pi_{c,\mathfrak{t}}(S_w^Y)x^c= d_w(\mathfrak{s}_J\mathfrak{t})k_w(c)x^{wc}+ \text{l.o.t.}; \end{align*}$$

hence, Theorem 6.3(1) implies that $\epsilon _w=d_w(\mathfrak {s}_J\mathfrak {t})k_w(c)$ . This concludes the proof.

Corollary 6.14. With the assumptions of Theorem 6.13, we have

  1. 1. $\pi _{c,\mathfrak {t}}(S_w^Y)x^c=0 \ \Leftrightarrow \ d_w(\mathfrak {s}_J\mathfrak {t})=0$ for $w\in W^J$ .

  2. 2. If $\mathfrak {t}$ is J-regular, then $\{\pi _{c,\mathfrak {t}}(S_w^Y)x^c\}_{w\in W^J}$ is a Y-eigenbasis of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ .

Proof. (1) is immediate from (6.9). For (2), it suffices to recall that the J-regularity of $\mathfrak {t}$ ensures that $d_w(\mathfrak {s}_J\mathfrak {t})\not =0$ for all $w\in W^J$ .

Theorem 6.15. Let $c\in C^J$ . Suppose that $\mathfrak {t}\in T_J^\prime $ is J-regular and that $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}\not =k_a^{-2}$ for all $a\in \Phi \setminus \Phi _J$ . Then $\mathfrak {s}_J\mathfrak {t}\in L^J$ , and $\mathcal {P}^{(c)}_{\mathfrak {t}}$ is irreducible.

Proof. The first statement follows immediately from the definition of $L^J\subseteq L_J$ (see Remark 3.12(1)).

Suppose that $M\subseteq \mathcal {P}^{(c)}_{\mathfrak {t}}$ is a nonzero subrepresentation. By the assumptions on $\mathfrak {t}$ , $\mathcal {P}^{(c)}_{\mathfrak {t}}$ is Y-semisimple with simple Y-spectrum, and $\{\pi _{c,\mathfrak {t}}(S_w^Y)x^c\}_{w\in W^{J}}$ is a Y-eigenbasis of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ . Hence, $\pi _{c,\mathfrak {t}}(S_w^Y)x^c\in M$ for some $w\in W^{J}$ . By (2.59), we then have

(6.10) $$ \begin{align} \pi_{c,\mathfrak{t}}(S_{w^{-1}}^Y)\pi_{c,\mathfrak{t}}(S_w^Y)x^c=\text{n}_w(\mathfrak{s}_J\mathfrak{t})x^c\in M, \end{align} $$

with $\text {n}_w\in \mathcal {P}$ defined by

(6.11) $$ \begin{align} \text{n}_w(x):=\prod_{a\in\Pi(w)}(k_a^{-1}-k_ax^{a^{\vee}})(k_a^{-1}-k_ax^{-a^{\vee}}). \end{align} $$

Note that $\Pi (w)\subseteq \Phi ^+\setminus \Phi _{J}^+$ by (2.14); hence, $n_w(\mathfrak {s}_J\mathfrak {t})\not =0$ by the assumption that $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}\not =k_a^{-2}$ for all $a\in \Phi \setminus \Phi _J$ . We conclude that $x^c\in M$ ; hence, $M=\mathcal {P}^{(c)}_{\mathfrak {t}}$ .

6.4 Closure relations

For $c\in C^J$ , $\mathfrak {t}\in T_J$ and $w\in W^J$ , the coefficients $e_{w,u;\mathfrak {t}}^J$ of $x^{uc}$ ( $u\in W^J$ ) in the expansion of $E_{wc}^J(x;\mathfrak {t})$ in quasi-monomials do not dependent on $c\in C^J$ (see Theorem 6.3). In this subsection, we express the coefficients $e_{w^\prime ,u^\prime ;\mathfrak {t}^\prime }^{J^\prime }$ in terms of $e_{w,u;\mathfrak {t}^\prime }^J$ when $J\subseteq J^\prime $ (i.e., when $C^{J^\prime }\subseteq \overline {C^J}$ ).

Recall the map $\text {pr}_{c,c^\prime }^{\mathfrak {t}^\prime }$ , defined in Corollary 4.7.

Proposition 6.16. Let $J\subseteq J^\prime $ . Fix a $J^\prime $ -regular $\mathfrak {t}^\prime \in T_{J^\prime }^\prime $ such that $\mathfrak {s}_J^{-1}\mathfrak {s}_{J^\prime }\mathfrak {t}^\prime \in T_J$ lies in $T_J^\prime $ and is J-regular. Let $c\in C^J$ and $c^\prime \in C^{J^\prime }$ . Then

(6.12) $$ \begin{align} \text{pr}_{c,c^\prime}^{\mathfrak{t}^\prime} \Big(k_w(c)E_{wc}^J(x;\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime)\Big)= \begin{cases} 0\quad &\mbox{ if }\, w\in W^J\setminus W^{J^\prime},\\ k_w(c^\prime)E_{wc^\prime}^{J^\prime}(x;\mathfrak{t}^\prime)\quad &\mbox{ if }\, w\in W^{J^\prime}. \end{cases} \end{align} $$

Proof. Let $w\in W^J$ . By (6.9), the left-hand side of (6.12) equals

$$\begin{align*}\frac{1}{d_w(\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime)}\text{pr}_{c,c^\prime}^{\mathfrak{t}^\prime}\big(\pi_{c,\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime}(S_w^Y)x^c \big). \end{align*}$$

By Proposition 6.8(1) and (6.9), it then suffices to show that

$$\begin{align*}\text{pr}_{c,c^\prime}^{\mathfrak{t}^\prime}\big(\pi_{c,\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime}(S_w^Y)x^c\big)= \pi_{c^\prime,\mathfrak{t}^\prime}(S_w^Y)x^{c^\prime}. \end{align*}$$

This follows from Corollary 4.7.

Keep the assumptions of Proposition 6.16. For $w\in W^J$ and $u^\prime \in W^{J^\prime }$ , consider the expression

(6.13) $$ \begin{align} e_{w,u^\prime;\mathfrak{t}^\prime}^{J,J^\prime}:=\sum_{u\in W^J\cap u^\prime W_{J^\prime}:\,u\leq_Bw} \frac{\kappa_{Du}(c^\prime)}{\kappa_{Du}(c)}e_{w,u;\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime}^J \end{align} $$

involving the coefficients of $E_{wc}^J(x;\mathfrak {s}_J^{-1}\mathfrak {s}_{J^\prime }\mathfrak {t}^\prime )$ in its expansion in quasi-monomials. The following consequence of Proposition 6.16 shows that (6.13) provides the coefficients of $E_{wc}^{J^\prime }(x;\mathfrak {t}^\prime )$ in its expansion in quasi-monomials when $w\in W^{J^\prime }$ .

Corollary 6.17. Keep the assumptions of Proposition 6.16. Then

  1. 1. If $w\in W^J\setminus W^{J^\prime }$ , then $e_{w,u^\prime ;\mathfrak {t}^\prime }^{J,J^\prime }=0$ for all $u^\prime \in W^{J^\prime }$ .

  2. 2. If $w\in W^{J^\prime }$ and $u^\prime \not \leq _Bw$ , then $e_{w,u^\prime ;\mathfrak {t}^\prime }^{J,J^\prime }=0$ .

  3. 3. If $w\in W^{J^\prime }$ , then $e_{w,w;\mathfrak {t}^\prime }^{J,J^\prime }=k_w(c^\prime )/k_w(c)$ .

  4. 4. If $u^\prime ,w\in W^{J^\prime }$ and $u^\prime \leq _Bw$ , then

    $$\begin{align*}e_{w,u^\prime;\mathfrak{t}^\prime}^{J^\prime}=\frac{k_w(c)}{k_w(c^\prime)}e_{w,u^\prime;\mathfrak{t}^\prime}^{J,J^\prime}. \end{align*}$$

Proof. By Corollary 4.7, we have

$$ \begin{align*} \text{pr}_{c,c^\prime}^t\big(E_{wc}^J(x;\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime)\big)=\sum_{u^\prime\in W^{J^\prime}}e_{w,u^\prime;\mathfrak{t}^\prime}^{J,J^\prime}x^{u^\prime c^\prime}\qquad (w\in W^J). \end{align*} $$

The result now follows from Proposition 6.16.

Consider the special case $J^\prime =[1,r]$ , $c^\prime =0$ and $\mathfrak {t}^\prime =1_T$ ; cf. Remark 4.8. In this case, $W^{[1,r]}=\{w_\mu \,\, | \,\, \mu \in Q^{\vee }\}$ . Let $J\subseteq J^\prime $ and $c\in C^J$ . Then we conclude from Proposition 6.16 that

$$\begin{align*}\text{pr}_{c,0}^{1_T}(E_{w_\lambda c}^J(x;\mathfrak{s}_J^{-1}t_{\text{sph}}))=\frac{k_{w_\lambda}(0)}{k_{w_\lambda}(c)}E_\lambda^{[1,r]} (x;1_T)\qquad (\lambda\in Q^{\vee}). \end{align*}$$

Corollary 6.17(3) then yields the expression

$$\begin{align*}e_{w_\lambda,w_\mu;1_T}^{[1,r]}=\frac{k_{w_\lambda}(c)}{k_{w_\lambda}(0)}\sum_{u\in W^J\cap w_\mu W_0:\, u\leq_Bw_\lambda}\frac{\kappa_{Du}(0)}{\kappa_{Du}(c)} e_{w_\lambda,u;\mathfrak{s}_J^{-1}t_{\text{sph}}}^J \end{align*}$$

for the coefficient of $x^\mu $ in the expansion of the nonsymmetric Macdonald polynomial $E_\lambda ^{[1,r]}(x;1_T)$ in monomials.

6.5 The normalised Y-eigenbasis and pseudo-duality

The normalised version of the quasi-polynomial $E_y^{J}(x;\mathfrak {t})$ that we will introduce in this subsection, requires the following additional constraint on $\mathfrak {t}\in T_J^\prime $ .

Definition 6.18. We say that $\mathfrak {t}\in T_J$ is J-generic if $\mathfrak {t}$ is J-regular and $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}\not =k_a^{-2}$ for all $a\in \Phi ^+\setminus \Phi _J^+$ .

Note that $\mathfrak {s}_J\mathfrak {t}\in L^J$ if $\mathfrak {t}\in T_J$ is J-generic. Furthermore, for $c\in C^J$ and J-generic $\mathfrak {t}\in T_J^\prime $ , the irreducibility criterion for the quasi-polynomial representation $\mathcal {P}^{(c)}_{\mathfrak {t}}$ (see Theorem 6.15) requires the additional constraints $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}\not =k_a^{-2}$ for all $a\in \Phi ^-\setminus \Phi _J^-$ .

For $w\in W$ , define $r_{w}\in \mathcal {P}$ by

(6.14) $$ \begin{align} r_{w}(x):=\prod_{a\in\Pi(w)}(k_a x^{a^{\vee}}-k_a^{-1}). \end{align} $$

Note that $\widetilde {S}^Y_w=S_w^Yr_w(Y^{-1})^{-1}$ in $\mathbb {H}^{Y-\text {loc}}$ by (2.60). For $w\in W^J$ and J-generic $\mathfrak {t}\in T_J$ , we have $r_{w}(\mathfrak {s}_J\mathfrak {t})\not =0$ by (2.14), and hence, we may define for $c\in C^J$ ,

(6.15) $$ \begin{align} P_{wc}^{J}(x;\mathfrak{t}):= \frac{\pi_{c,\mathfrak{t}}(S_w^Y)x^c}{r_w(\mathfrak{s}_J\mathfrak{t})}\in\mathcal{P}^{(c)}_{\mathfrak{t}}[w(\mathfrak{s}_J\mathfrak{t})]\qquad\quad (w\in W^J). \end{align} $$

Note that $\{P_y^J(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ is a Y-eigenbasis of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ by Corollary 6.10.

Remark 6.19. Adding to the assumptions of Proposition 6.16 the assumption that $\mathfrak {t}^\prime $ is $J^\prime $ -generic, we have

$$\begin{align*}\text{pr}_{c,c^\prime}^{\mathfrak{t}^\prime}\big(P_{wc}^J(x;\mathfrak{s}_J^{-1}\mathfrak{s}_{J^\prime}\mathfrak{t}^\prime)\big)= P_{wc^\prime}^{J^\prime}(x;\mathfrak{t}^\prime)\qquad (w\in W^{J^\prime}). \end{align*}$$

This follows from (6.15) and the fact that $\text {pr}_{c,c^\prime }^{\mathfrak {t}^\prime }(x^c)=x^{c^\prime }$ .

The precise relation of the normalised quasi-polynomials (6.15) with the monic Y-eigenbasis $\{E_y^J(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ is as follows.

Proposition 6.20. Let $c\in C^J$ . For J-generic $\mathfrak {t}\in T_J^\prime $ and $y\in \mathcal {O}_c$ , set

(6.16) $$ \begin{align} t_y:=w_y(\mathfrak{s}_J\mathfrak{t})=\mathfrak{s}_y\mathfrak{t}_y\in T. \end{align} $$

Then

$$\begin{align*}P_y^J(x;\mathfrak{t})=\frac{k(y)}{k(c)}\prod_{a\in\Phi^+: a(y)<0} \left(\frac{t_y^{-a^{\vee}}-1}{k_at_y^{-a^{\vee}}-k_a^{-1}}\right) E_y^J(x;\mathfrak{t})\qquad\quad (y\in\mathcal{O}_c). \end{align*}$$

Proof. By Theorem 6.13, we have

(6.17) $$ \begin{align} P_{wc}^J(x;\mathfrak{t})=\frac{d_w(\mathfrak{s}_J\mathfrak{t})k_w(c)}{r_w(\mathfrak{s}_J\mathfrak{t})}E_{wc}^J(x;\mathfrak{t})\qquad\quad (w\in W^J). \end{align} $$

Now substitute $w=w_y\in W^J$ ( $y\in \mathcal {O}_c$ ). Then $k_{w_y}(c)=k(y)/k(c)$ by (5.7), and

$$\begin{align*}d_{w_y}(\mathfrak{s}_J\mathfrak{t})= \prod_{a\in\Pi(w_y^{-1})}\big((\mathfrak{s}_J\mathfrak{t})^{-(w_y^{-1}a)^{\vee}}-1\big)= \prod_{a\in\Phi^+: a(y)<0}\big(t_y^{-a^{\vee}}-1\big) \end{align*}$$

by (6.6), $\Pi (w_y)=-w_y^{-1}\Pi (w_y^{-1})$ , (2.31) and Corollary 5.22. In a similar manner, one verifies that

$$\begin{align*}r_{w_y}(\mathfrak{s}_J\mathfrak{t})=\prod_{a\in\Phi^+: a(y)<0}\big(k_at_y^{-a^{\vee}}-k_a^{-1}\big). \end{align*}$$

Substituting these expressions in (6.17) gives the desired result.

Remark 6.21. Suppose that $1_T\in T_{[1,r]}^\prime $ and that $1_T$ is $[1,r]$ -generic (this holds true for suitably generic values of $k_a\in \mathbf {F}$ ; cf. Subsection 6.1). Then

$$\begin{align*}P_\mu^{[1,r]}(x;1_T)\in\mathcal{P}\qquad (\mu\in \mathcal{O}_0=Q^{\vee}) \end{align*}$$

are the normalised nonsymmetric Macdonald polynomials; that is, $P_\mu ^{[1,r]}(x;1_T)$ is the normalisation of the monic nonsymmetric Macdonald polynomials $E_\mu ^{[1,r]}(x;1_T)$ such that

$$\begin{align*}P_\mu^{[1,r]}(t_{\text{sph}};1_T)=1. \end{align*}$$

This follows from Proposition 6.20 and the evaluation formula for the nonsymmetric monic Macdonald polynomials,

$$\begin{align*}E_\mu^{[1,r]}(t_{\text{sph}};1_T)=\frac{k(0)}{k(\mu)}\prod_{a\in\Phi^+: a(\mu)<0} \left(\frac{k_a\mathfrak{s}_\mu^{-a^{\vee}}-k_a^{-1}}{\mathfrak{s}_\mu^{-a^{\vee}}-1}\right) \qquad (\mu\in Q^{\vee}); \end{align*}$$

see, for example, [Reference Macdonald29, (5.2.14)] and [Reference Cherednik6, §5].

We will now describe the action of $\delta (T_j)$ on $\mathcal {P}_{\mathfrak {t}}^{(c)}$ in terms of a discrete Demazure-Lusztig operator acting on $\mathcal {O}_c$ (see Theorem 6.26). We first need to analyse the action of the affine Weyl group W on $L_J$ in more detail.

For $w\in W$ , set

$$\begin{align*}\rho_w:=\prod_{a\in \Pi(w^{-1})}k_a^{2Da}\in T. \end{align*}$$

Lemma 6.22. If $w\in W_J$ and $t\in L_J$ , then $wt=\rho _wt$ .

Proof. This is correct for $w=s_j$ ( $j\in J$ ) since $\Pi (s_j)=\{\alpha _j\}$ and $t\in L_J$ (see (2.37)). We now apply induction to $\ell (w)$ . If $w=us_j$ with $u \in W_{J}$ , $j\in J$ and $\ell (u s_j)=\ell (u)+1$ , then

(6.18) $$ \begin{align} \rho_w=k_j^{2D(u\alpha_j)}\rho_{u} \end{align} $$

by Lemma 2.1, and

$$\begin{align*}wt=u(k_j^{2D\alpha_j}t) =(Du)(k_j^{2D\alpha_j})u t =k_j^{2D(u\alpha_j)}ut. \end{align*}$$

Applying the induction hypothesis to $ut$ and using (6.18), we obtain $wt=\rho _wt$ .

Suppose that $w\in W^J$ and $j\in [0,r]$ satisfy $s_jw\not \in wW_J$ . Then $s_jw\in W^J$ by Lemma 2.2; hence, both $wt$ and $s_jwt$ are Y-weights of $\mathbb {M}^J_t$ when $t\in \mathfrak {s}_JT_J^\prime \subseteq L_J$ .

We now analyse what happens with the Y-weights of $\mathbb {M}_t^J$ when $w\in W^J$ and $j\in [0,r]$ satisfy $s_jw\in wW_{J}$ .

Lemma 6.23. Let $t\in L_J$ , $w\in W^J$ and $j\in [0,r]$ such that $s_jw\in wW_{J}$ . Then

$$\begin{align*}(wt)^{\alpha_j^{\vee}}=k_j^{-2}; \end{align*}$$

that is, $j\in J(wt)$ (see (3.15)). Furthermore, $s_jwt=k_j^{2D\alpha _j}wt$ .

Proof. By Lemma 2.2, the assumption $s_jw\in wW_J$ implies that $s_jw=ws_{j^\prime }$ for some $j^\prime \in J$ , and $\ell (s_jw)=\ell (w)+1=\ell (ws_{j^\prime })$ . Then $w\alpha _{j^\prime }=\alpha _j$ , and hence,

$$\begin{align*}(wt)^{\alpha_j^{\vee}}=t^{(w^{-1}\alpha_j)^{\vee}}=t^{\alpha_{j^\prime}^{\vee}}=k_{j^\prime}^{-2}=k_j^{-2}. \end{align*}$$

The second statement follows from (2.37).

Corollary 6.24. Let $c\in C^J$ and $\mathfrak {t}\in T_J$ . Suppose that $w\in W^J$ and $j\in [0,r]$ such that $s_jw\in wW_J$ . Then

(6.19) $$ \begin{align} (\pi_{c,\mathfrak{t}}(\delta(T_j))-k_j)\pi_{c,\mathfrak{t}}(S_w^Y)x^c=0 \end{align} $$

if $k_j^2\not =1$ .

Proof. For the duration of the proof, we write $t:=\mathfrak {s}_J\mathfrak {t}\in L_J$ .

Let $w\in W^J$ and $j\in [0,r]$ such that $s_jw\in wW_J$ . Then $s_jw=ws_{j^\prime }$ for some $j^\prime \in J$ , and $\ell (s_jw)=\ell (w)+1=\ell (ws_{j^\prime })$ (see the proof of Lemma 6.23). Hence,

$$\begin{align*}\pi_{c,\mathfrak{t}}(S_j^Y)\pi_{c,\mathfrak{t}}(S_w^Y)x^c=\pi_{c,\mathfrak{t}}(S_{s_jw}^Y)x^c=\pi_{c,\mathfrak{t}}(S_w^YS_{j^\prime}^Y)x^c=0 \end{align*}$$

by (the proof of) Proposition 6.8(1). We have $\pi _{c,\mathfrak {t}}(S_w^Y)x^c\in \mathcal {P}_{\mathfrak {t}}^{(c)}[wt]$ , and $j\in J(wt)$ by Lemma 6.23; hence,

$$\begin{align*}\pi_{c,\mathfrak{t}}\big((Y^{-1})^{\alpha_j^{\vee}}\big)\pi_{c,\mathfrak{t}}(S_w^Y)x^c=k_j^{-2}\pi_{c,\mathfrak{t}}(S_w^Y)x^c. \end{align*}$$

Using the explicit expression (2.56) for $S_j^Y$ , we conclude that

$$\begin{align*}0=\pi_{c,\mathfrak{t}}(S_j^Y)\pi_{c,\mathfrak{t}}(S_w^Y)x^c=((k_j^{-2}-1)\pi_{c,\mathfrak{t}}(\delta(T_j))+k_j-k_j^{-1})\pi_{c,\mathfrak{t}}(S_w^Y)x^c. \end{align*}$$

The result then follows from the assumption that $k_j^2\not =1$ .

Remark 6.25. If $\mathfrak {t}\in T_J^\prime $ and $\mathfrak {t}$ is J-regular, then the condition $k_j^2\not =1$ may be omitted in Corollary 6.24. Indeed, assume that $\mathfrak {t}\in T_J^\prime $ is J-regular and that $k_j^2=1$ , and fix $w\in W^J$ and $j\in [0,r]$ with $s_jw\in wW_J$ . Then we have $s_jwt=wt$ by Lemma 6.23, where $t=\mathfrak {s}_J\mathfrak {t}$ . The cross relations (2.40) then show that $\pi _{c,\mathfrak {t}}(\delta (T_j))\pi _{c,\mathfrak {t}}(S_w^Y)x^c$ lies in $\mathcal {P}_{\mathfrak {t}}^{(c)}[wt]$ . Since $\mathfrak {t}\in T_J^\prime $ , it follows that $\pi _{c,\mathfrak {t}}(\delta (T_j))\pi _{c,\mathfrak {t}}(S_w^Y)x^c$ is a constant multiple of $\pi _{c,\mathfrak {t}}(S_w^Y)x^c$ . The constant multiple can be computed by expanding $\pi _{c,\mathfrak {t}}(S_w^Y)x^c$ in terms of the basis $\{m_{uc}^J(x;\mathfrak {t})\}_{u\in W^J}$ . Since $\mathfrak {t}$ is J-regular, $d_w(t)m_{wc}^J(x;\mathfrak {t})$ is the nonzero leading term in this expansion (see Proposition 6.8(2)). By the proof of Proposition 6.8, the leading term of $\pi _{c,\mathfrak {t}}(\delta (T_j))\pi _{c,\mathfrak {t}}(S_w^Y)x^c$ in its expansion along $\{m_{uc}^J(x;\mathfrak {t})\}_{u\in W^J}$ will then be $d_w(t)\pi _{c,\mathfrak {t}}(\delta (T_j))m_w^J(x;\mathfrak {t})=k_jd_w(t)m_w^J(x;\mathfrak {t})$ , where the equality follows from Lemma 3.9. We conclude that (6.19) is valid.

In the following theorem, we show that the action of $\delta (T_j)$ ( $0\leq j\leq r$ ) on the normalised Y-eigenbasis $\{P_y^J(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ is described by discrete Demazure-Lusztig operators acting on $y\in \mathcal {O}_c$ . We call this the pseudo-duality property of the normalised quasi-polynomials $P_y^J(x;\mathfrak {t})$ ( $y\in \mathcal {O}_c$ ).

Theorem 6.26. Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ . Suppose that $\mathfrak {t}$ is J-generic. Then

(6.20) $$ \begin{align} \pi_{c,\mathfrak{t}}(\delta(T_j))P_{y}^J(x;\mathfrak{t})=k_jP_{y}^J(x;\mathfrak{t})+ \left(\frac{k_j^{-1}-k_jt_y^{\alpha_j^{\vee}}}{1-t_y^{\alpha_j^{\vee}}}\right) \big(P_{s_jy}^J(x;\mathfrak{t})-P_{y}^J(x;\mathfrak{t})\big) \end{align} $$

for $0\leq j\leq r$ and $y\in \mathcal {O}_c$ . Here, $t_y\in T$ is defined by (6.16) and the right-hand side of (6.20) should always be read as $k_jP_y^J(x;\mathfrak {t})$ when $s_jy=y$

Proof. Let $y\in \mathcal {O}_c$ and $0\leq j\leq r$ .

The case that $s_jy=y$ follows from Remark 6.25. Note that in this case, we have $s_jw_y\in w_yW_J$ , so Lemma 6.23 gives

$$\begin{align*}t_{y}^{\alpha_j^{\vee}}=(\mathfrak{s}_J\mathfrak{t})^{(w_y^{-1}\alpha_j)^{\vee}}=k_j^{-2}; \end{align*}$$

hence, the right-hand side of (6.20) is well defined when $k_j^2\not =1$ and it simplifies to $k_jP_y^J(x;\mathfrak {t})$ .

In the remainder of the proof, we assume that $s_jy\not =y$ . This is equivalent to $s_jw_y=w_{s_jy}$ by Lemma 2.2 and Corollary 5.23. Then

$$\begin{align*}w_y^{-1}\alpha_j\in\Pi(w_y)\cup\Pi(s_jw_y)\subseteq\Phi\setminus\Phi_J \end{align*}$$

by Lemma 2.1 and (2.14), and

$$\begin{align*}t_{y}^{\alpha_j^{\vee}}=(\mathfrak{s}_J\mathfrak{t})^{(w_y^{-1}\alpha_j)^{\vee}}\not=1 \end{align*}$$

since $\mathfrak {t}$ is J-generic, and hence in particular J-regular. So the right-hand side of (6.20) is well defined, and formula (6.20) is equivalent to

(6.21) $$ \begin{align} \pi_{c,\mathfrak{t}}(\delta(T_j))P_{y}^J(x;\mathfrak{t})=\left(\frac{k_j-k_j^{-1}}{1-t_y^{\alpha_j^{\vee}}}\right)P_y^J(x;\mathfrak{t})+ \left(\frac{k_j^{-1}-k_jt_y^{\alpha_j^{\vee}}}{1-t_y^{\alpha_j^{\vee}}}\right)P_{s_jy}^J(x;\mathfrak{t}). \end{align} $$

To prove (6.21), we consider two cases.

Case 1: $\ell (w_{s_jy})=\ell (w_y)+1$ .

Consider the expansion

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_j))\pi_{c,\mathfrak{t}}(S_{w_y}^Y)x^c =\sum_{y^\prime\in\mathcal{O}_c}C_{y^\prime} \pi_{c,\mathfrak{t}}(S_{w_{y^\prime}}^Y)x^c \qquad (C_{y^\prime}\in\mathbf{F}) \end{align*}$$

in the Y-eigenbasis $\{\pi _{c,\mathfrak {t}}(S_{y^\prime }^Y)x^c\}_{y^{\prime }\in \mathcal {O}_c}$ of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ . Since $\ell (w_{s_jy})=\ell (w_y)+1$ , we then have

(6.22) $$ \begin{align} \begin{aligned} \pi_{c,\mathfrak{t}}&(S_{w_{s_jy}}^Y)x^c=\pi_{c,\mathfrak{t}}(S_j^YS_{w_y}^Y)x^c\\ =&\big((t_y^{\alpha_j^{\vee}}-1)\pi_{c,\mathfrak{t}}(\delta(T_j))+k_j-k_j^{-1}\big)\pi_{c,\mathfrak{t}}(S_{w_y}^Y)x^c\\ =&\big((t_y^{\alpha_j^{\vee}}-1)C_y+k_j-k_j^{-1}\big)\pi_{c,\mathfrak{t}}(S_{w_y}^Y)x^c +\sum_{y^\prime\in\mathcal{O}_c\setminus\{y\}}(t_y^{\alpha_j^{\vee}}-1) C_{y^\prime}\pi_{c,\mathfrak{t}}(S_{w_{y^\prime}}^Y)x^c. \end{aligned} \end{align} $$

The assumptions on $\mathfrak {t}$ ensure that the $t_y$ ( $y\in \mathcal {O}_c$ ) are pairwise different Y-weights of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ and that $\mathcal {P}^{(c)}_{\mathfrak {t}}[t_y]$ is spanned by $\pi _{c,\mathfrak {t}}(S_{w_y}^Y)x^c$ for $y\in \mathcal {O}_c$ . Comparing the different Y-weight contributions on both sides of (6.22) thus gives

$$ \begin{align*} \begin{aligned} C_{y^\prime}&=0\quad \text{ unless }\, y^\prime\in \{y,s_jy\},\\ C_y&=\frac{k_j-k_j^{-1}}{1-t_y^{\alpha_j^{\vee}}},\qquad C_{s_jy}=\frac{1}{t_y^{\alpha_j^{\vee}}-1}. \end{aligned} \end{align*} $$

We conclude that

(6.23) $$ \begin{align} \pi_{c,\mathfrak{t}}(\delta(T_j))\pi_{c,\mathfrak{t}}(S_{w_y}^Y)x^c =\left(\frac{k_j-k_j^{-1}}{1-t_y^{\alpha_j^{\vee}}}\right)\pi_{c,\mathfrak{t}}(S_{w_y}^Y)x^c+ \left(\frac{1}{t_y^{\alpha_j^{\vee}}-1}\right)\pi_{c,\mathfrak{t}}(S_{w_{s_jy}}^Y)x^c. \end{align} $$

Then (6.15) and the identity $r_{w_{s_jy}}(\mathfrak {s}_J\mathfrak {t})=(k_jt_y^{\alpha _j^{\vee }}-k_j^{-1})r_{w_y}(\mathfrak {s}_J\mathfrak {t})$ give (6.21).

Case 2: $\ell (w_{s_jy})=\ell (w_y)-1$ .

We can now apply the first case to $s_jy$ , yielding

(6.24) $$ \begin{align} \pi_{c,\mathfrak{t}}(\delta(T_j))P_{s_jy}^J(x;\mathfrak{t})= \left(\frac{k_j-k_j^{-1}}{1-t_{s_jy}^{\alpha_j^{\vee}}}\right)P_{s_jy}^{J}(x;\mathfrak{t}) +\left(\frac{k_j^{-1}-k_jt_{s_jy}^{\alpha_j^{\vee}}}{1-t_{s_jy}^{\alpha_j^{\vee}}}\right)P_{y}^J(x;\mathfrak{t}). \end{align} $$

Now act on both sides with $\delta (T_j)$ , apply the Hecke relation $\delta (T_j)^2=(k_j-k_j^{-1})\delta (T_j)+1$ to the left-hand side, and substitute (6.24) for the resulting $\pi _{c,\mathfrak {t}}(\delta (T_j))P_{s_jy}^J(x;\mathfrak {t})$ ’s in the equality. A straightforward computation, avoiding rescaling of the equation, then shows that

$$ \begin{align*} \begin{aligned} &\left(\frac{k_j^{-1}-k_jt_{s_jy}^{\alpha_j^{\vee}}}{1-t_{s_jy}^{\alpha_j^{\vee}}}\right)\pi_{c,\mathfrak{t}}(\delta(T_j))P_y^J(x;\mathfrak{t})\\ &\qquad=\frac{(k_j^{-1}-k_jt_{s_jy}^{\alpha_j^{\vee}})(k_j-k_j^{-1})}{(1-t_{s_jy}^{\alpha_j^{\vee}}) (1-t_{s_jy}^{-\alpha_j^{\vee}})}P_y^J(x;\mathfrak{t})+ \frac{(k_j^{-1}-k_jt_{s_jy}^{\alpha_j^{\vee}})(k_j^{-1}-k_jt_{s_jy}^{-\alpha_j^{\vee}})} {(1-t_{s_jy}^{\alpha_j^{\vee}})(1-t_{s_jy}^{-\alpha_j^{\vee}})}P_{s_jy}^J(x;\mathfrak{t}). \end{aligned} \end{align*} $$

Now dividing out the prefactor of $\pi _{c,\mathfrak {t}}(\delta (T_j))P_y^J(x;\mathfrak {t})$ (which is nonzero since $\mathfrak {t}$ is J-generic) and using that $t_{s_jy}^{\alpha _j^{\vee }}=t_y^{-\alpha _j^{\vee }}$ , we obtain the desired formula (6.21).

Remark 6.27.

  1. 1. In case of Cherednik’s basic representation $\mathcal {P}^{(0)}_{1_T}$ , Theorem 6.26 is an important intermediate step in proving the well-known duality formula

    $$\begin{align*}P_\mu^{[1,r]}(\mathfrak{s}_\nu;1_T)=P_\nu^{[1,r]}(\mathfrak{s}_\mu;1_T)\qquad (\mu,\nu\in Q^{\vee}) \end{align*}$$
    for the normalised nonsymmetric Macdonald polynomials; see [Reference Cherednik6, Thm 5.1]. An analogue of the duality formula for the quasi-polynomial generalisations $P_y^J(x;\mathfrak {t})$ of the normalised nonsymmetric Macdonald polynomials is not known.
  2. 2. The quasi-duality property (6.20), together with the eigenvalue equations

    $$\begin{align*}\pi_{c,\mathfrak{t}}(Y^\mu)P_y^J(x;\mathfrak{t})=t_y^{-\mu}P_y^J(x;\mathfrak{t})\qquad (y\in\mathcal{O}_c,\, \mu\in Q^{\vee}) \end{align*}$$
    for $c\in C^J$ and J-generic $\mathfrak {t}\in T_J^\prime $ , allows one to describe the $\mathbb {H}$ -action on $\mathcal {P}^{(c)}_{\mathfrak {t}}$ entirely in terms of an $\mathbb {H}$ -action on the space of $\mathbf {F}$ -valued functions on $\mathcal {O}_c$ , with $\delta (T_j)$ ( $0\leq j\leq r$ ) acting by discrete Demazure-Lusztig type operators (cf. Remark 3.13). The subspace of finitely supported $\mathbf {F}$ -valued functions on $\mathcal {O}_c$ forms a $\mathbb {H}$ -submodule, which alternatively can be obtained as a quotient of Cherednik’s [Reference Cherednik8, §3.4.2] $\mathbb {H}$ -representation on the space of finitely supported $\mathbf {F}$ -valued functions on W (cf. Remark 3.13(3)).

Corollary 6.28. Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ . Suppose that $\mathfrak {t}$ is J-generic. Then

(6.25) $$ \begin{align} \pi_{c,\mathfrak{t}}(S_j^Y)P_y^J(x;\mathfrak{t})=\big(k_jt_y^{\alpha_j^{\vee}}-k_j^{-1})P_{s_jy}^J(x;\mathfrak{t}) \end{align} $$

for $j\in [0,r]$ and $y\in \mathcal {O}_c$ , where $t_y\in T$ is defined by (6.16). In particular,

(6.26) $$ \begin{align} \pi_{c,\mathfrak{t}}(S_j^Y)P_y^J(x;\mathfrak{t})=0\,\,\mbox{ if }\, s_jy=y. \end{align} $$

Proof. If $s_jy=y$ , then $t_y^{\alpha _j^{\vee }}=k_j^{-2}$ ; hence, the right-hand side of (6.25) is zero and the resulting equation (6.26) follows from (2.56) and Remark 6.25. If $s_jy\not =y$ , then (6.25) follows from (2.56) and (6.20) by a straightforward computation.

Remark 6.29. Under additional assumptions on $\mathfrak {t}$ , a proof of Corollary 6.28 that avoids Theorem 6.26 is as follows.

Formula (6.25) follows directly from the definition of $P_y^J(x;\mathfrak {t})$ (see (6.15)) when $s_jw_y\not \in W^J$ , and when $s_jw_y\in W^J$ and $\ell (s_jw_y)=\ell (w_y)+1$ . In the first case, use Proposition 6.8(1) and Lemma 6.23, in the second case, use $s_jw_y=w_{s_jy}$ and $r_{s_jw_y}(x)=(k_jx^{(w_y^{-1}\alpha _j)^{\vee }}-k_j^{-1})r_{w_y}(x)$ . When $s_jw_y\in W^J$ and $\ell (s_jw_y)=\ell (w_y)-1$ , acting by $S_j^Y$ on (6.25) and applying (2.59) gives

$$\begin{align*}(k_jt_y^{\alpha_j^{\vee}}-k_j^{-1})\pi_{c,\mathfrak{t}}(S_j^Y)P_{s_jy}^J(x;\mathfrak{t})=(k_jt_y^{\alpha_j^{\vee}}-k_j^{-1})(k_jt_y^{-\alpha_j^{\vee}}-k_j^{-1})P_y^J(x;\mathfrak{t}). \end{align*}$$

This yields (6.25) if the scalar factor $k_jt_y^{\alpha _j^{\vee }}-k_j^{-1}$ can be divided out. This is ensured if $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}\not =k_a^{-2}$ for all $a\in \Phi ^-\setminus \Phi _J^-$ . Note that these are exactly the additional requirements on $\mathfrak {t}$ ensuring the irreducibility of $\pi _{c,\mathfrak {t}}$ ; see Theorem 6.15.

The action of $\delta (T_j)$ on the monic quasi-polynomials $E_y^J(x;\mathfrak {t})$ can be described as follows. Recall the constants $\kappa _v(y)\in \mathbf {F}^\times $ ( $y\in E$ , $v\in W_0$ ), defined by (4.10).

Corollary 6.30. Let $c\in C^J$ and assume that $\mathfrak {t}\in T_J^\prime $ is J-generic.

The monic quasi-polynomials $E_y^J(x;\mathfrak {t})\in \mathcal {P}^{(c)}_{\mathfrak {t}}$ ( $y\in \mathcal {O}_c$ ) satisfy for $0\leq j\leq r$ ,

$$ \begin{align*} \pi_{c,\mathfrak{t}}(\delta(T_j))E_y^J(x;\mathfrak{t})= \begin{cases} k_jE_y^J(x;\mathfrak{t})\qquad &\mbox{ if }\,\, \alpha_j(y)=0,\\ \Big(\frac{k_j-k_j^{-1}}{1-t_y^{\alpha_j^{\vee}}}\Big)E_y^J(x;\mathfrak{t})+\kappa_{Ds_j}(y)E_{s_jy}^J(x;\mathfrak{t}) \qquad &\mbox{ if }\,\, \alpha_j(y)>0, \end{cases} \end{align*} $$

where $t_y\in T$ is defined by (6.16). Furthermore, for $y\in \mathcal {O}_c$ such that $\alpha _j(y)<0$ ,

$$ \begin{align*} \pi_{c,\mathfrak{t}}(\delta(T_j))E_y^J(x;\mathfrak{t})&= \left(\frac{k_j-k_j^{-1}}{1-t_y^{\alpha_j^{\vee}}}\right)E_y^J(x;\mathfrak{t})\\ &\quad+ \kappa_{Ds_j}(y)\frac{(k_j-k_j^{-1}t_y^{\alpha_j^{\vee}})(k_j-k_j^{-1}t_y^{-\alpha_j^{\vee}})} {\big(1-t_y^{\alpha_j^{\vee}}\big)\big(1-t_y^{-\alpha_j^{\vee}}\big)}E_{s_jy}^J(x;\mathfrak{t}). \end{align*} $$

Proof. The statement when $\alpha _j(y)=0$ is clear. If $\alpha _j(y)>0$ , then $s_jw_y=w_{s_jy}\in W^J$ and $\ell (s_jw_y)=\ell (w_y)+1$ by Corollary 5.23. Furthermore, $t_y^{\alpha _j^{\vee }}\not =1$ ; see the proof of Theorem 6.26. Then (6.23) and (6.9) give

$$\begin{align*}\pi_{c,\mathfrak{t}}(\delta(T_j))E_y^J(x;\mathfrak{t})=\left(\frac{k_j-k_j^{-1}}{1-t_y^{\alpha_j^{\vee}}}\right) E_y^J(x;\mathfrak{t})+\frac{d_{s_jw_y}(\mathfrak{s}_J\mathfrak{t})k_{s_jw_y}(c)}{d_{w_y}(\mathfrak{s}_J\mathfrak{t})k_{w_y}(c)} \left(\frac{1}{t_y^{\alpha_j^{\vee}}-1}\right)E_{s_jy}^J(x;\mathfrak{t}). \end{align*}$$

Simplifying the second factor using (6.8), (5.7) and Lemma 5.7 yields the desired result.

If $\alpha _j(y)<0$ , then the formula follows from a direct computation using the previous case, the Hecke relation (2.39) and Lemma 5.7.

6.6 (Anti)symmetrisation

By (4.10), we have $\kappa _v(0)=\prod _{\alpha \in \Pi (v)}k_\alpha $ . The element

$$\begin{align*}\mathbf{1}_+:=\sum_{v\in W_0}\kappa_v(0)T_v=\kappa_{w_0}(0)^2\sum_{v\in W_0}\kappa_v(0)^{-1}T_{v^{-1}}^{-1} \end{align*}$$

is, up to a normalisation factor, the trivial idempotent of $H_0$ . Concretely, it satisfies

$$\begin{align*}T_i\mathbf{1}_+=k_i\mathbf{1}_+=\mathbf{1}_+T_i\qquad (1\leq i\leq r) \end{align*}$$

and $\mathbf {1}_+^2=\big (\sum _{v\in W_0}\kappa _v(0)^2\big )\mathbf {1}_+$ .

The following result follows by a direct computation.

Lemma 6.31. Let $i\in [1,r]$ . Then

$$ \begin{align*} S_i^X\mathbf{1}_+=(k_ix^{\alpha_i^{\vee}}-k_i^{-1})\mathbf{1}_+,\qquad \mathbf{1}_+S_i^Y=\mathbf{1}_+(k_i(Y^{-1})^{\alpha_i^{\vee}}-k_i^{-1}) \end{align*} $$

in $\mathbb {H}$ . In particular, $\widetilde {S}_i^X\mathbf {1}_+=\mathbf {1}_+$ in $\mathbb {H}^{X-\text {loc}}$ and $\mathbf {1}_+\widetilde {S}_i^Y=\mathbf {1}_+$ in $\mathbb {H}^{Y-\text {loc}}$ .

For $c\in C^J$ and $\mathfrak {t}\in T_J$ , consider the $\mathcal {P}^{W_0}$ -submodule

$$\begin{align*}\mathcal{P}^{(c),+}_{\mathfrak{t}}:=\pi_{c,\mathfrak{t}}(\mathbf{1}_+)\mathcal{P}^{(c)}_{\mathfrak{t}} \end{align*}$$

of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ . The interpretation of $\mathcal {P}^{(c),+}_{\mathfrak {t}}$ as $W_0$ -invariant quasi-polynomials requires the $\mathbb {H}^{X-\text {loc}}$ -module $\mathcal {Q}^{(c)}_{\mathfrak {t}}$ from the proof of Theorem 4.9, which contains $\mathcal {P}_{\mathfrak {t}}^{(c)}$ as $\mathbb {H}$ -submodule.

Lemma 6.32. Let $c\in C^J$ and $\mathfrak {t}\in T_J$ . The $\mathcal {Q}^{W_0}$ -submodule $\mathcal {Q}^{(c),+}_{\mathfrak {t}}:=\pi _{c,\mathfrak {t}}^{X-\text {loc}}(\mathbf {1}_+)\mathcal {Q}^{(c)}_{\mathfrak {t}}$ of $\mathcal {Q}_{\mathfrak {t}}^{(c)}$ is contained in $\{f\in \mathcal {Q}^{(c)}_{\mathfrak {t}} \,\, | \,\, \sigma _{c,\mathfrak {t}}(v)f=f\quad \forall \, v\in W_0\}$ .

Proof. For $f\in \mathcal {Q}_{\mathfrak {t}}^{(c),+}$ , we have $\pi ^{X-\text {loc}}_{c,\mathfrak {t}}(\widetilde {S}_v^X)f=f$ for all $v\in W_0$ by Lemma 6.31. The lemma now follows from the definition (4.18) of the $W\ltimes \mathcal {Q}$ -action $\sigma _{c,\mathfrak {t}}$ .

Definition 6.33. Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ . Define $E_y^{J,+}(x;\mathfrak {t})\in \mathcal {P}^{(c),+}_{\mathfrak {t}}$ ( $y\in \mathcal {O}_c$ ) by

$$\begin{align*}E_y^{J,+}(x;\mathfrak{t}):=\pi_{c,\mathfrak{t}}(\mathbf{1}_+)E_y^J(x;\mathfrak{t}). \end{align*}$$

If furthermore $\mathfrak {t}$ is J-generic, then we define normalised versions $P_y^{J,+}(x;\mathfrak {t})\in \mathcal {P}_{\mathfrak {t}}^{(c),+}$ of $E_y^{J,+}(x;\mathfrak {t})$ by

$$ \begin{align*} P_y^{J,+}(x;\mathfrak{t}):=\pi_{c,\mathfrak{t}}(\mathbf{1}_+)P_y^J(x;\mathfrak{t})\qquad (y\in\mathcal{O}_c). \end{align*} $$

Note that $E_y^{J,+}(x;\mathfrak {t})\in \mathcal {P}^{(c),+}_{\mathfrak {t}}$ ( $y\in \mathcal {O}_c$ ) for $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ satisfies the eigenvalue equations

(6.27) $$ \begin{align} \pi_{c,\mathfrak{t}}(p(Y))E_y^{J,+}(x;\mathfrak{t})=p(\mathfrak{s}_y^{-1}\mathfrak{t}_y^{-1})E_y^{J,+}(x;\mathfrak{t})\qquad \forall\, p\in\mathcal{P}^{W_0}, \end{align} $$

since $\mathcal {P}_Y^{W_0}$ is the center of H. The same holds true for $P_y^{J,+}(x;\mathfrak {t})$ .

Remark 6.34. If $1_T\in T_{[1,r]}^\prime $ , then $E_\mu ^{[1,r],+}(x;1_T)\in \mathcal {P}^{W_0}$ ( $\mu \in Q^{\vee }$ ) is, up to a normalisation factor, the symmetric Macdonald polynomial of degree $\mu \in Q^{\vee }$ (see, for example, [Reference Macdonald29, §5.3]).

Proposition 6.35. Let $c\in C^J$ and assume that $\mathfrak {t}\in T_J^\prime $ is J-generic.

  1. 1. For $i\in [1,r]$ such that $\alpha _i(y)>0$ , we have

    $$\begin{align*}E_{s_iy}^{J,+}(x;\mathfrak{t})=k_i^{\eta(\alpha_i(y))}\left(\frac{k_i^{-1}-k_it_y^{\alpha_i^{\vee}}}{1-t_y^{\alpha_i^{\vee}}}\right)E_y^{J,+}(x;\mathfrak{t}), \end{align*}$$
    where $t_y\in T$ is given by (6.16).
  2. 2. $P_{s_iy}^{J,+}(x;\mathfrak {t})=P_y^{J,+}(x;\mathfrak {t})$ for $i\in [1,r]$ and $y\in \mathcal {O}_c$ .

Proof. (1) Let $i\in [1,r]$ such that $\alpha _i(y)>0$ . Then $t_y^{\alpha _i^{\vee }}\not =1$ ; see the proof of Theorem 6.26. We furthermore have $\kappa _{s_i}(y)=k_i^{-\eta (\alpha _i(y))}$ ; hence, Corollary 6.30 gives

$$ \begin{align*} \begin{aligned} E_{s_iy}^{J,+}(x;\mathfrak{t})&=k_i^{\eta(\alpha_i(y))}\pi_{c,\mathfrak{t}}(\mathbf{1}_+) \left(\pi_{c,\mathfrak{t}}(T_i)-\left(\frac{k_i-k_i^{-1}}{1-t_y^{\alpha_i^{\vee}}}\right)\right) E_y^J(x;\mathfrak{t})\\ &=k_i^{\eta(\alpha_i(y))}\left(\frac{k_i^{-1}-k_it_y^{\alpha_i^{\vee}}}{1-t_y^{\alpha_i^{\vee}}}\right) E_y^{J,+}(x;\mathfrak{t}), \end{aligned} \end{align*} $$

where the second equality follows by a direct computation using $\mathbf {1}_+T_i=k_i\mathbf {1}_+$ .

(2) It suffices to prove the statement under the additional assumption that $\alpha _i(y)>0$ . Then $w_{s_iy}=s_iw_y$ , $\ell (w_{s_iy})=\ell (w_y)+1$ and $\Pi (w_{s_iy})=\{w_y^{-1}\alpha _i\}\cup \Pi (w_y)$ ; hence,

$$\begin{align*}P_{s_iy}^{J,+}(x;\mathfrak{t})=\frac{\pi_{c,\mathfrak{t}}(\mathbf{1}_+S_{w_{s_iy}}^Y)x^c}{r_{w_{s_iy}}(\mathfrak{s}_J\mathfrak{t})}= \frac{\pi_{c,\mathfrak{t}}(\mathbf{1}_+S_i^YS_{w_y}^Y)x^c}{r_{w_y}(\mathfrak{s}_J\mathfrak{t}) (k_it_y^{\alpha_i^{\vee}}-k_i^{-1})}. \end{align*}$$

Applying Lemma 6.31, we conclude that

$$\begin{align*}P_{s_iy}^{J,+}(x;\mathfrak{t})=\frac{\pi_{c,\mathfrak{t}}(\mathbf{1}_+(k_i^{-1}-k_i(Y^{-1})^{\alpha_i^{\vee}}))P_y^J(x;\mathfrak{t})} {(k_i^{-1}-k_it_y^{\alpha_i^{\vee}})}= \pi_{c,\mathfrak{t}}(\mathbf{1}_+)P_y^J(x;\mathfrak{t})=P_y^{J,+}(x;\mathfrak{t}), \end{align*}$$

as desired.

For $c\in \overline {C}_+$ , write $\mathcal {O}_c^+:=\mathcal {O}_c\cap \overline {E}_+$ , which is a fundamental domain for the $W_0$ -action on $\mathcal {O}_c$ .

Corollary 6.36. Let $c\in C^J$ and assume that $\mathfrak {t}\in T_J^\prime $ is J-generic.

Then $\{E_y^{J,+}(x;\mathfrak {t})\}_{y\in \mathcal {O}_c^+}$ and $\{P_y^{J,+}(x;\mathfrak {t})\}_{y\in \mathcal {O}_c^+}$ are bases of $\mathcal {P}^{(c),+}_{\mathfrak {t}}$ consisting of simultaneous eigenfunctions of the commuting operators $\pi _{c,\mathfrak {t}}(p(Y))$ ( $p\in \mathcal {P}^{W_0}$ ).

Proof. By Proposition 6.35 and (6.27), it suffices to show that $\{E_y^{J,+}(x;\mathfrak {t})\}_{y\in \mathcal {O}_c^+}$ is a linear independent set. But Theorem 6.26 shows that $E_y^{J,+}(x;\mathfrak {t})$ lies in subspace spanned by $E_{vy}^J(x;\mathfrak {t})$ ( $v\in W_0$ ); hence, the linear independence is a consequence of Theorem 6.3(3).

The coefficients in the expansion of $E_y^{J,+}(x;\mathfrak {t})$ as linear combination of the quasi-polynomials $E_{vy}^J(x;\mathfrak {t})$ ( $v\in W_0$ ) can be explicitly computed as follows.

Recall that $y_{\pm }$ denotes the unique element in $\overline {E}_{\pm }\cap W_0y$ . Write $W_{0,y_{\pm }}$ for the subgroup of $W_0$ consisting of the elements $v\in W_0$ fixing $y_{\pm }$ . It is a parabolic subgroup. Denote by $W_0^{y_{\pm }}$ the minimal coset representatives of $W_0/W_{0,y_{\pm }}$ . Denote by $g_y\in W_0$ the unique element of minimal length such that $y_-=g_yy$ . Note that $y\mapsto g_y^{-1}$ defines a bijection $W_0y\overset {\sim }{\longrightarrow } W_0^{y_-}$ . Observe furthermore that $g_{y_+}$ is the minimal coset representative of $w_0W_{0,y_+}$ . In particular, $y_-=g_{y_+}y_+=w_0y_+$ .

Theorem 6.37. Let $c\in C^J$ and assume that $\mathfrak {t}\in T_J^\prime $ is J-generic. Set $t_y:=w_y(\mathfrak {s}_J\mathfrak {t})$ for $y\in \mathcal {O}_c$ (see (6.16)). For $y\in \mathcal {O}_c$ , we have

(6.28) $$ \begin{align} E_y^{J,+}(x;\mathfrak{t})=\sum_{y^\prime\in W_0y}\text{C}_y^+(y^\prime)E_{y^\prime}^{J}(x;\mathfrak{t}) \end{align} $$

with the coefficients $C_y^+(y^\prime )\in \mathbf {F}$ determined by the following two properties:

(6.29) $$ \begin{align} C_{s_iy}^+(y^\prime)=k_i^{\eta(\alpha_i(y))}\left(\frac{k_i-k_i^{-1}t_y^{-\alpha_i^{\vee}}} {1-t_y^{-\alpha_i^{\vee}}}\right)C_y^+(y^\prime)\,\, \mbox{ for }\,\, 1\leq i\leq r\, \mbox{ such that }\, \alpha_i(y)>0 \end{align} $$

and

(6.30) $$ \begin{align} C_{y_+}^+(y^\prime)=\left(\sum_{v\in W_{0,y_+}}\kappa_v(0)^2\right)\kappa_{g_{y_+}}(0)\kappa_{g_{y_+}}(y_+) \prod_{\beta\in\Pi(g_{y^\prime}^{-1})}k_\beta^{-\eta(\beta(y_-))} \left(\frac{k_\beta-k_\beta^{-1}t_{y_-}^{-\beta^{\vee}}}{1-t_{y_-}^{-\beta^{\vee}}}\right) \end{align} $$

for $y^\prime \in W_0y$ .

Proof. Note that there exists an expansion of the form (6.28) for unique $C_y^+(y^\prime )\in \mathbf {F}$ ; cf. Corollary 6.36. It is clear that the two conditions (6.29) and (6.30) determine the coefficients $C_y(y^\prime )\in \mathbf {F}$ uniquely.

The recursion relation (6.29) follows from Proposition 6.35(1). It thus remains to prove (6.30).

Fix $y\in \mathcal {O}_c$ . Note that

$$\begin{align*}\mathbf{1}_+=\mathbf{1}_+^{y_+}\sum_{v\in W_{0,y_+}}\kappa_v(0)T_v \end{align*}$$

with $\mathbf {1}_+^{y_+}:=\sum _{u\in W_0^{y_+}}\kappa _u(0)T_u$ . Then Corollary 6.30 gives

$$\begin{align*}E_{y_+}^{J,+}(x;\mathfrak{t})=\left(\sum_{v\in W_{0,y_+}}\kappa_v(0)^2\right) \pi_{c,\mathfrak{t}}(\mathbf{1}_+^{y_+})E_{y_+}^{J}(x;\mathfrak{t}). \end{align*}$$

Note that for $u\in W_0^{y_+}$ and $1\leq i\leq r$ , one has $\alpha _i(uy_+)\gtrless 0$ iff $s_iu\in W_0^{y_+}$ and $\ell (s_iu)=\ell (u)\pm 1$ (this follows from the Weyl group variant of Corollary 5.23 with W and its fundamental alcove $\overline {C}_+$ replaced by $W_0$ and its fundamental Weyl chamber $\overline {E}_{+}$ ). Again invoking Corollary 6.30, we then get from Lemma 5.6,

(6.31) $$ \begin{align} C_{y_+}^+(y_-)=\left(\sum_{v\in W_{0,y_+}}\kappa_v(0)^2\right)\kappa_{g_{y_+}}(0)\kappa_{g_{y_+}}(y_+). \end{align} $$

We now derive the coefficients $C_{y_+}^+(y^\prime )$ for arbitrary $y^\prime \in W_0y$ recursively using the recursion relation

(6.32) $$ \begin{align} C_{y^{\prime\prime}}^+(s_iy^\prime)=C_{y^{\prime\prime}}^+(y^\prime)k_i^{-\eta(\alpha_i(y^\prime))} \left( \frac{k_i-k_i^{-1}t_{y^\prime}^{-\alpha_i^{\vee}}}{1-t_{y^\prime}^{-\alpha_i^{\vee}}}\right) \, \mbox{ for }\, y^\prime\in W_0y\, \mbox{ such that }\, \alpha_i(y^\prime)<0 \end{align} $$

for $y^{\prime \prime }\in W_0y$ . The recursion relation (6.32) follows from the identity

$$\begin{align*}\pi_{c,\mathfrak{t}}(T_i)E_{y^{\prime\prime}}^{J,+}(x;\mathfrak{t})=k_iE_{y^{\prime\prime}}^{J,+}(x;\mathfrak{t}), \end{align*}$$

expanding $E_{y^{\prime \prime }}^{J,+}(x;\mathfrak {t})$ as linear combination of quasi-polynomials $E_{y^{\prime \prime \prime }}^J(x;\mathfrak {t})$ ( $y^{\prime \prime \prime }\in W_0y$ ), and subsequently comparing the coefficient of $E_{y^\prime }^J(x;\mathfrak {t})$ on both sides using Corollary 6.30.

Now take $y^\prime \in W_0y$ and let $g_{y^\prime }^{-1}=s_{i_1}\cdots s_{i_\ell }$ be a reduced expression. Then $\Pi (g_{y^\prime }^{-1})=\{\beta _1,\ldots ,\beta _\ell \}$ with $\beta _m:=s_{i_\ell }\cdots s_{i_{m+1}}(\alpha _{i_m})$ for $1\leq m\leq \ell $ . Furthermore,

$$\begin{align*}\beta_m(y_-)=\alpha_{i_m}(s_{i_{m+1}}\cdots s_{i_\ell}y_-)<0\qquad\quad (1\leq m\leq\ell) \end{align*}$$

since $g_{y^\prime }^{-1}\in W_0^{y_-}$ (this follows from the Weyl group variant of Corollary 5.23 with W and $\overline {C}_+$ replaced by $W_0$ and $\overline {E}_{-}$ ). By repeated application of (6.32), we conclude that

$$\begin{align*}C_{y_+}^+(y^\prime)=C_{y_+}^+(y_-)\prod_{\beta\in\Pi(g_{y^\prime}^{-1})}k_\beta^{-\eta(\beta(y_-))}\left(\frac{k_\beta-k_\beta^{-1}t_{y_-}^{-\beta^{\vee}}}{1-t_y^{-\beta^{\vee}}}\right). \end{align*}$$

This completes the proof of the theorem.

When considering the Whittaker limit of $E_y^{J,+}(x;\mathfrak {t})$ in Subsection 6.8, it is convenient to use the explicit expansion formula for $E_y^{J,+}(x;\mathfrak {t})$ when $y\in \overline {E}_-\cap \mathcal {O}_c$ .

Corollary 6.38. Let $c\in C^J$ , and assume that $\mathfrak {t}\in T_J^\prime $ is J-generic. For $y\in \overline {E}_-\cap \mathcal {O}_c$ , we have

$$ \begin{align*} \begin{aligned} E_{y}^{J,+}(x;\mathfrak{t})&=\left(\sum_{v\in W_{0,y}}\kappa_{g_{y_+}^{-1}v}(0)^2\right)\prod_{\alpha\in\Pi(g_{y_+}^{-1})}\left(\frac{1-k_\alpha^{-2}t_{y}^{\alpha^{\vee}}} {1-t_{y}^{\alpha^{\vee}}}\right)\\ &\qquad\quad\times \sum_{v\in W_0^{y}}\frac{\kappa_v(y)}{\kappa_v(0)}\left(\prod_{\beta\in\Pi(v)}\left(\frac{1-k_\beta^{2}t_{y}^{\beta^{\vee}}}{1-t_{y}^{\beta^{\vee}}}\right)\right) E_{vy}^J(x;\mathfrak{t}), \end{aligned} \end{align*} $$

where $t_y\in T$ is defined by (6.16).

Proof. Using Lemma 5.6 and (6.29), we have for $y\in \overline {E}_-\cap \mathcal {O}_c$ ,

$$\begin{align*}C_{y}^+(y^\prime)= \frac{\kappa_{g_{y_+}^{-1}}(0)^2}{\kappa_{g_{y_+}}(0)\kappa_{g_{y_+}}(y_+)}\left(\prod_{\alpha\in\Pi(g_{y_+}^{-1})}\left( \frac{1-k_\alpha^{-2}t_{y}^{\alpha^{\vee}}}{1-t_{y}^{\alpha^{\vee}}}\right)\right)C_{y_+}^+(y^\prime). \end{align*}$$

Substituting (6.30), the desired result follows after a direct computation using the fact that $\kappa _v(0)=\prod _{\alpha \in \Pi (v)}k_\alpha $ and $\kappa _v(y)=\prod _{\alpha \in \Pi (v)}k_\alpha ^{-\eta (\alpha (y))}$ for $v\in W_0^y$ (cf. Lemma 5.6).

Within the ambient $\mathbb {H}^{X-\text {loc}}$ -module $\mathcal {Q}_{\mathfrak {t}}^{(c)}$ of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ , we can express $E_y^{J,+}(x;\mathfrak {t})$ as linear combination of the $\sigma _{c,\mathfrak {t}}(W_0)$ -translates of $E_y^J(x;\mathfrak {t})$ as follows.

Proposition 6.39. Let $c\in C^J$ and assume that $\mathfrak {t}\in T_J^\prime $ is J-generic. For $y\in \mathcal {O}_c$ , we have

$$\begin{align*}E_y^{J,+}(x;\mathfrak{t})=\sum_{v\in W_0}\kappa_v(0)^2\left(\prod_{\alpha\in\Phi_0^+}\frac{1-k_\alpha^{2\chi(v^{-1}\alpha)}x^{-\alpha^{\vee}}}{1-x^{-\alpha^{\vee}}}\right) \sigma_{c,\mathfrak{t}}(v)E_y^J(x;\mathfrak{t}) \end{align*}$$

where (recall) $\chi (\beta )=\pm 1$ if $\beta \in \Phi _0^{\pm }$ .

Proof. By [Reference Macdonald29, (5.5.14)], we have

$$\begin{align*}\beth^{-1}(\mathbf{1}_+)=\left(\sum_{v\in W_0}v\right)\prod_{\alpha\in\Phi_0^+}\left(\frac{1-k_\alpha^2x^{-\alpha^{\vee}}}{1-x^{-\alpha^{\vee}}}\right), \end{align*}$$

and hence, in view of (4.19) and the definition of $E_y^{J,+}(x;\mathfrak {t})$ ,

$$\begin{align*}E_y^{J,+}(x;\mathfrak{t})=\pi_{c,\mathfrak{t}}^{X-\text{loc}}(\mathbf{1}_+)E_y^J(x;\mathfrak{t})= \sum_{v\in W_0}\sigma_{c,\mathfrak{t}}(v)\left(E_y^J(x;\mathfrak{t})\prod_{\alpha\in\Phi_0^+}\left(\frac{1-k_\alpha^2x^{-\alpha^{\vee}}}{1-x^{-\alpha^{\vee}}}\right) \right) \end{align*}$$

in $\mathcal {Q}^{(c)}$ . The result now follows by a straightforward computation.

Similar results can be obtained with respect to Hecke anti-symmetrisation. Write

(6.33) $$ \begin{align} \mathbf{1}_-:=\sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v(0)^{-1}T_v=\kappa_{w_0}(0)^{-2}\sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v(0)T_{v^{-1}}^{-1}, \end{align} $$

which satisfies $T_i\mathbf {1}_-=-k_i^{-1}\mathbf {1}_-=\mathbf {1}_-T_i$ ( $1\leq i\leq r$ ).

Definition 6.40. Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ . Define $E_y^-(x;\mathfrak {t})\in \mathcal {P}_{\mathfrak {t}}^{(c)}$ ( $y\in \mathcal {O}_c$ ) by

(6.34) $$ \begin{align} E_y^{J,-}(x;\mathfrak{t}):=\pi_{c,\mathfrak{t}}(\mathbf{1}_-)E_y^J(x;\mathfrak{t}). \end{align} $$

We state below the basic formulas for $E_y^{J,-}(x;\mathfrak {t})$ , but we do not provide detailed proofs, since they are similar to the proofs for $E_y^{J,+}(x;\mathfrak {t})$ .

Let $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ . Suppose that $\mathfrak {t}$ is J-generic. Let $t_y\in T$ be given by (6.16). Then

$$\begin{align*}E_{s_iy}^{J,-}(x;\mathfrak{t})=-k_i^{\eta(\alpha_i(y))}\left(\frac{k_i-k_i^{-1}t_y^{\alpha_i^{\vee}}}{1-t_y^{\alpha_i^{\vee}}}\right)E_y^{J,-}(x;\mathfrak{t})\,\mbox{ for }\, 1\leq i\leq r\, \mbox{ such that }\, \alpha_i(y)>0 \end{align*}$$

and

(6.35) $$ \begin{align} E_y^{J,-}(x;\mathfrak{t})=\left(\prod_{\alpha\in\Phi_0^+}\frac{1-k_\alpha^{-2}x^{-\alpha^{\vee}}}{1-x^{-\alpha^{\vee}}}\right) \sum_{v\in W_0}(-1)^{\ell(v)}\sigma_{c,\mathfrak{t}}(v)E_y^J(x;\mathfrak{t}). \end{align} $$

Furthermore,

(6.36) $$ \begin{align} E_y^{J,-}(x;\mathfrak{t})=0\,\, \mbox{ if }\,\, y\not\in E^{\text{reg}}\cap\mathcal{O}_c \end{align} $$

and

(6.37) $$ \begin{align} E_y^{J,-}(x;\mathfrak{t})=\sum_{y^\prime\in W_0y}C_y^-(y^\prime)E_{y^\prime}^J(x;\mathfrak{t})\qquad (y\in\mathcal{O}_c) \end{align} $$

with $C_y^-(y^\prime )\in \mathbf {F}$ satisfying, for all $y^\prime \in W_0y$ ,

$$\begin{align*}C_{s_iy}^-(y^\prime)=k_i^{\eta(\alpha_i(y))}\left(\frac{k_i^{-1}-k_it_y^{-\alpha_i^{\vee}}}{t_y^{-\alpha_i^{\vee}}-1}\right)C_y^-(y^\prime)\, \mbox{ for }\, 1\leq i\leq r\, \mbox{ such that }\, \alpha_i(y)>0, \end{align*}$$

$C_{y_+}^-(y^\prime )=0$ if $y_+\in \overline {E}_+\setminus E_+$ and, for $y\in E^{\text {reg}}\cap \mathcal {O}_c$ ,

$$\begin{align*}C_{y_+}^-(y^\prime)=(-1)^{\ell(w_0)}\frac{\kappa_{w_0}(y_+)}{\kappa_{w_0}(0)}\prod_{\beta\in\Pi(g_{y^\prime}^{-1})} k_\beta^{-\eta(\beta(y_-))}\left(\frac{k_\beta^{-1}-k_\beta t_{y_-}^{-\beta^{\vee}}}{t_{y_-}^{-\beta^{\vee}}-1}\right)\qquad (y^\prime\in W_0y). \end{align*}$$

For the Whittaker limit of $E_y^{J,-}(x;\mathfrak {t})$ (see Subsection 6.8), it is again convenient to use the explicit expansion of $E_{y_-}^{J,-}(x;\mathfrak {t})$ in the quasi-polynomials $E_{vy_-}^J(x;\mathfrak {t})$ ( $v\in W_0$ ) when $y_-\in E_-$ .

Proposition 6.41. Let $c\in C^J$ and suppose that $\mathfrak {t}\in T_J^\prime $ is J-generic. For $y\in E_-\cap \mathcal {O}_c$ , we have

$$ \begin{align*} \begin{aligned} E_y^{J,-}(x;\mathfrak{t})=\kappa_{w_0}(0)^{-2}&\prod_{\alpha\in\Phi_0^+}\left(\frac{1-k_\alpha^{2}t_y^{\alpha^{\vee}}}{1-t_y^{\alpha^{\vee}}}\right)\\ &\times \sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v(0)\kappa_v(y)\left(\prod_{\beta\in\Pi(v)}\left(\frac{1-k_\beta^{-2}t_y^{\beta^{\vee}}}{1-t_y^{\beta^{\vee}}}\right)\right)E_{vy}^J(x;\mathfrak{t}), \end{aligned} \end{align*} $$

where $t_y\in T$ is defined by (6.16).

6.7 Pseudo-unitarity and orthogonality relations

Let $d\mapsto d^\ast $ be a nontrivial automorphism of $\mathbf {F}$ of order two, and write

$$ \begin{align*} \begin{aligned} \mathbf{F}_u^\times&:=\{d\in\mathbf{F}^\times \,\, | \,\, d^\ast=d^{-1} \},\\ \mathbf{F}_r&:=\{d\in\mathbf{F} \,\, | \,\, d^\ast=d\}. \end{aligned} \end{align*} $$

for the subgroup (resp. subfield) of unitary (resp. real) elements in $\mathbf {F}$ . Write $T_u:=\text {Hom}(Q^{\vee },\mathbf {F}_u^\times )$ and $T_r:=\text {Hom}(Q^{\vee },\mathbf {F}_r^\times )$ for the associated compact and real tori, respectively. We assume in this subsection that $q,k_a\in \mathbf {F}_u^\times $ for all $a\in \Phi $ . In case a root of q is required (see Subsection 3.1), then we assume it to lie in $\mathbf {F}_u^\times $ too. Note that under these assumptions, $\mathfrak {s}_J\in T_u$ for all $J\subsetneq [0,r]$ .

The formulas

$$\begin{align*}T_w^\ast=T_w^{-1}\quad (w\in W),\qquad (x^\mu)^*=x^{-\mu} \quad (\mu\in Q^{\vee}) \end{align*}$$

extend the automorphism $d\mapsto d^\ast $ of $\mathbf {F}$ to a $\mathbf {F}_r$ -algebra anti-involution $h\mapsto h^\ast $ of $\mathbb {H}$ . It satisfies $(Y^\mu )^*=Y^{-\mu }$ for all $\mu \in Q^{\vee }$ . The following definition is from [Reference Cherednik8, §3.6].

Definition 6.42. Let M be a $\mathbb {H}$ -module.

  1. 1. A $\mathbf {F}_r$ -bilinear map $(\cdot ,\cdot ): M\times M\rightarrow \mathbf {F}$ is said to be $\ast $ -sesquilinear if

    1. (a) $(\cdot ,\cdot )$ is $\mathbf {F}$ -linear in the first component,

    2. (b) $(m,m^\prime )=(m^\prime ,m)^*$ for all $m,m^\prime \in M$ .

  2. 2. We say that M is pseudo-unitarizable if there exists a nonzero $\ast $ -sesquilinear form $(\cdot ,\cdot )$ on M such that

    $$\begin{align*}(hm,m^\prime)=(m,h^*m^\prime)\qquad \forall\, h\in\mathbb{H},\,\,\,\forall\, m,m^\prime\in M. \end{align*}$$

Define $N^w\in \mathcal {Q}$ ( $w\in W$ ) by

(6.38) $$ \begin{align} N^w(x):=\prod_{a\in\Pi(w)}\left( \frac{k_a^{-1}x^{a^{\vee}}-k_a}{k_ax^{a^{\vee}}-k_a^{-1}}\right). \end{align} $$

The following theorem relates to [Reference Cherednik8, Thm. 3.6.1] via Remark 6.12.

Theorem 6.43. Let $c\in C^J$ and fix a J-generic $\mathfrak {t}\in T_J^\prime \cap \text {Hom}(Q^{\vee },\mathbf {F}_u^\times )$ .

  1. 1. There exists a unique $\ast $ -sesquilinear form $(\cdot ,\cdot )_{J,\mathfrak {t}}$ on $\mathcal {P}_{\mathfrak {t}}^{(c)}$ such that

    (6.39) $$ \begin{align} \big(P_y^J(x;\mathfrak{t}),P_{y^\prime}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}} :=\delta_{y,y^\prime}N^{w_y}(\mathfrak{s}_J\mathfrak{t}) \qquad \forall\, y,y^\prime\in\mathcal{O}_c. \end{align} $$
  2. 2. $\mathcal {P}_{\mathfrak {t}}^{(c)}$ is pseudo-unitarizable.

  3. 3. If $(\cdot ,\cdot )$ is a nonzero $\ast $ -sesquilinear form on $\mathcal {P}^{(c)}_{\mathfrak {t}}$ realising the pseudo-unitarity of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ , then $(\cdot ,\cdot )=d\,(\cdot ,\cdot )_{J,\mathfrak {t}}$ for some $d\in \mathbf {F}^\times _r$ .

Proof. Write $t_y:=\mathfrak {s}_y\mathfrak {t}_y$ for $y\in \mathcal {O}_c$ (see (6.16)). We have $t_y\in \text {Hom}(Q^{\vee },\mathbf {F}_u^\times )$ by the assumptions on $q,k_a$ and $\mathfrak {t}$ .

(1) For $w\in W^J$ the quadratic norm $N^w(\mathfrak {s}_J\mathfrak {t})$ is well defined since $\mathfrak {t}$ is J-generic (it may be zero). Furthermore, $\{P_y^J(x;\mathfrak {t})\}_{y\in \mathcal {O}_c}$ is a $\mathbf {F}$ -basis of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ . The result then follows directly from the observation that $N^w(\mathfrak {s}_J\mathfrak {t})\in \mathbf {F}_r$ for all $w\in W^J$ .

(2) We show that $\mathcal {P}^{(c)}_{\mathfrak {t}}$ is pseudo-unitary with respect to $(\cdot ,\cdot )_{J,\mathfrak {t}}$ . It suffices to show that

$$\begin{align*}\big(\pi_{c,\mathfrak{t}}(\delta(h))P_y^J(x;\mathfrak{t}),P_{y^\prime}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}}=\big(P_y^J(x;\mathfrak{t}), \pi_{c,\mathfrak{t}}(\delta(h)^*)P_{y^\prime}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}} \end{align*}$$

for $h=x^\mu $ ( $\mu \in Q^{\vee }$ ) and $h=T_j$ ( $0\leq j\leq r$ ).

For $h=x^\mu $ , we have $\delta (x^\mu )=Y^{-\mu }$ ; hence, $\delta (x^{\mu })^*=Y^\mu $ . Since $t_y^\mu \in \mathbf {F}_u$ and the $P_y^J(x;\mathfrak {t})$ are orthogonal with respect to $(\cdot ,\cdot )_{J,\mathfrak {t}}$ , we have

$$ \begin{align*} \begin{aligned} \big(\pi_{c,\mathfrak{t}}(Y^{-\mu})P_y^J(x;\mathfrak{t}),P_{y^\prime}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}}&= t_y^\mu\big(P_y^J(x;\mathfrak{t}),P_{y^\prime}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}}\\ &=t_{y^\prime}^\mu N^{w_y}(\mathfrak{s}_J\mathfrak{t})\delta_{y,y^\prime}= \big(P_y^J(x;\mathfrak{t}),\pi_{c,\mathfrak{t}}(Y^\mu)P_{y^\prime}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}} \end{aligned} \end{align*} $$

for $y,y^\prime \in \mathcal {O}_c$ .

Finally, consider the case that $h=T_j$ ( $0\leq j\leq r$ ). Then $\delta (T_j)^*=\delta (T_j^{-1})$ . By Theorem 6.26 and Corollary 5.23, the following two statements are equivalent:

  1. (a) For all $y,y^\prime \in \mathcal {O}_c$ ,

    $$\begin{align*}\big(\pi_{c,\mathfrak{t}}(\delta(T_j))P_y^J(x;\mathfrak{t}),P_{y^\prime}^J(x;\mathfrak{t}))_{J,\mathfrak{t}}= (P_y^J(x;\mathfrak{t}),\pi_{c,\mathfrak{t}}(\delta(T_j^{-1}))P_{y^\prime}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}}. \end{align*}$$
  2. (b) For $w\in W^J$ satisfying $s_jw\in W^J$ ,

    $$\begin{align*}\left(\frac{k_jt_{wc}^{\alpha_j^{\vee}}-k_j^{-1}}{t_{wc}^{\alpha_j^{\vee}}-1}\right) N^{s_jw}(\mathfrak{s}_J\mathfrak{t})=\left(\frac{k_j^{-1}t_{wc}^{\alpha_j^{\vee}}-k_j}{t_{wc}^{\alpha_j^{\vee}}-1}\right) N^w(\mathfrak{s}_J\mathfrak{t}). \end{align*}$$

Statement (b) can be checked directly using Lemma 2.1 and (6.38), which completes the proof.

(3) Suppose that $(\cdot ,\cdot )$ is a nonzero $\ast $ -sesquilinear form on $\mathcal {P}^{(c)}_{\mathfrak {t}}$ which also realises the pseudo-unitarity of $\mathcal {P}^{(c)}_{\mathfrak {t}}$ . Set

$$\begin{align*}d:=(x^c,x^c)\in\mathbf{F}_r. \end{align*}$$

We will show that $(\cdot ,\cdot )=d\,(\cdot ,\cdot )_{J,\mathfrak {t}}$ (and consequently, $d\in \mathbf {F}_r^\times $ ).

By the simplicity of the Y-spectrum $\mathcal {S}(\mathcal {P}_{\mathfrak {t}}^{(c)})$ , we have

$$\begin{align*}(P_y^{J}(x;\mathfrak{t}),P_{y^\prime}^J(x;\mathfrak{t}))=\delta_{y,y^\prime}\big(P_y^J(x;\mathfrak{t}),P_y^J(x;\mathfrak{t})\big)\qquad (y,y^\prime\in\mathcal{O}_c). \end{align*}$$

We have $(S_j^Y)^*=S_j^Y$ for $j\in [0,r]$ (this can be verified by looking at its $\delta $ -image and using (2.40)); hence,

$$\begin{align*}(S_w^Y)^*=S_{w^{-1}}^Y\qquad (w\in W). \end{align*}$$

For $w\in W^J$ , we then obtain

$$ \begin{align*} \begin{aligned} \big(P_{wc}^J(x;\mathfrak{t}),P_{wc}^J(x;\mathfrak{t})\big)&=\frac{\big(\pi_{c,\mathfrak{t}}(S_{w^{-1}}^YS_w^Y)x^c,x^c\big)}{r_w(\mathfrak{s}_J\mathfrak{t})^*r_w(\mathfrak{s}_J\mathfrak{t})}\\ &=\frac{(x^c,x^c)\text{n}_w(\mathfrak{s}_J\mathfrak{t})}{r_w(\mathfrak{s}_J\mathfrak{t})^*r_w(\mathfrak{s}_J\mathfrak{t})}\\ &=\big(x^c,x^c\big)N^w(\mathfrak{s}_J\mathfrak{t})= d\,\big(P_{wc}^J(x;\mathfrak{t}),P_{wc}^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}}, \end{aligned} \end{align*} $$

where the first equality is by definition of $P_y^J(x;\mathfrak {t})$ (see (6.15)), the second equality follows from (6.10), and the third equality from the explicit expressions (6.11), (6.14) and (6.38) of $\text {n}_w(x)$ , $r_w(x)$ and $N^w(x)$ , respectively. It follows that $(\cdot ,\cdot )=d\,(\cdot ,\cdot )_{J,\mathfrak {t}}$ , as desired.

It is easy to check that

$$\begin{align*}\Phi^+=\bigcup_{w\in W}\Pi(w). \end{align*}$$

For $J\subsetneq [0,r]$ , this implies that

$$\begin{align*}\Phi^+\setminus\Phi_J^+=\bigcup_{(w,u)\in W^J\times W_J}u\big(\Pi(w)\big). \end{align*}$$

In the present context, Theorem 6.43 gives the following sharpening of Theorem 6.15.

Corollary 6.44. Keep the assumptions of Theorem 6.43. Then

  1. 1. $(\cdot ,\cdot )_{J,\mathfrak {t}}$ is nondegenerate iff

    $$\begin{align*}(\mathfrak{s}_J\mathfrak{t})^{a^{\vee}}\not=k_a^2\qquad \forall\, a\in\bigcup_{w\in W^J}\Pi(w). \end{align*}$$
  2. 2. $\mathcal {P}_{\mathfrak {t}}^{(c)}$ is reducible if $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}=k_a^{2}$ for some $a\in \bigcup _{w\in W^J}\Pi (w)$ .

Proof. (1) By (6.39), the sesquilinear form $(\cdot ,\cdot )_{J,\mathfrak {t}}$ is degenerate iff $N^w(\mathfrak {s}_J\mathfrak {t})=0$ for some $w\in W^J$ . In view of (6.38), this is equivalent to $(\mathfrak {s}_J\mathfrak {t})^{a^{\vee }}=k_a^2$ for some $a\in \Pi (w)$ .

(2) The radical

$$\begin{align*}\text{Rad}_{J,\mathfrak{t}}:=\{p\in\mathcal{P}\,\, | \,\, (p,\cdot)_{J,\mathfrak{t}}\equiv 0\} \end{align*}$$

of $(\cdot ,\cdot )_{J,\mathfrak {t}}$ is a proper $\mathbb {H}$ -submodule of $\mathcal {P}_{\mathfrak {t}}^{(c)}$ by Theorem 6.43. It is nonzero iff $(\cdot ,\cdot )_{J,\mathfrak {t}}$ is degenerate. The result now follows from part (1).

Note that by $\Pi (w_y)=-w_y^{-1}\Pi (w_y^{-1})$ , (2.31), Corollary 5.22, and (6.38), we have the explicit expressions

$$\begin{align*}N^{w_y}(\mathfrak{s}_J\mathfrak{t})=\prod_{a\in\Phi^+: a(y)<0} \left(\frac{k_a^{-1}t_y^{-a^{\vee}}-k_a}{k_at_y^{-a^{\vee}}-k_a^{-1}}\right)\qquad (y\in\mathcal{O}_c) \end{align*}$$

for the quadratic norms, where $t_y=\mathfrak {s}_y\mathfrak {t}_y\in T$ . By Proposition 6.20, the quadratic norms for the monic quasi-polynomials $E_y^J(x;\mathfrak {t})$ are

$$\begin{align*}\big(E_y^J(x;\mathfrak{t}),E_y^J(x;\mathfrak{t})\big)_{J,\mathfrak{t}}=\prod_{a\in\Phi^+: a(y)<0}\frac{(k_at_y^{-a^{\vee}}-k_a^{-1})(k_a^{-1}t_y^{-a^{\vee}}-k_a)} {(t_y^{-a^{\vee}}-1)^2}\qquad (y\in\mathcal{O}_c). \end{align*}$$

Remark 6.45. In case of Cherednik’s polynomial representation $\pi =\pi _{0,1_T}$ , there exists a concrete realization of the $\ast $ -sesquilinear form $(\cdot ,\cdot )_{0,1_T}: \mathcal {P}\times \mathcal {P}\rightarrow \mathbf {F}$ . It is of the form

$$\begin{align*}(p_1,p_2)_{[1,r],1_T}=\text{ct}(p_1p_2^*\mathcal{W})\qquad (p_1,p_2\in\mathcal{P}) \end{align*}$$

with $p^\ast :=\sum _{\mu \in Q^{\vee }}d_\mu ^\ast x^{-\mu }$ for $p=\sum _{\mu \in Q^{\vee }}d_\mu x^\mu \in \mathcal {P}$ , with $\text {ct}$ the constant term map (picking the coefficient of $x^0$ in the expansion in monomials), and with $\mathcal {W}$ an explicit weight function (see, for example, [Reference Macdonald29, §5.1]). Such a realisation for $(\cdot ,\cdot )_{J,\mathfrak {t}}$ is not known when $(J,\mathfrak {t})\not =([1,r],1_T)$ .

6.8 The Whittaker limit

In this section, we take $\mathbf {F}=\mathbf {K}(q^{\frac {1}{2h}})$ for some field $\mathbf {K}$ of characteristic zero and we assume that ${}^{\text {sh}}k^{\frac {1}{2}}$ and ${}^{\text {lg}}k^{\frac {1}{2}}$ are in $\mathbf {K}$ . Set

$$\begin{align*}\overline{\mathcal{P}}^{(c)}:=\bigoplus_{y\in\mathcal{O}_c}\mathbf{K}x^y\subset\mathbf{K}[E]\qquad (c\in\overline{C}_+). \end{align*}$$

Let $\overline {\mathcal {Q}}$ be the quotient field of $\overline {\mathcal {P}}:=\overline {\mathcal {P}}^{(0)}$ , and set

$$\begin{align*}\mathbf{K}_{\overline{\mathcal{Q}}}[E]:=\overline{\mathcal{Q}}\otimes_{\overline{\mathcal{P}}}\mathbf{K}[E]. \end{align*}$$

For $c\in \overline {C}_+$ , let $\overline {\mathcal {Q}}^{(c)}\subset \mathbf {K}_{\overline {\mathcal {Q}}}[E]$ be the $\overline {\mathcal {Q}}$ -submodule generated by $\overline {\mathcal {P}}^{(c)}$ . We view $\overline {\mathcal {Q}}^{(c)}$ as $\mathbf {K}$ -submodule of $\mathcal {Q}^{(c)}$ in the natural way.

Denote by $\overline {H}_0\subset \overline {H}$ the finite and affine Hecke algebra over $\mathbf {K}$ , respectively. They are regarded as $\mathbf {K}$ -subalgebras of H in the natural way.

Lemma 6.46.

  1. 1. The formulas

    (6.40) $$ \begin{align} (fx^y)\blacktriangleleft T_i:=k_i^{\chi_{\mathbb{Z}}(\alpha_i(y))}s_i(fx^{y})+(k_i-k_i^{-1})\left(\frac{fx^y-s_i(f)x^{y-\lfloor \alpha_i(y)\rfloor\alpha_i^{\vee}}}{1-x^{\alpha_i^{\vee}}}\right) \end{align} $$
    for $1\leq i\leq r$ , $f\in \overline {\mathcal {Q}}$ and $y\in E$ define a $\mathbf {K}$ -linear right $\overline {H}_0$ -action on $\mathbf {K}_{\overline {\mathcal {Q}}}[E]$ .
  2. 2. The $\overline {H}_0$ -action (6.40) on $\mathbf {K}_{\overline {\mathcal {Q}}}[E]$ extends to a $\mathbf {K}$ -linear right $\overline {H}$ -action by

    (6.41) $$ \begin{align} (fx^y)\blacktriangleleft T_0:=(fx^{y-\varphi^{\vee}})\blacktriangleleft T_{s_\varphi}^{-1} \end{align} $$
    for $f\in \overline {\mathcal {Q}}$ and $y\in E$ .
  3. 3. $\mathbf {K}[E]$ , $\overline {\mathcal {Q}}^{(c)}$ and $\overline {\mathcal {P}}^{(c)}$ are $\overline {H}$ -submodules of $\mathbf {K}_{\overline {\mathcal {Q}}}[E]$ for all $c\in \overline {C}_+$ .

  4. 4. Let $c\in C^J$ . The restriction $\pi _{c,\mathfrak {t}}^{X-\text {loc}}\vert _{\delta (H)}$ of the X-localised quasi-polynomial representation $\pi _{c,\mathfrak {t}}^{X-\text {loc}}: \mathbb {H}^{X-\text {loc}}\rightarrow \text {End}(\mathcal {Q}^{(c)})$ to $\delta (H)\subset \mathbb {H}^{X-\text {loc}}$ does not depend on $\mathfrak {t}\in T_J$ . Furthermore,

    (6.42) $$ \begin{align} f\blacktriangleleft h=\pi_{c,\mathfrak{t}}^{X-\text{loc}}(\delta(h))(f)\qquad \forall\, f\in\overline{\mathcal{Q}}^{(c)},\, \forall\, h\in\overline{H}. \end{align} $$

Proof. Let $c\in C^J$ and $\mathfrak {t}\in T_J$ . For $i\in [1,r]$ , $f\in \overline {\mathcal {Q}}$ and $y\in \mathcal {O}_c$ , we have in $\overline {\mathcal {Q}}^{(c)}$ ,

(6.43) $$ \begin{align} \begin{aligned} \pi_{c,\mathfrak{t}}^{X-\text{loc}}(\delta(T_i))(fx^y)&=\pi_{c,\mathfrak{t}}^{X-\text{loc}}(T_i)(fx^y)\\ &=k_i^{\chi_{\mathbb{Z}}(\alpha_i(y))}s_i(fx^{y})+(k_i-k_i^{-1})\left(\frac{fx^y-s_i(f)x^{y-\lfloor \alpha_i(y)\rfloor\alpha_i^{\vee}}}{1-x^{\alpha_i^{\vee}}}\right) \end{aligned} \end{align} $$

by (4.17). Furthermore, we have by (2.47),

(6.44) $$ \begin{align} \pi_{c,\mathfrak{t}}^{X-\text{loc}}(\delta(T_0))(fx^y)=\pi_{c,\mathfrak{t}}^{X-\text{loc}}(T_{s_\varphi}^{-1}x^{-\varphi^{\vee}})(fx^y)=\pi_{c,\mathfrak{t}}^{X-\text{loc}}(\delta(T_{s_\varphi}^{-1}))(fx^{y-\varphi^{\vee}}). \end{align} $$

This shows that $\pi _{c,\mathfrak {t}}^{X-\text {loc}}\vert _{\delta (H)}$ is independent of $\mathfrak {t}\in T_J$ . Hence, (6.42) defines a $\mathfrak {t}$ -independent $\mathbf {K}$ -linear right $\overline {H}$ -action $\blacktriangleleft $ on $\overline {\mathcal {Q}}^{(c)}$ . By the explicit formulas (6.43) and (6.44), the $\overline {H}$ -action $\blacktriangleleft $ on $\overline {\mathcal {Q}}^{(c)}$ is characterised by the formulas (6.40) and (6.41). The remaining statements now follow immediately.

The adjoint torus over $\mathbf {K}$ is denoted by

$$\begin{align*}\overline{T}:=\text{Hom}(Q^{\vee},\mathbf{K})\subset T=\text{Hom}(Q^{\vee},\mathbf{F}). \end{align*}$$

Let $J\subsetneq [0,r]$ and set

$$\begin{align*}\overline{T}_J^{\text{red}}:=\{\mathfrak{t}\in\overline{T} \,\, | \,\, \mathfrak{t}^{D\alpha_j^{\vee}}=1\quad \forall\, j\in J \}. \end{align*}$$

Choose $\lambda _J\in C^J$ such that $\lambda _J\in \frac {1}{2h}P^{\vee }$ (see Subsection 6.1). Recall that $q^{\lambda _J}\in T_J$ is the character which takes the value

$$\begin{align*}q^{\langle \lambda_J,\alpha^{\vee}\rangle}=q_\alpha^{\alpha(\lambda_J)}\in\mathbb{Q}\big(q^{\frac{1}{2h}}\big)\subset\mathbf{F} \end{align*}$$

at the coroot $\alpha ^{\vee }\in \Phi _0^{\vee }$ (see Subsection 6.1). Note that

(6.45) $$ \begin{align} q^{\lambda_J}\overline{T}_J^{\text{red}}\subseteq\{\mathfrak{t}\in T_J^\prime \,\, | \,\, \mathfrak{t}\, \mbox{ is }\, J\text{-generic}\}. \end{align} $$

Let $\mathbf {F}_{\text {reg}}\subset \mathbf {F}$ be the subring consisting of the elements $d\in \mathbf {F}=\mathbf {K}(q^{\frac {1}{2h}})$ which are regular at $q^{\,-\frac {1}{2h}}=0$ . Specialisation at $q^{-\frac {1}{2h}}=0$ defines a ring homomorphism $\mathbf {F}_{\text {reg}}\rightarrow \mathbf {K}$ , $d\mapsto \overline {d}$ which we extend to a ring homomorphism

$$\begin{align*}\mathbf{F}_{\text{reg}}[E]\rightarrow\mathbf{K}[E],\qquad f\mapsto\overline{f}:=\sum_{y\in E}\overline{d}_yx^y\quad (f=\sum_{y\in E}d_yx^y\in\mathbf{F}_{\text{reg}}[E]). \end{align*}$$

We call $\overline {f}\in \mathbf {K}[E]$ the Whittaker limit of $f\in \mathbf {F}_{\text {reg}}[E]$ . Write

$$\begin{align*}\mathcal{P}_{\text{reg}}^{(c)}:=\bigoplus_{y\in\mathcal{O}_c}\mathbf{F}_{\text{reg}}x^y. \end{align*}$$

Proposition 6.47. Let $c\in C^J$ , $\mathfrak {t}\in \overline {T}_J^{\text {red}}$ and $y\in \mathcal {O}_c$ . Then

  1. 1. $E_y^J(x;q^{\lambda _J}\mathfrak {t})\in \mathcal {P}_{\text {reg}}^{(c)}$ .

  2. 2. The Whittaker limit $\overline {E}_y^J(x)\in \overline {\mathcal {P}}^{(c)}$ of $E_y^J(x;q^{\lambda _J}\mathfrak {t})$ is independent of $\lambda _J$ and $\mathfrak {t}$ .

  3. 3. For $j\in [0,r]$ , we have

    (6.46) $$ \begin{align} \overline{E}_{s_jy}^J(x)\blacktriangleleft T_j= \begin{cases} \kappa_{Ds_j}(y)^{-1}\overline{E}_y^J(x)\quad &\mbox{ if }\, \alpha_j(y)<0,\\ k_j\overline{E}_y^J(x)\quad &\mbox{ if }\, \alpha_j(y)=0. \end{cases} \end{align} $$

Remark 6.48. Note that $\kappa _{Ds_j}(y)\not =k_j^{-1}$ when $\alpha _j(y)=0$ , so it is necessary to distinguish the two cases in part (3) of the proposition (an explicit formula for $\overline {E}_{s_jy}^J(x)\blacktriangleleft T_j$ when $\alpha _j(y)>0$ follows from (6.46) and the Hecke relation (2.39) for $T_j$ ).

Proof. For $y\in \mathcal {O}_c$ , write $t_y:=w_y(q^{\lambda _J}\mathfrak {s}_J\mathfrak {t})\in T$ (see (6.45) and (6.16)). Then

(6.47) $$ \begin{align} t_y^{\alpha_j^{\vee}}=q_{\alpha_j}^{\alpha_j(w_y\lambda_J)}(\mathfrak{s}_y\mathfrak{t}_y)^{D\alpha_j^{\vee}} \end{align} $$

with $q_{\alpha _j}^{\alpha _j(w_y\lambda _j)}\in \mathbb {Q}\big (q^{\frac {1}{2h}}\big )$ and $(\mathfrak {s}_y\mathfrak {t}_y)^{D\alpha _j^{\vee }}\in \mathbf {K}$ , and Corollary 6.30 gives

(6.48) $$ \begin{align} \begin{aligned} \pi_{c,q^{\lambda_J}\mathfrak{t}}(\delta(T_j))E_{s_jy}^J(x;q^{\lambda_J}\mathfrak{t})+ \ &\left(\frac{k_j-k_j^{-1}}{t_{y}^{-\alpha_j^{\vee}}-1}\right)E_{s_jy}^J(x;q^{\lambda_J}\mathfrak{t})\\ &\quad= \kappa_{Ds_j}(y)^{-1}E_y^J(x;q^{\lambda_J}\mathfrak{t})\quad \mbox{ if }\, \alpha_j(y)<0 \end{aligned} \end{align} $$

(we used here that $\kappa _{Ds_i}(s_iy)=\kappa _{Ds_i}(y)^{-1}$ by (5.8), and $t_{s_jy}^{\alpha _j^{\vee }}=t_y^{-\alpha _j^{\vee }}$ ). We now use (6.48) to prove (1) and (2) by induction to $\ell (w_y)$ .

If $\ell (w_y)=0$ , then $y=c$ and $E_c^J(x;q^{\lambda _J}\mathfrak {t})=x^c\in \mathcal {P}_{\text {reg}}^{(c)}$ . Furthermore, $\overline {E}_c^J(x)=x^c$ .

Let $y\in \mathcal {O}_c$ with $\ell (w_y)>0$ and suppose that for all $y^\prime \in \mathcal {O}_c$ with $\ell (w_{y^\prime })<\ell (w_y)$ , we have $E_{y^\prime }^J(x;q^{\lambda _J}\mathfrak {t})\in \mathcal {P}_{\text {reg}}^{(c)}$ with $\overline {E}_{y^\prime }^J(x)$ independent of $\lambda _J$ and $\mathfrak {t}$ . Take $0\leq j\leq r$ such that $\ell (s_jw_y)=\ell (w_y)-1$ . By Corollary 5.23, we have $\alpha _j(y)<0$ and $s_jw_y=w_{s_jy}$ – in particular, $\ell (w_{s_jy})<\ell (w_y)$ . Since both c and $\lambda _J$ lie in $C^J$ and $\alpha _j(y)=\alpha _j(w_yc)<0$ , we also have $\alpha _j(w_y\lambda _J)<0$ by Corollary 5.23(1). Taking also (6.47) into account, we conclude that the left-hand side of (6.48) lies in $\mathcal {P}_{\text {reg}}^{(c)}$ . Hence, $E_y^J(x;q^{\lambda _J}\mathfrak {t})\in \mathcal {P}_{\text {reg}}^{(c)}$ , and the Whittaker limit of (6.48) gives

$$\begin{align*}\overline{E}_{y}^J(x)=\kappa_{Ds_j}(y)\overline{E}_{s_jy}^J(x)\blacktriangleleft T_j. \end{align*}$$

The right-hand side does not depend on $\lambda _J$ and $\mathfrak {t}$ by the induction hypothesis. This completes the proof of the induction step.

The Whittaker limit of (6.48) yields the first case of (6.46). The second case is immediate from Corollary 6.30.

Corollary 6.49. Let $c\in C^J$ and $y\in \mathcal {O}_c$ . Then

$$ \begin{align*} \overline{E}_y^J(x)=\frac{k(c)}{k(y)}\,x^c\blacktriangleleft T_{w_y^{-1}}. \end{align*} $$

Proof. This follows from Proposition 6.47, Corollary 5.23 and (5.8) by induction to $\ell (w_y)$ .

In Theorem 6.51, we will give an explicit formula for $\overline {E}_{y}^J(x)$ only involving the $\overline {H}_0$ -action on $\mathbf {K}[E]$ . We need for this the following preparatory lemma.

Lemma 6.50. If $c\in \overline {C}_+$ and $y \in \mathcal {O}_{c}\cap \overline {E}_-$ , then $y = vc + \mu $ for some $v \in W_0$ and $\mu \in Q^{\vee }\cap \overline {E}_-$ .

Proof. We may assume without loss of generality that $E=E^\prime $ (i.e., $\Phi _0$ spans E).

First, suppose that $y \in \mathcal {O}_c\cap E_-$ . Write $y=vc+\mu $ with $v\in W_0$ and $\mu \in Q^{\vee }$ . For $\alpha \in \Phi _0^{+}$ , we have $\alpha (\mu ) = \alpha (y)- \alpha (vc) < 1$ , since $\alpha (y)<0$ and $|\alpha (vc)| \leq 1$ . Since $\mu \in Q^{\vee }$ , it follows that $\alpha (\mu ) \leq 0$ , so $\mu \in Q^{\vee }\cap \overline {E}_-$ .

The general case follows by continuity. Indeed, let $y \in \mathcal {O}_{c}\cap \overline {E}_-$ . Let $\{y_\ell \}_{\ell =1}^{\infty }$ be a sequence in $E_-$ converging to y. Denote $c_{y_\ell }\in \overline {C}_+$ by $c_\ell $ . By the first paragraph, we have $y_\ell = v_\ell c_\ell + \mu _\ell $ for some $v_\ell \in W_0$ and $\mu _\ell \in Q^{\vee }\cap \overline {E}_-$ . Since $W_0$ is finite, we may assume by passing to a subsequence that $v_\ell = v\in W_0$ is constant. Since $\overline {C}_{+}$ is compact (by the assumption that $E=E^\prime $ ), we may assume by passing to a subsequence that $c_\ell $ converges to some $c^\prime \in \overline {C}_{+}$ . Since $Q^{\vee }\cap \overline {E}_-$ is discrete, it then follows that $\mu _\ell =\mu $ is constant for $\ell $ sufficiently large. We then have $y = vc^\prime + \mu $ , so $c^\prime = c$ . The result now follows, since $\mu \in Q^{\vee }\cap \overline {E}_-$ .

Recall that $g_y\in W_0$ denotes the unique element of minimal length such that $y_-=g_yy$ .

Theorem 6.51. For $c\in C^J$ and $y\in \mathcal {O}_c$ , we have

(6.49) $$ \begin{align} \overline{E}_y^J(x)=\frac{k(y_-)}{k(y)}\,x^{y_-}\blacktriangleleft T_{g_y^{-1}}^{-1}. \end{align} $$

Proof. First, we prove the theorem for $y\in \mathcal {O}_c\cap \overline {E}_-$ .

In this case, the right-hand side of (6.49) is simply $x^y$ . Since $\overline {E}_y^J(x)=x^y+\text {l.o.t.}$ it suffices to show that $\overline {E}_y^J(x)\sim x^y$ , where we write $p\sim p^\prime $ if p and $p^\prime $ in $\overline {\mathcal {P}}^{(c)}$ differ by a nonzero constant multiple.

Write $y\in \mathcal {O}_c\cap \overline {E}_-$ as $y=vc+\mu $ with $v\in W_0$ and $\mu \in Q^{\vee }\cap \overline {E}_-$ ; see Lemma 6.50. Set $w:=\tau (\mu )v\in W$ . Then $w=w_yh_y$ for a unique $h_y\in W_J$ and

$$\begin{align*}\ell(v)+\ell(\tau(\mu))=\ell(w)=\ell(w_y)+\ell(h_y), \end{align*}$$

where the first equality is due to (2.13). It follows that

(6.50) $$ \begin{align} \delta(T_{w^{-1}})=\delta(T_{v^{-1}}Y^{-\mu})=x^\mu T_v \end{align} $$

and $T_{w_y^{-1}}=T_{h_y^{-1}}^{-1}T_{w^{-1}}$ . Since $x^c\blacktriangleleft T_{h_y^{-1}}^{-1}\sim x^c$ by Proposition 6.47(3), we get from Corollary 6.49 that

$$ \begin{align*} \overline{E}_y^J(x)\sim x^c\blacktriangleleft T_{w_y^{-1}}\sim x^c\blacktriangleleft T_{w^{-1}}. \end{align*} $$

By Lemma 6.46 and (6.50), we then have for $\mathfrak {t}\in T_J$ ,

$$ \begin{align*} \overline{E}_y^J(x)\sim\pi_{c,\mathfrak{t}}(\delta(T_{w^{-1}}))x^c=x^\mu\pi_{c,\mathfrak{t}}(T_v)x^c\sim x^{\mu+vc}=x^y, \end{align*} $$

where the final step is due to Corollary 5.10. Hence, $\overline {E}_y^J(x)=x^y$ , which completes the proof of (6.49) when $y\in \mathcal {O}_c\cap \overline {E}_-$ .

Suppose now that $y\in \mathcal {O}_c$ is arbitrary. Take a reduced expression $g_y^{-1}=s_{i_1}\cdots s_{i_\ell }$ . Since $g_y^{-1}\in W_0^{y_-}$ , formula (2.12) implies that

$$\begin{align*}\alpha_{i_m}(s_{i_{m+1}}\cdots s_{i_\ell}y_-)<0 \end{align*}$$

for $1\leq m\leq \ell $ . Hence, by (5.8) and by repeated application of Proposition 6.47(3),

$$\begin{align*}\overline{E}^J_y(x)=\overline{E}^J_{g_y^{-1}y_-}(x)= \frac{k(y_-)}{k(y)}\overline{E}^J_{y_-}(x)\blacktriangleleft T_{g_y^{-1}}^{-1}. \end{align*}$$

The result now follows since $\overline {E}_{y^-}^J(x)=x^{y_-}$ by the previous paragraph.

Let $c\in C^J$ , $\mathfrak {t}\in \overline {T}^{\text {red}}_J$ and $y\in \mathcal {O}_c$ . By Lemma 6.46 and Proposition 6.47, both $E_y^{J,+}(x;q^{\lambda _J}\mathfrak {t})$ and $E_y^{J,-}(x;q^{\lambda _J}\mathfrak {t})$ lie in $\mathbf {F}_{\text {reg}}[E]$ , and their Whittaker limits $\overline {E}_y^{J,+}(x)$ and $\overline {E}_y^{J,-}(x)$ do not depend on $\lambda _J$ and $\mathfrak {t}$ . We now take a close look at $\overline {E}_y^{J,-}(x)$ , since they are closely related to spherical metaplectic Whittaker functions (see Section 10). We list the main formulas for $\overline {E}_y^{J,+}(x)$ at the end of the subsection.

First, note that

$$\begin{align*}\overline{E}_y^{J,-}(x)=0\quad \mbox{ if }\, y\not\in E^{\text{reg}}\cap\mathcal{O}_c \end{align*}$$

by (6.36). Furthermore, $\overline {E}_{vy}^{J,-}(x)$ ( $v\in W_0$ ) is a nonzero constant multiple of $\overline {E}_y^{J,-}(x)$ when $y\in E^{\text {reg}}\cap \mathcal {O}_c$ .

Corollary 6.52. Suppose $c\in C^J$ and $y\in E_-\cap \mathcal {O}_c$ . Then

(6.51) $$ \begin{align} \overline{E}_y^{J,-}(x)=\kappa_{w_0}(0)^{-2}\sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v(0)\kappa_v(y)\overline{E}_{vy}^J(x). \end{align} $$

Furthermore,

(6.52) $$ \begin{align} \kappa_v(y)\overline{E}_{vy}^J(x)=x^y\blacktriangleleft T_v^{-1}\qquad\quad(v\in W_0). \end{align} $$

Proof. We first prove (6.52). For $v\in W_0$ , we have $g_{vy}=v^{-1}$ since $y\in E_-$ , and hence by (6.49),

$$\begin{align*}\overline{E}_{vy}^J(x)=\frac{k(y)}{k(vy)}x^y\blacktriangleleft T_v^{-1}. \end{align*}$$

Then use that $k(vy)/k(y)=k_v(y)=\kappa _v(y)$ , since $y\in E^{\text {reg}}$ .

There are two ways one can arrive at (6.51). First, one can take the Whittaker limit of (6.34) after fixing $\mathfrak {t}:=q^{\lambda _J}\mathfrak {t}^\prime \in T_J^\prime $ with $\mathfrak {t}^\prime \in \overline {T}_J^{\text {red}}$ . By Lemma 6.46, Proposition 6.47 and (6.49), it then follows that

(6.53) $$ \begin{align} \overline{E}_y^{J,-}(x)=\kappa_{w_0}(0)^{-2}\sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v(0)\,x^y\blacktriangleleft T_v^{-1}, \end{align} $$

and (6.51) follows from (6.52). Second, one can take the Whittaker limit of the formula in Proposition 6.41 after fixing $\mathfrak {t}:=q^{\lambda _J}\mathfrak {t}^\prime \in T_J^\prime $ with $\mathfrak {t}^\prime \in \overline {T}_J^{\text {red}}$ . To perform the specialisation in this case, one notes that for $\beta \in \Phi _0^+$ ,

$$\begin{align*}t_y^{\beta^{\vee}}=q_\beta^{\beta(w_y\lambda_J)}\big(\mathfrak{s}_y\mathfrak{t}_y\big)^{\beta^{\vee}} \end{align*}$$

and the q-power is strictly negative, since $y=w_yc\in E_-$ implies $w_y\lambda _J\in E_-$ . The Whittaker limit of the formula in Proposition 6.41 then gives (6.51) (compare with the proof of Corollary 6.38).

Lemma 6.53. The formulas

$$ \begin{align*} \begin{aligned} s_i\blacktriangleright x^y&:=\frac{k_i^{\chi_{\mathbb{Z}}(\alpha_i(y))}(x^{\alpha_i^{\vee}}-1)}{(k_ix^{\alpha_i^{\vee}}-k_i^{-1})}x^{s_iy}+\frac{(k_i-k_i^{-1})}{(k_ix^{\alpha_i^{\vee}}-k_i^{-1})} x^{y-\lfloor\alpha_i(y)\rfloor\alpha_i^{\vee}},\\ f\blacktriangleright x^y&:=fx^y \end{aligned} \end{align*} $$

for $i\in [1,r]$ , $f\in \overline {\mathcal {Q}}$ and $y\in \mathcal {O}_c$ define a left $W_0\ltimes \overline {\mathcal {Q}}$ -action on $\mathbf {K}_{\overline {\mathcal {Q}}}[E]$ . Furthermore, if $c\in C^J$ , then $\sigma _{c,\mathfrak {t}}\vert _{W_0\ltimes \overline {\mathcal {Q}}}$ does not depend on $\mathfrak {t}\in T_J$ , and

(6.54) $$ \begin{align} v\blacktriangleright (fx^y)=\sigma_{c,\mathfrak{t}}(v)(fx^y)\qquad\quad (v\in W_0,\, f\in\overline{\mathcal{Q}},\, y\in\mathcal{O}_c). \end{align} $$

Proof. It suffices to note that (6.54) is correct for $v=s_i$ ( $1\leq i\leq r$ ) and $y\in \mathcal {O}_c$ . This follows immediately from the definition of $\sigma _{c,\mathfrak {t}}$ ; see Theorem 4.9.

In the following lemma, we describe the right $\overline {H}_0$ -action on $\mathbf {K}_{\overline {\mathcal {Q}}}[E]$ in terms of Demazure-Lusztig type operators.

Lemma 6.54. For $i\in [1,r]$ and $f\in \mathbf {K}_{\overline {\mathcal {Q}}}[E]$ , we have

(6.55) $$ \begin{align} f\blacktriangleleft T_i=k_if+k_i^{-1}\left(\frac{1-k_i^2x^{\alpha_i^{\vee}}}{1-x^{\alpha_i^{\vee}}}\right)\big(s_i\blacktriangleright f-f\big). \end{align} $$

Proof. This follows from (6.42), (6.54) and Remark 4.11.

Proposition 6.55. Let $c\in C^J$ and $y\in \mathcal {O}_c$ . Then

$$\begin{align*}\overline{E}_y^{J,-}(x)=\left(\prod_{\alpha\in\Phi_0^+}\frac{1-k_\alpha^{-2}x^{-\alpha^{\vee}}}{1-x^{-\alpha^{\vee}}}\right)\sum_{v\in W_0}(-1)^{\ell(v)}v\blacktriangleright\overline{E}_y^J(x). \end{align*}$$

In particular,

(6.56) $$ \begin{align} \overline{E}_y^{J,-}(x)=\left(\prod_{\alpha\in\Phi_0^+}\frac{1-k_\alpha^{-2}x^{-\alpha^{\vee}}}{1-x^{-\alpha^{\vee}}}\right)\sum_{v\in W_0}(-1)^{\ell(v)}v\blacktriangleright x^y\qquad (y\in E_-\cap\mathcal{O}_c). \end{align} $$

Proof. The first formula follows by taking the Whittaker limit of (6.35) after fixing $\mathfrak {t}=q^{\lambda _J}\mathfrak {t}^\prime \in T_J^\prime $ with $\mathfrak {t}^\prime \in \overline {T}_J^{\text {red}}$ . Then (6.56) follows from (6.49).

We now give the main formulas for $\overline {E}_y^{J,+}(x)$ . Let $c\in C^J$ and $y\in \mathcal {O}_c$ . Then $E_{vy}^{J,+}(x)$ ( $v\in W_0$ ) is a nonzero scalar multiple of $E_y^{J,+}(x)$ , and

$$\begin{align*}\overline{E}_y^{J,+}(x)=\kappa_{w_0}(0)^2\sum_{v\in W_0}\kappa_v(0)^{-1}\overline{E}_y^J(x)\blacktriangleleft T_v^{-1}. \end{align*}$$

In particular, for $y\in \overline {E}_-\cap \mathcal {O}_c$ , it follows from (6.46) and (6.52) that

(6.57) $$ \begin{align} \begin{aligned} \overline{E}_y^{J,+}(x)&=\kappa_{w_0}(0)^2\left(\sum_{v\in W_{0,y}}\kappa_v(0)^{-2}\right)\sum_{v\in W_0^y}\kappa_v(0)^{-1}x^y\blacktriangleleft T_v^{-1}\\ &=\kappa_{w_0}(0)^2\left(\sum_{v\in W_{0,y}}\kappa_v(0)^{-2}\right)\sum_{v\in W_0^y}\frac{\kappa_v(y)}{\kappa_v(0)}\overline{E}_{vy}^J(x). \end{aligned} \end{align} $$

The latter formula can also be directly obtained by taking the Whittaker limit of the formula in Corollary 6.38 after taking $\mathfrak {t}=q^{\lambda _J}\mathfrak {t}^\prime $ with $\mathfrak {t}^\prime \in \overline {T}^{\text {red}}_J$ . Then one uses the fact that $\beta (w_y\lambda _J)<0$ for $\beta \in \Pi (v)$ ( $v\in W_0^y$ ) since $\beta (y)<0$ , and that $\sum _{v\in W_{0,y}}\kappa _{g_{y_+}^{-1}v}(0)^2=\kappa _{w_0}(0)^2\sum _{v\in W_{0,y}}\kappa _v(0)^{-2}$ .

Taking the Whittaker limit of the formula in Proposition 6.39, we obtain the expression

$$\begin{align*}\overline{E}_y^{J,+}(x)=\left(\prod_{\alpha\in\Phi_0^+}\frac{1}{1-x^{-\alpha^{\vee}}}\right)\sum_{v\in W_0}\kappa_v(0)^2\left(\prod_{\alpha\in\Phi_0^+}\big(1-k_\alpha^{2\chi(v^{-1}\alpha)}x^{-\alpha^{\vee}}\big)\right)\, v\blacktriangleright\overline{E}_y^J(x). \end{align*}$$

In particular, for $y\in \overline {E}_-\cap \mathcal {O}_c$ ,

(6.58) $$ \begin{align} \begin{aligned} \overline{E}_y^{J,+}(x)&=\left(\prod_{\alpha\in\Phi_0^+}\frac{1}{1-x^{-\alpha^{\vee}}}\right) \sum_{v\in W_0}\kappa_v(0)^2\left(\prod_{\alpha\in\Phi_0^+}\big(1-k_\alpha^{2\chi(v^{-1}\alpha)}x^{-\alpha^{\vee}}\big)\right)\,v\blacktriangleright x^y\\ &=\sum_{v\in W_0}\left(\prod_{\alpha\in\Phi_0^+}\left(\frac{1-k_\alpha^2x^{-v\alpha^{\vee}}}{1-x^{-v\alpha^{\vee}}}\right)\right)\,v\blacktriangleright x^y; \end{aligned} \end{align} $$

cf. the proof of Proposition 6.39 for the second equality.

Remark 6.56. For $\mu \in P^{\vee }$ and $v\in W_0$ , we have $v\blacktriangleright x^\mu =\sigma _{c,\mathfrak {t}}(v)x^\mu =x^{v\mu }$ by (6.54) and Remark 4.10(1). Hence, the right-hand side of (6.58) for $y=\mu \in P^{\vee ,-}\cap \mathcal {O}_c$ gives the familiar explicit expression of Macdonald’s spherical function on a $\mathfrak {p}$ -adic Lie group [Reference Macdonald26] (which is the Hall-Littlewood polynomial in type A). Since $E_\mu ^{[1,r],+}(x)$ ( $\mu \in \overline {E}_-\cap Q^{\vee }$ ) is a symmetric Macdonald polynomial by Remark 6.34, and the monic symmetric Macdonald polynomials are invariant under the simultaneous inversion of the parameters q and $k_\alpha $ ( $\alpha \in \Phi _0$ ), this is in agreement with the interpretation of Macdonald’s spherical function as the $q\rightarrow 0$ limit of the symmetric Macdonald polynomial, see [Reference Macdonald27, §10].

For $\mu \in P^{\vee ,-}$ , write

$$\begin{align*}s_\mu(x)=\left(\prod_{\alpha\in\Phi_0^+}\frac{1}{1-x^{-\alpha^{\vee}}}\right)\sum_{v\in W_0}(-1)^{\ell(v)}x^{-\rho^{\vee}+v(\mu+\rho^{\vee})}= \sum_{v\in W_0}\prod_{\alpha\in\Phi_0^+}\left(\frac{x^{v\mu}}{1-x^{-v\alpha^{\vee}}}\right), \end{align*}$$

which is Weyl’s formula for the character of the associated finite dimensional irreducible representation of the complex simply connected semisimple Lie group with underlying root system $\Phi _0^{\vee }$ (the highest weight $\mu $ is relative to $\Phi _0^{\vee ,-}$ ). It is the Schur function for type A.

Recall that $\xi _y$ is the unique element in $y+P^{\vee }$ such that $0\leq \alpha _i(\xi _y)<1$ for $i\in [1,r]$ (see Remark 4.12).

Proposition 6.57. Let $c\in C^J$ and $y\in \overline {E}_-\cap \mathcal {O}_c$ . For the quasi-polynomial $\overline {E}_y^{J,+}(x)=\overline {E}_y^{J,+}(x;\mathbf {k})$ in the limiting case $\mathbf {k}=\mathbf {0}$ , we have

$$\begin{align*}\overline{E}_y^{J,+}(x;\mathbf{0})=x^{\xi_y}s_{y-\xi_y}(x). \end{align*}$$

Proof. Note that $y-\xi _y\in P^{\vee ,-}$ since $y\in \overline {E}_-$ . By (6.54) and Remark 4.12, we have for $v\in W_0$ ,

$$\begin{align*}v\blacktriangleright_{\mathbf{k}=\mathbf{0}}x^y=\sigma_{c,\mathfrak{t}}^{\mathbf{k}=\mathbf{0}}(v)x^y=x^{\xi_y+v(y-\xi_y)}, \end{align*}$$

and hence, the result follows from (6.58).

We will see in Section 10 that the $\overline {E}_y^J(x)$ and $\overline {E}_y^{J,-}(x)$ are, after some elementary modifications, the metaplectic Iwahori-Whittaker and metaplectic spherical Whittaker functions from [Reference Patnaik and Puskas34] and [Reference Chinta, Gunnells and Puskas15], respectively. Hence, $E_{y}^{J}(x;\mathfrak {t})$ and $E_{y}^{J,-}(x;\mathfrak {t})$ are q-analogues of metaplectic Whittaker functions, depending on the additional parameters $\mathfrak {t}\in T_J$ . The results in this subsection then relates to the metaplectic theory as follows:

  1. 1. Lemma 6.53 corresponds to Chinta’s and Gunnells’ [Reference Chinta and Gunnells13, Reference Chinta and Gunnells14] Weyl group action on rational functions, used to construct the local parts of Weyl group multiple Dirichlet series (see also [Reference Sahi, Stokman and Venkateswaran38, §3.3]).

  2. 2. Formula (6.52) corresponds via (6.55) with Patnaik’s and Puskas’ [Reference Patnaik and Puskas34, Cor. 5.4] expression of the metaplectic Iwahori-Whittaker function in terms of metaplectic Demazure-Lusztig operators.

  3. 3. Formula (6.56) corresponds to McNamara’s [Reference McNamara30, Thm. 15.2] Casselman-Shalika type formula for the metaplectic spherical Whittaker functions.

  4. 4. Formula (6.53) corresponds Chinta’s, Gunnells’ and Puskas’ [Reference Chinta, Gunnells and Puskas15, Thm. 16] expression of the metaplectic spherical Whittaker function in terms of metaplectic Demazure-Lusztig operators.

In Section 10, we give further details about these connections with metaplectic representation theory and metaplectic Whittaker functions.

7 Quasi-polynomial representations of extended double affine Hecke algebras

In this section, we generalise the results of the previous sections to extended root data. We assume throughout this section that J is a proper subset of $[0,r]$ .

7.1 Extended root datum and extended affine Weyl groups

For $e\in \mathbb {Z}_{>0}$ , let $(\mathcal {L}^{\times 2})_e$ be the set of pairs $(\Lambda ,\Lambda ^\prime )$ with $\Lambda ,\Lambda ^\prime \in \mathcal {L}$ (see (2.27)) satisfying the additional requirement that

(7.1) $$ \begin{align} \langle \Lambda,\Lambda^\prime\rangle\subseteq e^{-1}\mathbb{Z}. \end{align} $$

Our choice of normalisation of the root system $\Phi _0$ implies that $Q^{\vee }\subseteq Q$ ; hence, $(Q^{\vee },\Lambda ^\prime )\in (\mathcal {L}^{\times 2})_e$ for all $\Lambda ^\prime \in \mathcal {L}$ .

Furthermore, if either $\Lambda $ or $\Lambda ^\prime $ is contained in $E^\prime $ , then $(\Lambda ,\Lambda ^\prime )\in (\mathcal {L}^{\times 2})_{e}$ when $h\vert e$ . Indeed, if either $\Lambda $ or $\Lambda ^\prime $ is contained in $E^\prime $ , then $\langle \Lambda ,\Lambda ^\prime \rangle \subseteq \langle P^{\vee },P^{\vee }\rangle $ , and $\langle P^{\vee },P^{\vee }\rangle \subseteq h^{-1}\mathbb {Z}$ since $Q^{\vee }\subseteq Q$ and $P^{\vee }\subseteq \frac {1}{h}Q^{\vee }$ (see Proposition 2.10).

Note that if $(\Lambda _2,\Lambda _2^\prime )\in (\mathcal {L}^{\times 2})_e$ and if $\Lambda _1,\Lambda _1^\prime \in \mathcal {L}$ are such that $\Lambda _1\subseteq \Lambda _2$ and $\Lambda _1^\prime \subseteq \Lambda _2^\prime $ , then $(\Lambda _1,\Lambda _1^\prime )\in (\mathcal {L}^{\times 2})_e$ .

Definition 7.1. For $(\Lambda _1,\Lambda _1^\prime ),(\Lambda _2,\Lambda _2^\prime )\in (\mathcal {L}^{\times 2})_e$ , we write

$$\begin{align*}(\Lambda_1,\Lambda_1^\prime)\leq (\Lambda_2,\Lambda_2^\prime) \end{align*}$$

if $\Lambda _1\subseteq \Lambda _2$ and $\Lambda _1^\prime \subseteq \Lambda _2^\prime $ .

Definition 7.1 turns $(\mathcal {L}^{\times 2})_e$ into a partially ordered set with least element $(Q^{\vee },Q^{\vee })$ .

The lattice $\Lambda ^\prime \subset E$ is $W_0$ -invariant. The resulting group

$$\begin{align*}W_{\Lambda^\prime}:=W_0\ltimes \Lambda^\prime \end{align*}$$

is called the extended affine Weyl group. It contains W as normal subgroup, and $W_{\Lambda ^\prime }/W\simeq \Lambda ^\prime /Q^{\vee }$ . We denote its elements by $v\tau (\lambda )$ ( $v\in W_0$ , $\lambda \in \Lambda ^\prime $ ), and write $(w,y)\mapsto wy$ ( $w\in W_{\Lambda ^\prime }$ , $y\in E$ ) for the $W_{\Lambda ^\prime }$ -action on E by reflections and translations. Extend the group epimorphism $D: W\twoheadrightarrow W_0$ to a group epimorphism $D: W_{\Lambda ^\prime }\twoheadrightarrow W_0$ by $D(v\tau (\lambda )):=v$ for $v\in W_0$ and $\lambda \in \Lambda ^\prime $ .

The contragredient $W_{\Lambda ^\prime }$ -action provides a linear action on the space $E^*\oplus \mathbb {R}$ of affine linear functionals on E. The action is explicitly given by (2.3). The condition that $\alpha (\Lambda ^\prime )\subseteq \mathbb {Z}$ for all $\alpha \in \Phi _0$ implies that the affine root system $\Phi \subset E^*\oplus \mathbb {R}$ is $W_{\Lambda ^\prime }$ -stable. We extend the length function $\ell : W\rightarrow \mathbb {Z}_{\geq 0}$ to $W_{\Lambda ^\prime }$ using formula (2.11). It satisfies (2.13) for $\mu \in \Lambda ^\prime $ and $v\in W_0$ .

In Definition 2.3, we introduced, for $y\in E$ , the element $w_y\in W$ as the unique element of minimal length such that $w_yc_y=y$ , as well as the finite Weyl group element $v_y:=w_y^{-1}\tau (\mu _y)\in W_0$ , where $\mu _y=w_y(0)$ . For $y=\lambda \in Q^{\vee }$ one has $\mu _\lambda =\lambda $ ; hence,

(7.2) $$ \begin{align} \tau(\lambda)=w_\lambda v_\lambda \qquad\quad (\lambda\in Q^{\vee}). \end{align} $$

We now generalise the factorisation (7.2) to $\lambda \in \Lambda ^\prime $ as follows.

Let $\lambda \in \Lambda ^\prime $ . By, for example, [Reference Macdonald29, (2.4.5)], there exists a unique $w_{\lambda ,\Lambda ^\prime }\in \tau (\lambda )W_0\subset W_{\Lambda ^\prime }$ of minimal length. The resulting finite Weyl group element

$$\begin{align*}v_{\lambda,\Lambda^\prime}:=w_{\lambda,\Lambda^\prime}^{-1}\tau(\lambda)\in W_0 \qquad (\lambda\in\Lambda^\prime) \end{align*}$$

is the shortest element in $W_0$ such that $v_{\lambda ,\Lambda ^\prime }\lambda =\lambda _-$ . Furthermore, $\ell (w_{\lambda ,\Lambda ^\prime }v)=\ell (w_{\lambda ,\Lambda ^\prime })+\ell (v)$ for $v\in W_0$ (for details see, for example, [Reference Macdonald29, §2.4]). We thus have the decomposition

$$\begin{align*}\tau(\lambda)=w_{\lambda,\Lambda^\prime}v_{\lambda,\Lambda^\prime}\qquad\quad (\lambda\in\Lambda^\prime) \end{align*}$$

with $\ell (\tau (\lambda ))= \ell (w_{\lambda ,\Lambda ^\prime })+\ell (v_{\lambda ,\Lambda ^\prime })$ . Note that $w_{\lambda ,\Lambda ^\prime }\not =w_\lambda $ for $\lambda \in \Lambda ^\prime \setminus Q^{\vee }$ , but we do have

(7.3) $$ \begin{align} w_{\lambda,\Lambda^\prime}=w_\lambda,\qquad\quad v_{\lambda,\Lambda^\prime}=v_\lambda\qquad (\lambda\in Q^{\vee}). \end{align} $$

To compare $w_{\lambda ,\Lambda ^\prime }$ and $w_\lambda $ for $\lambda \in \Lambda ^\prime \setminus Q^{\vee }$ , we first need to describe the elements of the subgroup

$$\begin{align*}\Omega_{\Lambda^\prime}:=\{\omega\in W_{\Lambda^\prime} \,\, | \,\, \ell(\omega)=0\} \end{align*}$$

of $W_{\Lambda ^\prime }$ in detail. It stabilises $C_+$ and $\Delta $ . For $j\in [0,r]$ and $\omega \in \Omega _{\Lambda ^\prime }$ , we write $\omega (j)\in [0,r]$ for the index such that $\omega (\alpha _j)=\alpha _{\omega (j)}$ . Then we have $\omega s_j\omega ^{-1}=s_{\omega (j)}$ .

We now recall some well-known facts about the abelian quotient group $\Lambda ^\prime /Q^{\vee }$ and $\Omega _{\Lambda ^\prime }$ (see, for example, [Reference Macdonald29, §2.5] for details). First, note that $\mu +Q^{\vee }=W\mu $ for all $\mu \in \Lambda ^\prime $ , since $\Lambda ^\prime \in \mathcal {L}$ . Hence, the set

$$\begin{align*}\Lambda_{\text{min}}^\prime:=\overline{C}_+\cap \Lambda^\prime=\{\lambda\in\Lambda^\prime \,\,\, | \,\,\, 0\leq \alpha(\lambda)\leq 1\quad \forall\, \alpha\in\Phi_0^+\} \end{align*}$$

of minuscule weights in $\Lambda ^\prime $ forms a complete set of representatives of $\Lambda ^\prime /Q^{\vee }$ . The inverse of the set-theoretic bijection $\Lambda _{\text {min}}^\prime \overset {\sim }{\longrightarrow }\Lambda ^\prime /Q^{\vee }$ , $\zeta \mapsto \zeta +Q^{\vee }$ is concretely given by $\mu +Q^{\vee }\mapsto c_\mu $ for $\mu \in \Lambda ^\prime $ . The following result is well known (see, for example, [Reference Macdonald29, §2.5]).

Proposition 7.2.

  1. 1. The assignment $\omega \mapsto \omega (0)+Q^{\vee }$ defines a group isomorphism $\Omega _{\Lambda ^\prime }\overset {\sim }{\longrightarrow }\Lambda ^\prime /Q^{\vee }$ . Its inverse is explicitly given by $\mu +Q^{\vee }\mapsto w_{c_\mu ,\Lambda ^\prime }$ for $\mu \in \Lambda ^\prime $ .

  2. 2. $W_{\Lambda ^\prime }=\Omega _{\Lambda ^\prime }\ltimes W$ .

A reduced expression of $w\in W_{\Lambda ^\prime }$ is an expression $w=\omega s_{j_1}\cdots s_{j_\ell }$ with $\omega \in \Omega _{\Lambda ^\prime }$ , $0\leq j_i\leq r$ and $\ell =\ell (w)$ . The roots in $\Pi (w)$ are still described by (2.12).

The elements $v_{\zeta ,\Lambda ^\prime }\in W_0$ ( $\zeta \in \Lambda _{\text {min}}^\prime $ ) satisfy

(7.4) $$ \begin{align} \begin{aligned} \Pi(v_{\zeta,\Lambda^\prime})&=\{\alpha\in\Phi^+_0 \,\, | \,\, \alpha(\zeta)=1\},\\ \Phi^+_0\setminus\Pi(v_{\zeta,\Lambda^\prime})&=\{\alpha\in\Phi^+_0 \,\, | \,\, \alpha(\zeta)=0\}; \end{aligned} \end{align} $$

see [Reference Macdonald29, §2.5]. Furthermore, for $\zeta \in \Lambda _{\text {min}}^\prime $ , we have

$$\begin{align*}\zeta^\prime:=-w_0\zeta=-v_{\zeta,\Lambda}\zeta\in \Lambda_{\text{min}}^\prime, \end{align*}$$

$w_{\zeta ^\prime ,\Lambda ^\prime }=w_{\zeta ,\Lambda ^\prime }^{-1}$ , $v_{\zeta ^\prime ,\Lambda ^\prime }=v_{\zeta ,\Lambda ^\prime }^{-1}$ (see [Reference Macdonald29, (2.5.9)]), and

$$\begin{align*}w_{\zeta,\Lambda^\prime}y =\zeta+v_{\zeta,\Lambda^\prime}^{-1}y\qquad (y\in E). \end{align*}$$

The minuscule weights admit the following explicit description. Recall the functions $n_i: \Phi _0\rightarrow \mathbb {Z}$ for $1\leq i\leq r$ , which encode the expansion coefficients of the roots in simple roots: $\alpha =\sum _{i=1}^rn_i(\alpha )\alpha _i$ for all $\alpha \in \Phi _0$ . Define

(7.5) $$ \begin{align} I_{\Lambda^\prime}:=\{i\in [1,r] \,\,\, | \,\,\, n_i(\varphi)=1 \,\, \& \,\, (\varpi_i^{\vee}+E_{\text{co}})\cap \Lambda^\prime\not=\emptyset\}. \end{align} $$

For $i\in I_{\Lambda ^\prime }$ , $(\varpi _i^{\vee }+E_{\text {co}})\cap \Lambda ^\prime $ is a coset in $\Lambda ^\prime /(\Lambda ^{\prime }\cap E_{\text {co}})$ , and we fix a representative $\varpi _{i,\Lambda ^\prime }^{\vee }$ once and for all. The set $\Lambda _{\text {min}}^\prime $ of minuscule weights is invariant for the translation action of $\Lambda ^{\prime }\cap E_{\text {co}}$ .

Lemma 7.3. $\{0\}\cup \{\varpi _{i,\Lambda ^\prime }^{\vee }\}_{i\in I_{\Lambda ^\prime }}$ is a complete set of representatives of the $\Lambda ^{\prime }\cap E_{\text {co}}$ -orbits in $\Lambda _{\text {min}}^\prime $ .

Proof. See, for example, [Reference Macdonald29, §2.5] and [Reference Stokman, Koornwinder and Stokman44, §9.3.4].

For $\lambda \in \Lambda ^\prime $ , the element $w_{\lambda ,\Lambda ^\prime }\in W_{\Lambda ^\prime }$ is related to the shortest element $w_\lambda \in W$ such that $\lambda =w_\lambda c_\lambda $ in the following way.

Lemma 7.4. For $\lambda \in \Lambda ^\prime $ , we have $c_\lambda \in \Lambda ^\prime $ and $w_{c_\lambda ,\Lambda ^\prime }\in \Omega _{\Lambda ^\prime }$ . Furthermore,

(7.6) $$ \begin{align} w_\lambda=w_{\lambda,\Lambda^\prime}w_{c_\lambda,\Lambda^\prime}^{-1} \end{align} $$

in $W_{\Lambda ^\prime }$ .

Proof. Fix $\lambda \in \Lambda ^\prime $ and write $\zeta :=c_\lambda $ . Then clearly, $\zeta \in \overline {C}_+\cap \Lambda ^\prime =\Lambda _{\text {min}}^\prime $ . Hence, $w_{\zeta ,\Lambda ^\prime }\in \Omega _{\Lambda ^\prime }$ by Proposition 7.2, proving the first statement.

For the second statement, first note that

$$\begin{align*}w_{\lambda,\Lambda^\prime} w_{\zeta,\Lambda^\prime}^{-1}(0)=\lambda-v_{\lambda,\Lambda^\prime}^{-1}v_{\zeta,\Lambda^\prime}\zeta. \end{align*}$$

This lies in $Q^{\vee }$ since $\lambda \in \zeta +Q^{\vee }$ ; hence, $w_{\lambda ,\Lambda ^\prime }w_{\zeta ,\Lambda ^\prime }^{-1}\in W$ .

Now note that

$$\begin{align*}w_{\lambda,\Lambda^\prime}w_{\zeta,\Lambda^\prime}^{-1}(\zeta)= w_{\lambda,\Lambda^\prime}(0)=\tau(\lambda)(0)=\lambda. \end{align*}$$

So $w_{\lambda ,\Lambda ^\prime }w_{\zeta ,\Lambda ^\prime }^{-1}$ and $w_\lambda $ are two elements in W mapping $\zeta $ to $\lambda $ . Since $w_\lambda \in W^\zeta $ , this implies that $\ell (w_{\lambda ,\Lambda ^\prime })=\ell (w_{\lambda ,\Lambda ^\prime }w_{\zeta ,\Lambda ^\prime }^{-1})\geq \ell (w_\lambda )$ (the equality follows from the fact that $w_{\zeta ,\Lambda ^\prime }^{-1}\in \Omega _{\Lambda ^\prime }$ ). However, $w_\lambda w_{\zeta ,\Lambda ^\prime }\in \tau (\lambda )W_0$ , so by definition of $w_{\lambda ,\Lambda ^\prime }\in W_{\Lambda ^\prime }$ , we have $\ell (w_\lambda )=\ell (w_\lambda w_{\zeta ,\Lambda ^\prime })\geq \ell (w_{\lambda ,\Lambda ^\prime })$ with equality iff $w_\lambda w_{\zeta ,\Lambda ^\prime }=w_{\lambda ,\Lambda ^\prime }$ . Combining the two inequalities, we conclude that $\ell (w_{\lambda ,\Lambda ^\prime })=\ell (w_\lambda )$ , and hence, $w_\lambda =w_{\lambda ,\Lambda ^\prime }w_{\zeta ,\Lambda ^\prime }^{-1}$ .

Remark 7.5. Note that $c_\lambda =0$ iff $\lambda \in Q^{\vee }$ , in which case, $w_{0,\Lambda ^\prime }=1$ and (7.6) reduces to (7.3).

7.2 The extended double affine Hecke algebra

Fix lattices $(\Lambda ,\Lambda ^\prime )\in (\mathcal {L}^{\times 2})_e$ and assume that $q\in \mathbf {F}^\times $ has an $e^{\text {th}}$ root $q^{\frac {1}{e}}\in \mathbf {F}^\times $ , which we fix once and for all. We write

$$\begin{align*}q^{\ell/e}:=(q^{\frac{1}{e}})^{\ell},\qquad q_\alpha^{\ell/e}:=(q^{\frac{1}{e}})^{\frac{2\ell}{\|\alpha\|^2}} \end{align*}$$

for $\ell \in \mathbb {Z}$ (recall that $2/\|\alpha \|^2\in \mathbb {Z}$ by the choice of normalisation of $\Phi _0$ ).

The extended affine Weyl group $W_{\Lambda ^\prime }$ then acts on $T_\Lambda $ by q-dilations and reflections. In particular, $\tau (\lambda )\in W_{\Lambda ^\prime }$ ( $\lambda \in \Lambda ^\prime $ ) acts by $\tau (\lambda )t=q^\lambda t$ (here, $q^\lambda \in T_{\Lambda }$ is the torus element mapping $\mu \in \Lambda $ to $q^{\langle \lambda ,\mu \rangle }$ ). Later on, we will also use the $W_\Lambda $ -action on $T_{\Lambda ^\prime }$ by q-dilations and reflections.

Remark 7.6. The natural map $j_\Lambda : T_{P^{\vee }}\rightarrow T_\Lambda $ from Corollary 2.9 is W-equivariant. Furthermore, if $(P^{\vee },\Lambda ^\prime )\in (\mathcal {L}^{\times 2})_e$ , then $j_\Lambda $ is $W_{\Lambda ^\prime }$ -equivariant if either $\Lambda \subset E^\prime $ or $\Lambda ^\prime \subset E^\prime $ .

The contragredient representation gives a $W_{\Lambda ^\prime }$ -action on $\mathcal {P}_\Lambda $ by algebra automorphisms. The explicit formulas for the action are given by (2.32). The action of the subgroup $\Omega _{\Lambda ^\prime }$ on $\mathcal {P}_\Lambda $ is given by

(7.7) $$ \begin{align} w_{\zeta,\Lambda^\prime}(x^\lambda)=q^{-\langle v_{\zeta,\Lambda^\prime}\zeta,\lambda\rangle}x^{v_{\zeta,\Lambda^\prime}^{-1}\lambda} \end{align} $$

for $\zeta \in \Lambda _{\text {min}}^\prime $ and $\lambda \in \Lambda $ . The formulas $T_j\mapsto T_{\omega (j)}$ ( $\omega \in \Omega _{\Lambda ^\prime }$ , $0\leq j\leq r$ ) extend the action of $\Omega _{\Lambda ^\prime }$ on $\mathcal {P}=\mathcal {P}_{Q^{\vee }}$ to an action of $\Omega _{\Lambda ^\prime }$ on $\mathbb {H}$ by algebra automorphisms.

Definition 7.7. The smashed product algebra $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }:=\Omega _{\Lambda ^\prime }\ltimes \mathbb {H}$ is the Y-extended double affine Hecke algebra.

Concretely, we have the commutation relation $\omega T_j=T_{\omega (j)}\omega $ for $\omega \in \Omega _{\Lambda ^\prime }$ and $0\leq j\leq r$ in $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ , as well as the commutation relations

(7.8) $$ \begin{align} w_{\zeta,\Lambda^\prime}x^\mu=q^{-\langle v_{\zeta,\Lambda^\prime}\zeta,\mu\rangle}x^{v_{\zeta,\Lambda^\prime}^{-1}\mu}\,w_{\zeta,\Lambda^\prime} \end{align} $$

in $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ for $\zeta \in \Lambda _{\text {min}}^\prime $ and $\mu \in Q^{\vee }$ .

The extended affine Hecke algebra $H_{\Lambda ^\prime }$ is the subalgebra $\Omega _{\Lambda ^\prime }\ltimes H$ of $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ . Multiplication defines a $\mathbf {F}$ -linear isomorphism $\mathcal {P}\otimes H_{\Lambda ^\prime }\overset {\sim }{\longrightarrow }\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ . If $w=\omega s_{j_1}\cdots s_{j_\ell }\in W_{\Lambda ^\prime }$ ( $\omega \in \Omega _{\Lambda ^\prime }$ , $0\leq j_i\leq r$ ) is a reduced expression, then

$$\begin{align*}T_w:=\omega T_{j_1}\cdots T_{j_\ell}\in H_{\Lambda^\prime} \end{align*}$$

is well defined.

There exists a unique group homomorphism $\Lambda ^\prime \rightarrow H_{\Lambda ^\prime }$ , $\lambda \mapsto Y^\lambda $ , such that

(7.9) $$ \begin{align} Y^{\lambda}=T_{\tau(\lambda)}\qquad\quad (\lambda\in\Lambda^{\prime +}:=\Lambda^\prime\cap \overline{E}_+). \end{align} $$

Note that $Y^\mu $ for $\mu \in Q^{\vee }$ reduces to the corresponding element in H, as defined in Subsection 2.7. Then

$$\begin{align*}\{T_vY^\lambda \,\,\, | \,\,\, v\in W_0,\, \lambda\in\Lambda^\prime\} \end{align*}$$

is a basis of $H_{\Lambda ^\prime }$ , and the cross relations (2.46) in $H_{\Lambda ^\prime }$ hold true for $\mu \in \Lambda ^\prime $ and $1\leq i\leq r$ .

The elements in $\Omega _{\Lambda ^\prime }$ , viewed as elements in $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ , factorise as follows:

(7.10) $$ \begin{align} w_{\zeta,\Lambda^\prime}=Y^{\zeta}T_{v_{\zeta,\Lambda^\prime}}^{-1}\qquad (\zeta\in \Lambda_{\text{min}}^\prime). \end{align} $$

Definition 7.8. The extended double affine Hecke algebra $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ is the algebra generated by $\mathcal {P}_\Lambda $ and $H_{\Lambda ^\prime }$ subject to the cross relations (2.40) for $0\leq j\leq r$ and $p\in \mathcal {P}_\Lambda $ , as well as the relations (7.8) for $\zeta \in \Lambda _{\text {min}}^\prime $ and $\mu \in \Lambda $ .

Note that

(7.11) $$ \begin{align} \mathbb{H}_{\Lambda,\Lambda^\prime}\simeq\Omega_{\Lambda^\prime}\ltimes\mathbb{H}_{\Lambda,Q^{\vee}} \end{align} $$

with the action of $\Omega _{\Lambda ^\prime }$ on $\mathbb {H}_{\Lambda ,Q^{\vee }}$ by algebra automorphisms defined by (7.7) and $\omega (T_j):=T_{\omega (j)}$ for $\omega \in \Omega _{\Lambda ^\prime }$ and $0\leq j\leq r$ . The algebra $\mathbb {H}_{\Lambda ,Q^{\vee }}$ is the X-extended double affine Hecke algebra.

The duality anti-involution $\delta $ of $\mathbb {H}$ defined in Theorem 2.14 extends to an anti-algebra isomorphism $\delta :\mathbb {H}_{\Lambda ,\Lambda ^\prime }\rightarrow \mathbb {H}_{\Lambda ^\prime ,\Lambda }$ by

$$\begin{align*}\delta(Y^\lambda)=x^{-\lambda}\qquad (\lambda\in\Lambda^\prime). \end{align*}$$

Note that $\delta $ restricts to an anti-algebra isomorphism $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }\overset {\sim }{\longrightarrow }\mathbb {H}_{\Lambda ^\prime ,Q^{\vee }}$ , and

$$\begin{align*}\delta\big(w_{\zeta,\Lambda^\prime}\big)=T_{v_{\zeta,\Lambda^\prime}^{-1}}^{-1}x^{-\zeta}\qquad (\zeta\in\Lambda_{\text{min}}^\prime). \end{align*}$$

The extended X-intertwiners are defined by

$$\begin{align*}S_{\omega w}^X:= \omega S_w^X\in\mathbb{H}_{Q^{\vee},\Lambda^\prime}\qquad (\omega\in\Omega_{\Lambda^\prime}, w\in W). \end{align*}$$

They satisfy (2.51) for $w\in W_{\Lambda ^\prime }$ and $p\in \mathcal {P}_\Lambda $ . The extended Y-intertwiners are

$$\begin{align*}S_w^Y:=\delta(S_{w^{-1}}^X)\in\mathbb{H}_{\Lambda,Q^{\vee}} \qquad (w\in W_\Lambda). \end{align*}$$

In particular,

(7.12) $$ \begin{align} S_{w_{\zeta,\Lambda}}^Y=\delta\big(w_{\zeta,\Lambda}^{-1})=x^\zeta T_{v_{\zeta,\Lambda}^{-1}}\qquad (\zeta\in \Lambda_{\text{min}}). \end{align} $$

The extended Y-intertwiners $S_w^Y$ ( $w\in W_\Lambda $ ) satisfy (2.59) for $p\in \mathcal {P}_{\Lambda ^\prime }$ .

7.3 The extended quasi-polynomial representation

Recall the definition of the base-point $\mathfrak {s}_y\in T_{\Lambda ^\prime }$ ( $y\in E$ ) from (5.2). We have the following extension of Lemma 5.3.

Lemma 7.9. For $\omega \in \Omega _{\Lambda }$ and $y\in E$ , we have $(D\omega )\mathfrak {s}_y=\mathfrak {s}_{\omega y}$ in $T_{\Lambda ^\prime }$ .

Proof. Let $\zeta \in \Lambda _{\text {min}}$ such that $\omega =w_{\zeta ,\Lambda }=\tau (\zeta )v_{\zeta ,\Lambda }^{-1}$ (see Proposition 7.2). By the definition (5.2) of $\mathfrak {s}_y$ , we have

(7.13) $$ \begin{align} v_{\zeta,\Lambda}^{-1}\mathfrak{s}_y=\mathfrak{s}_{w_{\zeta,\Lambda}y} \prod_{\alpha\in\Phi_0^+\setminus\Pi(v_{\zeta,\Lambda})}k_\alpha^{(\eta(\alpha(v_{\zeta,\Lambda}^{-1}y))- \eta(\alpha(w_{\zeta,\Lambda}y)))\alpha} \prod_{\beta\in\Pi(v_{\zeta,\Lambda})}k_\beta^{-(\eta(-\beta(v_{\zeta,\Lambda}^{-1}y))+\eta(\beta(w_{\zeta,\Lambda}y)))\beta} \end{align} $$

in $T_{\Lambda ^\prime }$ . The second line of (7.4) gives

$$\begin{align*}\eta(\alpha(v_{\zeta,\Lambda}^{-1}y))- \eta(\alpha(w_{\zeta,\Lambda}y))=\eta(\alpha(v_{\zeta,\Lambda}^{-1}y))- \eta(\alpha(v_{\zeta,\Lambda}^{-1}y))=0 \end{align*}$$

for $\alpha \in \Phi _0^+\setminus \Pi (v_{\zeta ,\Lambda })$ ; hence, the product over $\alpha $ in (7.13) is equal to $1$ . Formula (5.1) and the first line of (7.4) imply that the product over $\beta $ in (7.13) is equal to $1$ .

For $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ , we extend the quasi-polynomial representation $\pi _{c,\mathfrak {t}\vert _{Q^{\vee }}}$ of $\mathbb {H}$ to $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ under suitably further restrictions on $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ . It is convenient to discuss the extension to the Y-extended double affine Hecke algebra $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ first, in which case, we do not need to impose additional conditions on $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ .

Let $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ . Recall the notation $\mathfrak {t}_y:=w_y\mathfrak {t}\in T_{\Lambda ^\prime }$ ( $y\in \mathcal {O}_c$ ) from Definition 4.1. We then have a $\mathfrak {t}$ -dependent $W_{\Lambda ^\prime }\ltimes \mathcal {P}$ -action $(w,p)\mapsto w_{\mathfrak {t}}p$ on $\mathcal {P}^{(c)}$ , with the action of $v\in W_0$ and $\tau (\mu )$ ( $\mu \in \Lambda ^\prime $ ) given by (4.3). Note that

(7.14) $$ \begin{align} \big(w_{\zeta,\Lambda^\prime}\big)_{\mathfrak{t}}(x^y)= \mathfrak{t}_{v_{\zeta,\Lambda^\prime}^{-1}y}^{-\zeta}x^{v_{\zeta,\Lambda^\prime}^{-1}y}\qquad (\zeta\in\Lambda_{\text{min}}^\prime,\, y\in\mathcal{O}_c). \end{align} $$

Lemma 7.10. Let $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ .

The quasi-polynomial representation $\pi _{c,\mathfrak {t}\vert _{Q^{\vee }}}: \mathbb {H}\rightarrow \text {End}(\mathcal {P}^{(c)})$ (see Theorem 4.5) extends to a representation $\pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime }: \mathbb {H}_{Q^{\vee },\Lambda ^\prime }\rightarrow \text {End}(\mathcal {P}^{(c)})$ by

$$\begin{align*}\pi_{c,\mathfrak{t}}^{Q^{\vee},\Lambda^\prime}(\omega)(x^y):=\omega_{\mathfrak{t}}(x^y) \qquad (\omega\in\Omega_{\Lambda^\prime},\, y\in\mathcal{O}_c). \end{align*}$$

Following the notations from Theorem 4.5, we denote the $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ -module $(\mathcal {P}^{(c)}, \pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime })$ again by $\mathcal {P}^{(c)}_{\mathfrak {t}}$ .

Proof. Note that $\mathfrak {t}|_{Q^{\vee }}\in T_J$ ; hence, the quasi-polynomial representation $\pi _{c,\mathfrak {t}\vert _{Q^{\vee }}}$ of $\mathbb {H}$ is well defined. Since

(7.15) $$ \begin{align} w_{\mathfrak{t}}(px^y)=w(p)w_{\mathfrak{t}}(x^y)\qquad (w\in W_{\Lambda^\prime}, p\in\mathcal{P}, y\in\mathcal{O}_c), \end{align} $$

the commutation relations $\omega x^\mu =\omega (x^\mu )\omega $ in $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ for $\omega \in \Omega _{\Lambda ^\prime }$ and $\mu \in Q^{\vee }$ are respected by the proposed extension $\pi _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime }$ of $\pi _{c,\mathfrak {t}\vert _{Q^{\vee }}}$ to $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ . It remains to check that the same is true for the relations $\omega T_j=T_{\omega (j)}\omega $ ( $\omega \in \Omega _{\Lambda ^\prime }$ , $0\leq j\leq r$ ). Note that $k_{\omega (j)}=k_j$ , $\omega s_j=s_{\omega (j)}\omega $ and

$$\begin{align*}\chi_{\mathbb{Z}}(\alpha_{\omega(j)}(D\omega(y)))= \chi_{\mathbb{Z}}(\alpha_{\omega(j)}(\omega y))=\chi_{\mathbb{Z}}(\alpha_j(y)); \end{align*}$$

hence, it suffices to show that

$$\begin{align*}\omega_{\mathfrak{t}}\big(\nabla_j(x^y)\big)=\nabla_{\omega(j)}(\omega_{\mathfrak{t}}(x^y)) \end{align*}$$

for $\omega \in \Omega _{\Lambda ^\prime }$ , $0\leq j\leq r$ and $y\in \mathcal {O}_c$ .

Fix $\omega \in \Omega _{\Lambda ^\prime }$ and $0\leq j\leq r$ . By (2.31), which holds true for $w\in W_{\Lambda ^\prime }$ and $a\in \Phi $ , and (7.15) we have

(7.16) $$ \begin{align} \omega_{\mathfrak{t}}\big(\nabla_j(x^y)\big)= \left(\frac{1-x^{-\lfloor D\alpha_j(y)\rfloor \alpha_{\omega(j)}^{\vee}}}{1-x^{\alpha_{\omega(j)}^{\vee}}} \right)\,\omega_{\mathfrak{t}}(x^y). \end{align} $$

Now note that $D\alpha _{\omega (j)}((D\omega )y)=D(\omega ^{-1}\alpha _{\omega (j)})(y)= D\alpha _j(y)$ ; hence, the right-hand side of (7.16) is equal to $\nabla _{\omega (j)}(\omega _{\mathfrak {t}}(x^y))$ , as desired.

In terms of the new notation from Lemma 7.10, the quasi-polynomial representation $\pi _{c,\mathfrak {t}}$ is equal to $\pi _{c,\mathfrak {t}}^{Q^{\vee },Q^{\vee }}$ .

In the same way, Theorem 4.9 extends as follows.

Lemma 7.11. Let $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ . The formulas

$$ \begin{align*} \begin{aligned} \sigma_{c,\mathfrak{t}}^{Q^{\vee},\Lambda^\prime}(s_j)(x^y)&:= \frac{k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}(x^{\alpha_j^{\vee}}-1)}{(k_jx^{\alpha_j^{\vee}}-k_j^{-1})} s_{j,\mathfrak{t}}(x^y)+\frac{(k_j-k_j^{-1})}{(k_jx^{\alpha_j^{\vee}}-k_j^{-1})}x^{y-\lfloor D\alpha_j(y) \rfloor\alpha_j^{\vee}},\\ \sigma_{c,\mathfrak{t}}^{Q^{\vee},\Lambda^\prime}(\omega)(x^y)&:=\omega_{\mathfrak{t}}(x^y),\\ \sigma_{c,\mathfrak{t}}^{Q^{\vee},\Lambda^\prime}(f)(x^y)&:=fx^y \end{aligned} \end{align*} $$

for $0\leq j\leq r$ , $y\in \mathcal {O}_c$ , $\omega \in \Omega _{\Lambda ^\prime }$ and $f\in \mathcal {Q}$ define a representation

$$\begin{align*}\sigma_{c,\mathfrak{t}}^{Q^{\vee},\Lambda^\prime}: W_{\Lambda^\prime}\ltimes\mathcal{Q}\rightarrow\text{End}(\mathcal{Q}^{(c)}). \end{align*}$$

As the next step, we introduce a natural extension of the quasi-polynomial representation from $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ to $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ . Its representation space will be

$$\begin{align*}\ \mathcal{P}_\Lambda^{(c)}:=\bigoplus_{y\in\mathcal{O}_{\Lambda,c}}\mathbf{F}x^y, \end{align*}$$

where $\mathcal {O}_{\Lambda ,c}$ is the $W_\Lambda $ -orbit of $c\in C^J$ in E. We realise it as a quotient of the induced $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -module

$$\begin{align*}V_{c,\mathfrak{t}}:=\mathbb{H}_{\Lambda,\Lambda^\prime}\otimes_{\mathbb{H}_{Q^{\vee},\Lambda^\prime}}\mathcal{P}_{\mathfrak{t}}^{(c)} \end{align*}$$

under suitable further restrictions on $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ .

We first consider the $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -module $V_{c,\mathfrak {t}}$ in more detail.

Lemma 7.12. Let $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ . As a $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ -module, we have

$$\begin{align*}V_{c,\mathfrak{t}} \simeq \bigoplus_{\omega\in\Omega_\Lambda}\mathcal{P}_{\omega\mathfrak{t}}^{(\omega c)}. \end{align*}$$

Proof. Applying $\delta $ to the decomposition (7.11), it follows that the subspaces

$$\begin{align*}V_{c,\mathfrak{t}}(\omega):=\text{span}\{S_\omega^Y\otimes_{\mathbb{H}_{Q^{\vee},\Lambda^\prime}}f \,\,\, | \,\,\, f\in\mathcal{P}_{\mathfrak{t}}^{(c)}\} \qquad (\omega\in\Omega_\Lambda) \end{align*}$$

are $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ -submodules of $V_{c,\mathfrak {t}}$ , and

$$\begin{align*}V_{c,\mathfrak{t}}=\bigoplus_{\omega\in\Omega_\Lambda}V_{c,\mathfrak{t}}(\omega). \end{align*}$$

It is easy to check that $V_{c,\mathfrak {t}}(\omega )\simeq \mathcal {P}_{\omega \mathfrak {t}}^{(\omega c)}$ as $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ -modules, with the isomorphism mapping $S_\omega ^Y\otimes _{\mathbb {H}_{Q^{\vee },\Lambda ^\prime }}x^c$ to $x^{\omega c}$ .

Let $c\in C^J$ and write

$$\begin{align*}\Omega_{\Lambda,J}:=\{\omega\in\Omega_{\Lambda} \,\, | \,\, \omega(J)=J\}. \end{align*}$$

Denote by $W_{\Lambda ,c}$ the fix-point subgroup in $W_\Lambda $ of c, and by $\Omega _{\Lambda ,c}$ the fix-point subgroups in $\Omega _{\Lambda }$ of c. Then $\Omega _{\Lambda ,c}\subseteq \Omega _{\Lambda ,J}$ and $W_{\Lambda ,c}=\Omega _{\Lambda ,c}\ltimes W_J$ . Write $\Omega _{\Lambda }^c$ for a complete set of representatives of $\Omega _{\Lambda }/\Omega _{\Lambda ,c}$ . The $W_\Lambda $ -orbit $\mathcal {O}_{\Lambda ,c}:=W_\Lambda c$ then decomposes as

$$\begin{align*}\mathcal{O}_{\Lambda,c}=\bigsqcup_{\omega\in\Omega^c_\Lambda}\mathcal{O}_{\omega c}. \end{align*}$$

In particular, we have

$$\begin{align*}\mathcal{P}_\Lambda^{(c)}= \bigoplus_{\omega\in\Omega_{\Lambda}^c}\mathcal{P}^{(\omega c)}. \end{align*}$$

as $\mathcal {P}$ -modules.

We now look for a quotient of the induced $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -module $V_{c,\mathfrak {t}}$ admitting a quasi-polynomial realisation on $\mathcal {P}_{\Lambda }^{(c)}$ . For this, we need to restrict to the following parameter subset of $T_{\Lambda ^\prime ,J}$ to the following subset.

Definition 7.13. For $c\in C^J$ , write

$$\begin{align*}{}^\Lambda T_{\Lambda^\prime}^{c}:=\{\mathfrak{t}\in T_{\Lambda^\prime,J} \,\, | \,\, \omega\mathfrak{t}=\mathfrak{t}\,\,\, \forall\,\omega\in\Omega_{\Lambda,c}\}. \end{align*}$$

The second and third example below should be compared to Lemma 3.3.

Example 7.14.

  1. 1. For $\omega \in \Omega _\Lambda $ and $c\in \overline {C}_+$ , we have $\omega ({}^{\Lambda }T_{\Lambda ^\prime }^{c})={}^{\Lambda }T_{\Lambda ^\prime }^{\omega c}$ .

  2. 2. For $c\in C^J$ , we have ${}^{Q^{\vee }}T_{\Lambda ^\prime }^c=T_{\Lambda ^\prime ,J}$ . Furthermore, the restriction map $T_{\Lambda ^\prime ,J}\rightarrow T_J$ , $s\mapsto s\vert _{Q^{\vee }}$ restricts to a map ${}^\Lambda T_{\Lambda ^\prime }^c\rightarrow {}^\Lambda T_{Q^{\vee }}^c$ (Lemma 3.3(2) does not have an apparent extension to the present context).

  3. 3. For all $c\in \overline {C}_+$ , the subset ${}^{\Lambda }T_{\Lambda ^\prime }^c$ of $T_{\Lambda ^\prime }$ is a $T_{\Lambda ^\prime ,[1,r]}$ -subset (use here the fact that $T_{\Lambda ,[1,r]}\subseteq T_{\Lambda }^{W_0}$ ).

  4. 4. If $\zeta \in \Lambda _{\text {min}}$ , then

    $$\begin{align*}{}^{\Lambda}T_{\Lambda^\prime}^\zeta=q^\zeta T_{\Lambda^\prime,[1,r]}. \end{align*}$$

    Indeed, note that ${}^{\Lambda }T_{\Lambda ^\prime }^0=T_{\Lambda ^\prime ,[1,r]}$ since $\Omega _{\Lambda ,0}=\{1\}$ and $\mathbf {J}(0)=[1,r]$ , and so

    $$\begin{align*}{}^{\Lambda}T_{\Lambda^\prime}^\zeta=w_{\zeta,\Lambda}({}^{\Lambda}T_{\Lambda^\prime}^0)=q^\zeta v_{\zeta,\Lambda}^{-1}(T_{\Lambda^\prime,[1,r]})= q^\zeta T_{\Lambda^\prime,[1,r]}. \end{align*}$$

    Here, we used part (2) and the fact that $w_{\zeta ,\Lambda }=\tau (\zeta )v_{\zeta ,\Lambda }^{-1}\in \Omega _\Lambda $ maps $0$ to $\zeta $ .

The lifting properties of characters as described in Lemma 3.3 do not seem to extend easily to ${}^\Lambda T_{\Lambda ^\prime }^c$ .

For $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^{c}$ , the induced module $V_{c,\mathfrak {t}}$ admits the following natural family of intertwiners.

Lemma 7.15. Let $c\in C^J$ and $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ . For each $\omega \in \Omega _{\Lambda ,c}$ , there exists a unique intertwiner

$$\begin{align*}\phi_\omega\in\text{End}_{\mathbb{H}_{\Lambda,\Lambda^\prime}} \big(V_{c,\mathfrak{t}}\big) \end{align*}$$

such that

(7.17) $$ \begin{align} \phi_\omega\big(S_{\omega^\prime}^Y\otimes_{\mathbb{H}_{Q^{\vee},\Lambda^\prime}}x^c\big) =S_{\omega^\prime\omega}^Y\otimes_{\mathbb{H}_{Q^{\vee},\Lambda^\prime}}x^c \qquad (\omega^\prime\in\Omega_\Lambda). \end{align} $$

Proof. Let $\omega ^\prime \in \Omega _\Lambda $ . Since $\omega \in \Omega _{\Lambda ,c}$ , the condition $\omega \mathfrak {t}=\mathfrak {t}$ and the proof of Lemma 7.12 provide isomorphisms

$$\begin{align*}V_{c,\mathfrak{t}}(\omega^\prime)\overset{\sim}{\longrightarrow} \mathcal{P}_{\omega^\prime\mathfrak{t}}^{(\omega^\prime c)}\overset{\sim}{\longrightarrow} V_{c,\mathfrak{t}}(\omega^\prime\omega) \end{align*}$$

of $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ -modules. The resulting $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ -linear isomorphism $V_{c,\mathfrak {t}}(\omega ^\prime )\overset {\sim }{\longrightarrow } V_{c,\mathfrak {t}}(\omega ^\prime \omega )$ is characterised by

$$\begin{align*}S_{\omega^\prime}^Y\otimes_{\mathbb{H}_{Q^{\vee},\Lambda^\prime}}x^c\mapsto S_{\omega^\prime\omega}^Y\otimes_{\mathbb{H}_{Q^{\vee},\Lambda^\prime}}x^c. \end{align*}$$

Hence, (7.17) gives rise to a well-defined intertwiner $\phi _\omega \in \text {Hom}_{\mathbb {H}_{Q^{\vee },\Lambda ^\prime }}(V_{c,\mathfrak {t}})$ . A straightforward check shows that $\phi _\omega $ also intertwines the $\delta (\Omega _\Lambda )$ -action on $V_{c,\mathfrak {t}}$ .

Recall that for $c\in C^J$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,J}$ , we introduced W-translates $\mathfrak {t}_y\in T_{\Lambda ^\prime }$ for $y\in \mathcal {O}_c$ in Definition 4.1. In case $\mathfrak {t}\in {}^{\Lambda }T^c_{\Lambda ^\prime }\subseteq T_{\Lambda ^\prime ,J}$ , we extend this construction to $W_\Lambda $ -translates as follows.

Definition 7.16. Let $c\in C^J$ and $\mathfrak {t}\in {}^{\Lambda }T^c_{\Lambda ^\prime }$ . Define $\mathfrak {t}_{y;c}\in T_{\Lambda ^\prime }$ for $y\in \mathcal {O}_{\Lambda ,c}$ by

(7.18) $$ \begin{align} \mathfrak{t}_{wc;c}:=w\mathfrak{t} \qquad (w\in W_\Lambda). \end{align} $$

Note that (7.18) is well defined since $\mathfrak {t}\in {}^{\Lambda }T^c_{\Lambda ^\prime }$ implies that $w\mathfrak {t}=\mathfrak {t}$ for all $w\in W_{\Lambda ,c}$ . Furthermore, $\mathfrak {t}_{y;c}=\mathfrak {t}_y$ for $y\in \mathcal {O}_c$ , and more generally,

(7.19) $$ \begin{align} \mathfrak{t}_{y;c}=(\omega\mathfrak{t})_y\,\,\,\mbox{ when }\, y\in\mathcal{O}_{\omega c}\, \mbox{ and }\, \omega\in\Omega_\Lambda. \end{align} $$

Furthermore, note that for $\omega \in \Omega _\Lambda $ , we have $\omega \big ({}^{\Lambda }T_{\Lambda ^\prime }^c\big )={}^{\Lambda }T_{\Lambda ^\prime }^{\omega c}\subseteq T_{\Lambda ^\prime ,\omega (J)}$ by Example 7.14(1), and

(7.20) $$ \begin{align} (\omega\mathfrak{t})_{y;\omega c}=\mathfrak{t}_{y;c}\,\,\, \mbox{ when }\, y\in\mathcal{O}_{\Lambda,c}\, \mbox{ and }\, \omega\in\Omega_\Lambda. \end{align} $$

For $w\in W_\Lambda $ and $y\in \mathcal {O}_{\Lambda ,c}$ , we have $\mathfrak {t}_{wy;c}=w\mathfrak {t}_{y;c}$ . For translations, this implies that

(7.21) $$ \begin{align} \mathfrak{t}_{y+\lambda;c}=q^{\lambda}\mathfrak{t}_{y;c}\qquad (y\in\mathcal{O}_{\Lambda,c},\, \lambda\in\Lambda). \end{align} $$

Lemma 4.2 now immediately generalises as follows.

Lemma 7.17. For $c\in C^J$ and $\mathfrak {t}\in {}^{\Lambda }T^c_{\Lambda ^\prime }$ , the extended affine Weyl group $W_{\Lambda ^\prime }$ acts linearly on $\mathcal {P}_\Lambda ^{(c)}$ by

$$ \begin{align*} \begin{aligned} v_{\mathfrak{t};c}(x^y)&:=x^{vy}\qquad\qquad\qquad (v\in W_0),\\ \tau(\mu)_{\mathfrak{t};c}(x^y)&:= \mathfrak{t}_{y;c}^{-\mu}x^y \qquad\qquad\quad (\mu\in\Lambda^\prime) \end{aligned} \end{align*} $$

for $y\in \mathcal {O}_{\Lambda ,c}$ .

Note that $w_{\mathfrak {t};c}(pf)=w(p)w_{\mathfrak {t};c}(f)$ for $w\in W_{\Lambda ^\prime }$ , $p\in \mathcal {P}_\Lambda $ and $f\in \mathcal {P}_\Lambda ^{(c)}$ , and

(7.22) $$ \begin{align} \begin{aligned} s_{0,\mathfrak{t};c}(x^y)&= \mathfrak{t}_{y;c}^{\varphi^{\vee}}x^{s_\varphi y},\\ (w_{\zeta,\Lambda^\prime})_{\mathfrak{t};c}(x^y)&= \mathfrak{t}_{v_{\zeta,\Lambda^\prime}^{-1}y;c}^{-\zeta}x^{v_{\zeta,\Lambda^\prime}^{-1}y} \end{aligned} \end{align} $$

for $y\in \mathcal {O}_{\Lambda ,c}$ and $\zeta \in \Lambda _{\text {min}}^\prime $ (cf. (7.14)). Note furthermore that $\big (w_{\zeta ,\Lambda ^\prime }\big )_{\mathfrak {t};c}(x^y)$ is a constant multiple of $x^{(Dw_\zeta )y}$ , and that for $c\in C^J$ , $\omega \in \Omega _{\Lambda }$ and $\mathfrak {t}\in {}^{\Lambda }T^c_{\Lambda ^\prime }$ ,

(7.23) $$ \begin{align} w_{\mathfrak{t};c}(x^y)=w_{\omega\mathfrak{t}}(x^y)\qquad (w\in W_{\Lambda^\prime},\, y\in\mathcal{O}_{\omega c}). \end{align} $$

We now have the following extension of Lemma 7.10 to $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ .

Theorem 7.18. Let $c\in C^J$ and $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ .

  1. 1. The formulas

    (7.24) $$ \begin{align} \begin{aligned} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(T_j)x^y&:=k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}s_{j,\mathfrak{t};c}(x^y) +(k_j-k_j^{-1})\nabla_j(x^y),\\ \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(x^\lambda)x^y&:=x^{y+\lambda},\\ \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(\omega)x^y&:=\omega_{\mathfrak{t};c}(x^y) \end{aligned} \end{align} $$
    for $0\leq j\leq r$ , $\lambda \in \Lambda $ , $\omega \in \Omega _{\Lambda ^\prime }$ and $y\in \mathcal {O}_{\Lambda ,c}$ define a representation
    $$\begin{align*}\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}: \mathbb{H}_{\Lambda,\Lambda^\prime}\rightarrow\text{End} (\mathcal{P}_\Lambda^{(c)}). \end{align*}$$

    We denote the resulting $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -module by $\mathcal {P}_{\Lambda ,\mathfrak {t}}^{(c)}$ .

  2. 2. As $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ -modules,

    (7.25) $$ \begin{align} \mathcal{P}_{\Lambda,\mathfrak{t}}^{(c)}=\bigoplus_{\omega\in\Omega_\Lambda^c}\mathcal{P}_{\omega\mathfrak{t}}^{(\omega c)}. \end{align} $$
  3. 3. $\mathcal {P}_{\Lambda ,\mathfrak {t}}^{(c)}$ is a quotient of $V_{c,\mathfrak {t}}$ .

Proof. Fix $\omega \in \Omega ^c_\Lambda $ . We first show that

(7.26) $$ \begin{align} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(h)x^y=\pi_{\omega c,\omega\mathfrak{t}}^{Q^{\vee},\Lambda^\prime}(h)x^y\qquad (y\in\mathcal{O}_{\omega c}) \end{align} $$

for $h=T_j$ ( $0\leq j\leq r$ ), $h=x^\mu $ ( $\mu \in Q^{\vee }$ ) and $h=\omega ^\prime $ ( $\omega ^\prime \in \Omega _{\Lambda ^\prime }$ ). Observe that the right-hand side is well defined since $\omega c\in C^{\omega (J)}$ and $\omega \mathfrak {t}\in T_{\Lambda ^\prime ,\omega (J)}$ . Formula (7.26) is trivially correct for $h=x^\mu $ ( $\mu \in Q^{\vee }$ ), and it follows from (7.23) when $h=\omega ^\prime $ ( $\omega ^\prime \in \Omega _{\Lambda ^\prime }$ ). For $h=T_j$ ( $0\leq j\leq r$ ), formula (7.26) follows from the observation that

$$ \begin{align*} \begin{aligned} s_{a,\omega\mathfrak{t}}(x^y)= (\omega\mathfrak{t})_y^{-m\alpha^{\vee}} x^{s_\alpha y}= \mathfrak{t}_{y;c}^{-m\alpha^{\vee}}x^{s_\alpha y}= s_{a,\mathfrak{t};c}(x^y) \end{aligned} \end{align*} $$

for $y\in \mathcal {O}_{\omega c}$ and $a=(\alpha ,m)\in \Phi $ , where we have used (7.19) in the second equality.

Proof of (1): In view of Lemma 7.10 and the previous paragraph, it suffices to show that the formulas (7.24) respect the defining relations of $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ involving the generators $x^\lambda $ ( $\lambda \in \Lambda $ ). The properties of the action $(w,f)\mapsto w_{\mathfrak {t};c}(f)$ of $w\in W_{\Lambda ^\prime }$ on $f\in \mathcal {P}_\Lambda ^{(c)}$ mentioned directly after Lemma 7.17 imply that $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }$ respect the relation $\omega x^\lambda =\omega (x^\lambda )\omega $ in $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ ( $\omega \in \Omega _{\Lambda ^\prime }$ , $\lambda \in \Lambda $ ). Using (4.6), the cross relations (2.40) for the operators $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(T_j)$ and $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(x^\lambda )$ ( $0\leq j\leq r$ , $\lambda \in \Lambda $ ) are verified by a direct computation.

Proof of (2): This is immediate from (7.26) and part (1).

Proof of (3): This is clear from part (2) and Lemma 7.12.

Remark 7.19.

  1. 1. Cherednik’s polynomial representation of the extended double affine Hecke algebra is

    $$\begin{align*}\pi_{0,1_{T_{\Lambda^\prime}}}^{\Lambda,\Lambda^\prime}: \mathbb{H}_{\Lambda,\Lambda^\prime}\rightarrow\text{End}(\mathcal{P}_\Lambda). \end{align*}$$
  2. 2. Let $c\in \overline {C}_+$ and $c^\prime \in \overline {C}_+\cap \mathcal {O}_{\Lambda ,c}$ . Then $c^\prime =\omega c$ for a unique $\omega \in \Omega _\Lambda $ , and

    $$\begin{align*}\pi_{\omega\mathfrak{t};\omega c}^{\Lambda,\Lambda^\prime}=\pi_{\mathfrak{t};c}^{\Lambda,\Lambda^\prime}\qquad\quad (\omega\in\Omega_\Lambda) \end{align*}$$
    by Example 7.14(2) and (7.20). In particular, for $\zeta \in \Lambda _{\text {min}}$ , we have $\mathcal {P}_\Lambda ^{(\zeta )}=\mathcal {P}_\Lambda $ and
    (7.27) $$ \begin{align} \pi_{\zeta,\mathfrak{t}}^{\Lambda,\Lambda^\prime}=\pi_{0,w_{\zeta,\Lambda}^{-1}\mathfrak{t}}^{\Lambda,\Lambda^\prime}=\pi_{0,q^{-\zeta}\mathfrak{t}}^{\Lambda,\Lambda^\prime} \end{align} $$
    for $\mathfrak {t}\in {}^\Lambda T_{\Lambda ^\prime }^\zeta =q^\zeta T_{\Lambda ^\prime ,[1,r]}$ (see Example 7.14(4)). For the second equality of (7.27), we used that $w_{\zeta ,\Lambda ^\prime }^{-1}\mathfrak {t}=v_{\zeta ,\Lambda ^\prime }(q^{-\zeta }\mathfrak {t})=q^{-\zeta }\mathfrak {t}$ since $q^{-\zeta }\mathfrak {t}\in T_{\Lambda ^\prime ,[1,r]}\subseteq T_{\Lambda ^\prime }^{W_0}$ .

We now compare extended quasi-polynomial representations for different choices of lattices $(\Lambda ,\Lambda ^\prime )$ .

Proposition 7.20. Let $(\Lambda _1,\Lambda _1^\prime ), (\Lambda _2,\Lambda _2^\prime )\in (\mathcal {L}^{\times 2})_e$ such that $(\Lambda _1,\Lambda _1^\prime )\leq (\Lambda _2,\Lambda _2^\prime )$ . Let $c\in C^J$ and $\mathfrak {t}\in {}^{\Lambda _2}T_{\Lambda _2^\prime }^c$ .

Then the inclusion map $\mathcal {P}_{\Lambda _1}^{(c)}\hookrightarrow \mathcal {P}_{\Lambda _2}^{(c)}$ defines an embedding

$$\begin{align*}\mathcal{P}_{\Lambda_1,\mathfrak{t}\vert_{\Lambda_1^\prime}}^{(c)}\hookrightarrow\mathcal{P}_{\Lambda_2,\mathfrak{t}}^{(c)} \end{align*}$$

of $\mathbb {H}_{\Lambda _1,\Lambda _1^\prime }$ -modules, where we view $\mathbb {H}_{\Lambda _1,\Lambda _1^\prime }$ as subalgebra of $\mathbb {H}_{\Lambda _2,\Lambda _2^\prime }$ in the natural way.

Proof. We have $\Omega _{\Lambda _1}\subseteq \Omega _{\Lambda _2}$ , and hence, the restriction map $T_{\Lambda _2^\prime }\rightarrow T_{\Lambda _1^\prime }$ , $\mathfrak {t}\mapsto \mathfrak {t}\vert _{\Lambda _1^\prime }$ restricts to a group homomorphism

$$\begin{align*}{}^{\Lambda_2}T_{\Lambda_2^\prime}^c\rightarrow{}^{\Lambda_1}T_{\Lambda_1^\prime}^c. \end{align*}$$

Furthermore, a direct check shows that

$$\begin{align*}\pi^{\Lambda_1,\Lambda_1^\prime}_{c,\mathfrak{t}\vert_{\Lambda_1^\prime}}(h)=\pi^{\Lambda_2,\Lambda_2^\prime}_{c,\mathfrak{t}}(h)\vert_{\mathcal{P}_{\Lambda_1}^{(c)}} \qquad \forall\, h\in\mathbb{H}_{\Lambda_1,\Lambda_1^\prime} \end{align*}$$

for $\mathfrak {t}\in {}^{\Lambda _2}T_{\Lambda _2^\prime }^c$ .

The action of the commuting family of operators $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(Y^\mu )$ ( $\mu \in \Lambda ^\prime $ ) on the cyclic vector $x^c$ is described as follows.

Lemma 7.21. Let $c\in C^J$ and $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ . Then

(7.28) $$ \begin{align} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(Y^\mu)x^c=(\mathfrak{s}_J\mathfrak{t})^{-\mu}x^c \qquad \forall\,\mu\in\Lambda^\prime. \end{align} $$

Proof. By Proposition 7.20 and Theorem 4.5(2), formula (7.28) is correct for $\mu \in Q^{\vee }$ . It thus suffices to prove the lemma for $\mu \in \Lambda _{\text {min}}^\prime $ . Fix $\zeta \in \Lambda _{\text {min}}^\prime $ . Then

$$ \begin{align*} \begin{aligned} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(Y^\zeta)x^c= \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(w_{\zeta,\Lambda^\prime})\pi_{c,\mathfrak{t}}(T_{v_{\zeta,\Lambda^\prime}})x^c &=\kappa_{v_{\zeta,\Lambda^\prime}}(c)\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(w_{\zeta,\Lambda^\prime})x^{v_{\zeta,\Lambda^\prime}c}\\ &= \kappa_{v_{\zeta,\Lambda^\prime}}(c)\big(w_{\zeta,\Lambda^\prime}\big)_{\mathfrak{t};c}(x^{v_{\zeta,\Lambda^\prime}c})= \kappa_{v_{\zeta,\Lambda^\prime}}(c) \mathfrak{t}^{-\zeta}x^c \end{aligned} \end{align*} $$

with the first equality by (7.10) and Proposition 7.20, the second equality by Corollary 5.10 and the fourth equality by (7.22). The result now follows from the fact that

$$\begin{align*}\mathfrak{s}_J^{-\zeta}=\mathfrak{s}_c^{-\zeta}= \prod_{\alpha\in\Phi_0^+} k_\alpha^{-\eta(\alpha(c))\alpha(\zeta)}= \kappa_{v_{\zeta,\Lambda^\prime}}(c), \end{align*}$$

where the last equality follows from (7.4).

Corollary 5.10 now extends as follows.

Proposition 7.22. Let $c\in C^J$ and $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ . Define the surjective linear map $\psi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }: \mathbb {H}_{\Lambda ,\Lambda ^\prime }\twoheadrightarrow \mathcal {P}_\Lambda ^{(c)}$ by

$$\begin{align*}\psi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(x^\lambda T_v Y^{\mu}):= \kappa_v(c)(\mathfrak{s}_J\mathfrak{t})^{-\mu}x^{\lambda+vc} \qquad (\lambda\in\Lambda,\, v\in W_0,\, \mu\in\Lambda^\prime). \end{align*}$$

Then

(7.29) $$ \begin{align} \psi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(h)=\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(h)x^c\qquad \qquad\forall\,h\in\mathbb{H}_{\Lambda,\Lambda^\prime}. \end{align} $$

Proof. First, note that it suffices to prove (7.29) for $h\in H_{\Lambda ^\prime }$ since $\psi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(x^\lambda h^\prime )= x^\lambda \psi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(h^\prime )$ and $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(x^\lambda h)x^c= x^\lambda \pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(h)(x^c)$ for $\lambda \in \Lambda $ and $h^\prime \in H_{\Lambda ^\prime }$ . By a similar argument, it follows from Lemma 7.21 and Theorem 7.18(2) that it suffices to prove (7.29) for $h\in H_0$ . But $\psi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }\vert _{\mathbb {H}}=\psi _{c,\mathfrak {t}\vert _{Q^{\vee }}}$ ; hence, this follows from Corollary 5.10.

Corollary 7.23. Let $c\in \overline {C}_+$ and $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ . Then

$$\begin{align*}\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_\omega^Y)x^c= \kappa_{D\omega}(c)x^{\omega c}\qquad (\omega\in\Omega_\Lambda). \end{align*}$$

In particular, $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(S_\omega ^Y)x^c=x^c$ for $\omega \in \Omega _{\Lambda ,c}$ , and

$$\begin{align*}S_\omega^Y\otimes_{\mathbb{H}_{Q^{\vee},\Lambda^\prime}}f\mapsto \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_\omega^Y)f\qquad (\omega\in\Omega_\Lambda,\, f\in\mathcal{P}^{(c)}) \end{align*}$$

defines an epimorphism $V_{c,\mathfrak {t}}\twoheadrightarrow \mathcal {P}_{\Lambda ,\mathfrak {t}}^{(c)}$ of $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -modules.

Proof. Write $\omega =w_{\zeta ,\Lambda }\in \Omega _\Lambda $ ( $\zeta \in \Lambda _{\text {min}}$ ), Then we have

$$\begin{align*}\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_{w_{\zeta,\Lambda}}^Y)x^c=\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(x^\zeta T_{v_{\zeta,\Lambda}^{-1}})x^c=\kappa_{v_{\zeta,\Lambda}^{-1}}(c)x^{\zeta+v_{\zeta,\Lambda}^{-1}c}= \kappa_{Dw_{\zeta,\Lambda}}(c)x^{w_{\zeta,\Lambda}c}, \end{align*}$$

where we used Proposition 7.22 in the second equality. The last statement now follows from Theorem 7.18 and the proof of Lemma 7.12.

7.4 Twist parameters

The group isomorphism $\Lambda ^\prime /Q^{\vee }\overset {\sim }{\longrightarrow }\Omega _{\Lambda ^\prime }$ from Proposition 7.2(1) allows one to identify ${}^{\Lambda }T_{\Lambda ^\prime }^0=T_{\Lambda ^\prime ,[1,r]}\simeq \text {Hom}(\Lambda ^\prime /Q^{\vee },\mathbf {F}^\times )$ with the character group $\text {Hom}(\Omega _{\Lambda ^\prime },\mathbf {F}^\times )$ of $\Omega _{\Lambda ^\prime }$ . Concretely, the multiplicative character of $\Omega _{\Lambda ^\prime }$ corresponding to $\mathfrak {t}\in T_{\Lambda ^\prime ,[1,r]}$ is given by $w_{\zeta ,\Lambda ^\prime }\mapsto \mathfrak {t}^\zeta $ ( $\zeta \in \Lambda _{\text {min}}^\prime $ ). Note that $w_{c_\mu ,\Lambda ^\prime }\mapsto \mathfrak {t}^\mu $ for $\mu \in \Lambda ^\prime $ .

Definition 7.24. For $\mathfrak {t}\in T_{\Lambda ^\prime ,[1,r]}$ , let $\Xi _{\mathfrak {t}}$ be the algebra automorphism of the extended double affine Hecke algebra $\mathbb {H}_{\Lambda ,\Lambda ^\prime }\simeq \Omega _{\Lambda ^\prime }\ltimes \mathbb {H}_{\Lambda ,Q^{\vee }}$ defined by $h\mapsto h$ ( $h\in \mathbb {H}_{\Lambda ,Q^{\vee }}$ ) and $w_{\zeta ,\Lambda ^\prime }\mapsto \mathfrak {t}^{-\zeta } w_{\zeta ,\Lambda ^\prime }$ ( $\zeta \in \Lambda _{\text {min}}^\prime $ ).

In particular, $\Xi _{\mathfrak {t}}(x^\lambda )=x^\lambda $ and $\Xi _{\mathfrak {t}}(T_j)=T_j$ for $\lambda \in \Lambda $ and $j\in [0,r]$ .

Proposition 7.25. Let $c\in C^J$ , $\mathfrak {t}^\prime \in {}^{\Lambda }T_{\Lambda ^\prime }^c$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,[1,r]}$ . Then

(7.30) $$ \begin{align} \pi_{c,\mathfrak{t}^\prime\mathfrak{t}}^{\Lambda,\Lambda^\prime}=\pi_{c,\mathfrak{t}^\prime}^{\Lambda,\Lambda^\prime}\circ\Xi_{\mathfrak{t}}. \end{align} $$

Furthermore, $\Xi _{\mathfrak {t}}(Y^\mu )=\mathfrak {t}^{-\mu }Y^\mu $ for all $\mu \in \Lambda ^\prime $ .

Proof. Fix $\mathfrak {t}^\prime \in {}^{\Lambda }T_{\Lambda ^\prime }^c$ and $\mathfrak {t}\in T_{\Lambda ^\prime ,[1,r]}$ . Then $\mathfrak {t}^\prime \mathfrak {t}\in {}^{\Lambda ^\prime }T_{\Lambda }^c$ by Example 7.14(3), and

(7.31) $$ \begin{align} (\mathfrak{t}^\prime\mathfrak{t})_{y;c}=\mathfrak{t}^\prime_{y;c}\mathfrak{t}\qquad \forall\, y\in\mathcal{O}_{\Lambda,c}. \end{align} $$

Indeed, if $g\in W_{\Lambda }$ is such that $y=gc$ , then $(\mathfrak {t}^\prime \mathfrak {t})_{y;c}=g(\mathfrak {t}^\prime \mathfrak {t})=(g\mathfrak {t}^\prime )((Dg)\mathfrak {t})=\mathfrak {t}^\prime _{y;c}\mathfrak {t}$ since $Dg\in W_0$ and $\mathfrak {t}\in T_{\Lambda ^\prime }^{W_0}$ .

Let $\Theta _{\mathfrak {t}}: W_{\Lambda ^\prime }\rightarrow \mathbf {F}^\times $ be the group homomorphism defined by $w\mapsto 1$ ( $w\in W$ ) and $w_{\zeta ,\Lambda ^\prime }\mapsto \mathfrak {t}^{-\zeta }$ ( $\zeta \in \Lambda ^\prime _{\text {min}}$ ). Since $\tau (\mu )\in w_{c_\mu ,\Lambda ^\prime }W$ for all $\mu \in \Lambda ^\prime $ , we have

$$\begin{align*}\Theta_{\mathfrak{t}}(\tau(\mu))=\mathfrak{t}^{-\mu}\qquad\quad \forall\, \mu\in\Lambda^\prime. \end{align*}$$

We claim that

(7.32) $$ \begin{align} w_{\mathfrak{t}^\prime\mathfrak{t};c}(x^y)=\Theta_{\mathfrak{t}}(w)w_{\mathfrak{t}^\prime;c}(x^y) \end{align} $$

for $w\in W_{\Lambda ^\prime }$ and $y\in \mathcal {O}_{\Lambda ,c}$ . Formula (7.32) is trivial for $w\in W_0$ . If $w=\tau (\mu )\in W$ ( $\mu \in \Lambda ^\prime $ ), then it follows from (7.31) that

$$\begin{align*}\tau(\mu)_{\mathfrak{t}^\prime\mathfrak{t};c}(x^y)=(\mathfrak{t}^\prime_{y;c}\mathfrak{t})^{-\mu}x^y=\Theta_{\mathfrak{t}}(\tau(\mu))\tau(\mu)_{\mathfrak{t}^\prime;c}(x^y) \end{align*}$$

for $y\in \mathcal {O}_{\Lambda ,c}$ , which completes the proof of (7.32).

Formula (7.30) follows immediately from (7.32). Finally, we have $\Xi _{\mathfrak {t}}(Y^\mu )=\mathfrak {t}^{-\mu }Y^\mu $ for $\mu \in \Lambda ^\prime $ since $Y^\mu \in w_{c_\mu ,\Lambda ^\prime }H$ by (7.9).

Proposition 7.25 implies, under suitable generic conditions on $\mathfrak {t}\in {}^\Lambda T_{\Lambda ^\prime }^c$ , that the simultaneous eigenfunctions of the commuting operators $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(Y^\mu )$ ( $\mu \in \Lambda ^\prime $ ) will only depend on the coset $\mathfrak {t}T_{\Lambda ^\prime ,[1,r]}$ inside ${}^\Lambda T_{\Lambda ^\prime }^c$ . The precise statement will be given in Theorem 7.26(3).

It follows from Remark 7.19 and Proposition 7.25 that for $\zeta \in \Lambda _{\text {min}}$ and $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^\zeta =q^{-\zeta }T_{\Lambda ^\prime ,[1,r]}$ , the extended quasi-polynomial representation

$$\begin{align*}\pi_{\zeta,\mathfrak{t}}^{\Lambda,\Lambda^\prime}: \mathbb{H}_{\Lambda,\Lambda^\prime}\rightarrow\text{End}(\mathcal{P}_\Lambda) \end{align*}$$

is the twisted version $\pi _{0,1_{T_{\Lambda ^\prime }}}^{\Lambda ,\Lambda ^\prime }\circ \Xi _{q^{-\zeta }\mathfrak {t}}$ of the Cherednik representation $\pi _{0,1_{T_{\Lambda ^\prime }}}^{\Lambda ,\Lambda ^\prime }$ . The corresponding commuting Y-operators are related by the formula

$$\begin{align*}\pi_{\zeta,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(Y^\mu)=q^{\langle\zeta,\mu\rangle}\mathfrak{t}^{-\mu}\,\pi_{0,1_{T_{\Lambda^\prime}}}^{\Lambda,\Lambda^\prime}(Y^\mu) \qquad\quad\forall\,\mu\in\Lambda^\prime. \end{align*}$$

7.5 The extended eigenvalue equations

In this subsection, we consider the monic quasi-polynomial simultaneous eigenfunctions of $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(Y^\mu )$ ( $\mu \in \Lambda ^\prime $ ).

Theorem 7.26. Let $c\in C^J$ , and let $\mathcal {O}$ be a W-orbit in $\mathcal {O}_{\Lambda ,c}$ . Let $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ such that

(7.33) $$ \begin{align} (\mathfrak{s}_y\mathfrak{t}_{y;c})\vert_{\Lambda^\prime}\not=(\mathfrak{s}_{y^\prime}\mathfrak{t}_{y^\prime;c})\vert_{\Lambda^\prime}\quad\mbox{ when }\,\, y,y^\prime\in \mathcal{O}\,\, \mbox{ and }\,\, y\not=y^\prime. \end{align} $$

For all $y\in \mathcal {O}$ , we have

  1. 1. There exists a unique simultaneous eigenfunction

    $$\begin{align*}E_{y;c}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t})\in\mathcal{P}^{(c)}_\Lambda \end{align*}$$
    of the commuting operators $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }(Y^\mu )$ ( $\mu \in \Lambda ^\prime $ ) satisfying
    (7.34) $$ \begin{align} E_{y;c}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t})=x^{y}+\text{l.o.t.} \end{align} $$
  2. 2. We have

    (7.35) $$ \begin{align} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(Y^\mu)E_{y;c}^{\Lambda,\Lambda^\prime}(\cdot;\mathfrak{t})= (\mathfrak{s}_y\mathfrak{t}_{y;c})^{-\mu}E_{y;c}^{\Lambda,\Lambda^\prime}(\cdot;\mathfrak{t})\qquad\quad\forall\, \mu\in\Lambda^\prime. \end{align} $$
  3. 3. If $\mathfrak {t}$ satisfies the more stringent genericity conditions

    (7.36) $$ \begin{align} (\mathfrak{s}_y\mathfrak{t}_{y;c})\vert_{Q^{\vee}}\not=(\mathfrak{s}_{y^\prime}\mathfrak{t}_{y^\prime;c})\vert_{Q^{\vee}}\quad\mbox{ when }\,\, y,y^\prime\in \mathcal{O}\,\, \mbox{ and }\,\, y\not=y^\prime \end{align} $$
    then
    (7.37) $$ \begin{align} E_{y;c}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t})=E_y^{\omega(J)}(x;(\omega\mathfrak{t})\vert_{Q^{\vee}}), \end{align} $$
    where $\omega \in \Omega _\Lambda $ is such that $\mathcal {O}=\mathcal {O}_{\omega c}$ .

Proof. (1)&(2). It suffices to prove that

(7.38) $$ \begin{align} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(Y^\mu)x^y=(\mathfrak{s}_y\mathfrak{t}_{y;c})^{-\mu}x^y+\text{l.o.t.}\qquad\quad (\mu\in\Lambda^\prime). \end{align} $$

By Proposition 7.20, Theorem 7.18(2), (7.20) and Proposition 5.28, this is correct for $\mu \in Q^{\vee }$ . A straightforward extension of the proof of Proposition 5.28 shows that it holds true for all $\mu \in \Lambda ^\prime $ .

(3) The extra assumption on $\mathfrak {t}$ implies that $(\omega \mathfrak {t})\vert _{Q^{\vee }}\in T_{\omega (J)}^\prime $ ; hence, $E_y^{\omega (J)}\big (x;(\omega \mathfrak {t})\vert _{Q^{\vee }}\big )$ is well defined. The result is now an immediate consequence of Theorem 7.18(2) and Theorem 6.3(1).

As a special case of Theorem 7.26(3), we have for $c\in C^J$ and $\mathfrak {t}\in T_J^\prime $ ,

$$\begin{align*}E_y^J(x;\mathfrak{t})=E_{y;c}^{Q^{\vee},Q^{\vee}}(x;\mathfrak{t})\qquad\quad (y\in\mathcal{O}_c). \end{align*}$$

The index c is to indicate that $E_{y;c}^{\Lambda ,\Lambda ^\prime }(x;\mathfrak {t})$ depends on the choice of the representative of $\mathcal {O}_{\Lambda ,c}$ inside $\overline {C}_+$ . The quasi-polynomials for different choices of representatives are related as follows.

Corollary 7.27. Let $c\in \overline {C}_+$ and choose a W-orbit $\mathcal {O}$ in $\mathcal {O}_{\Lambda ,c}$ . Fix $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ satisfying (7.33). Then $\omega \mathfrak {t}\in T_{\Lambda ^\prime }$ lies in $\omega \mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^{\omega c}$ and satisfies (7.33), and

(7.39) $$ \begin{align} E_{y;\omega c}^{\Lambda,\Lambda^\prime}(x;\omega\mathfrak{t})=E_{y;c}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t})\qquad \forall\, y\in\mathcal{O}. \end{align} $$

Proof. This follows immediately from the characterisation of $E_{y;c}^{\Lambda ,\Lambda ^\prime }(x;\mathfrak {t})$ as given in Theorem 7.26(1).

The following is an extension of Theorem 7.26(3).

Corollary 7.28. Let $(\Lambda _1,\Lambda _1^\prime ), (\Lambda _2,\Lambda _2^\prime )\in (\mathcal {L}^{\times 2})_e$ such that $(\Lambda _1,\Lambda _1^\prime )\leq (\Lambda _2,\Lambda _2^\prime )$ . Let $c\in \overline {C}_+$ and choose a W-orbit $\mathcal {O}$ in $\mathcal {O}_{\Lambda _1,c}$ . Fix $\mathfrak {t}\in {}^{\Lambda _2}T_{\Lambda _2^\prime }^c$ such that (7.33) holds true for $\Lambda ^\prime =\Lambda _1^\prime $ . Then for all $y\in \mathcal {O}$ ,

(7.40) $$ \begin{align} E_{y;c}^{\Lambda_1,\Lambda_1^\prime}(x;\mathfrak{t}\vert_{\Lambda_1^\prime})=E_{y;c}^{\Lambda_2,\Lambda_2^\prime}(x;\mathfrak{t}). \end{align} $$

In particular, the quasi-polynomial $E_{y;c}^{\Lambda _1,\Lambda _1^\prime }(x;\mathfrak {t}\vert _{\Lambda _1^\prime })$ ( $y\in \mathcal {O}$ ) satisfies the extended eigenvalue equations

(7.41) $$ \begin{align} \pi_{c,\mathfrak{t}}^{\Lambda_2,\Lambda_2^\prime}(Y^\mu)E_{y;c}^{\Lambda_1,\Lambda_1^\prime}(\cdot;\mathfrak{t}\vert_{\Lambda_1^\prime})= (\mathfrak{s}_y\mathfrak{t}_{y;c})^{-\mu}E_{y;c}^{\Lambda_1,\Lambda_1^\prime}(\cdot;\mathfrak{t}\vert_{\Lambda_1^\prime})\qquad\quad\forall\, \mu\in\Lambda_2^\prime. \end{align} $$

Proof. Note that (7.33) also holds true for $\Lambda ^\prime =\Lambda _2^\prime $ since $\Lambda _1^\prime \subseteq \Lambda _2^\prime $ . Hence, both sides of (7.40) are well defined. The result now follows from Theorem 7.26(1) and Proposition 7.20.

Remark 7.29. Let $\zeta \in \Lambda \cap E_{\text {co}}$ and impose the assumptions of Theorem 7.26(1). Then (7.33) also holds true for $\tau (\zeta )\mathcal {O}$ since $\mathfrak {s}_{y+\zeta }=\mathfrak {s}_y$ , and

$$ \begin{align*} E_{y+\zeta;c}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t})= x^\zeta E_{y;c}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t})\qquad\quad (\zeta\in\Lambda\cap E_{\text{co}}) \end{align*} $$

for all $y\in \mathcal {O}$ (compare with Lemma 6.6(2) and its proof).

Under appropriate generic conditions on $\mathfrak {t}$ , one can now transport the results on the quasi-polynomial eigenfunctions $E_y^J(x;\mathfrak {t})$ from Section 6.1 to $E_{y;c}^{\Lambda ,\Lambda ^\prime }(x;\mathfrak {t})$ using Theorem 7.26(3). We finish this subsection by giving an explicit example.

For an element $w\in W_\Lambda $ , define $d_w\in \mathcal {P}$ and $k_w(y)$ by (6.6) and (5.7), respectively.

Proposition 7.30. Let $c\in C^J$ and choose a W-orbit $\mathcal {O}$ in $\mathcal {O}_{\Lambda ,c}$ . Let $\omega \in \Omega _\Lambda $ such that $\mathcal {O}=\mathcal {O}_{\omega c}$ . Fix $\mathfrak {t}\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ satisfying (7.36) and suppose that $\mathfrak {t}$ is J-regular. Then

  1. 1. For $y\in \mathcal {O}$ , we have

    $$\begin{align*}E_{y;c}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t})= d_{w_y\omega}(\mathfrak{s}_J\mathfrak{t})^{-1}k_{w_y\omega}(c)^{-1} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_{w_y\omega}^Y)x^c. \end{align*}$$
  2. 2. Assume that (7.36) also holds true for the W-orbit $\omega ^{-1}\mathcal {O}=\mathcal {O}_c$ in $\mathcal {O}_{\Lambda ,c}$ . Then

    $$\begin{align*}E_{\omega y;c}^{\Lambda,\Lambda^\prime}(\cdot;\mathfrak{t})=k_\omega(y)^{-1}\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_\omega^Y)E_y^J(\cdot;\mathfrak{t}\vert_{Q^{\vee}})\qquad\quad (y\in\mathcal{O}_c). \end{align*}$$

Proof. (1) Fix $y\in \mathcal {O}_{\omega c}$ . Then Corollary 7.23, (7.26) and Proposition 7.20 give

$$\begin{align*}\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_{w_y\omega}^Y)x^c= \kappa_{D\omega}(c)\pi_{\omega c,\omega\mathfrak{t}}(S_{w_y}^Y)x^{\omega c}. \end{align*}$$

Applying Theorem 6.13, we get

$$\begin{align*}\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_{w_y\omega}^Y)x^c= k_{w_y}(\omega c)\kappa_{D\omega}(c)d_{w_y}(\mathfrak{s}_{\omega(J)}\omega\mathfrak{t}) E_y^{\omega(J)}(x;(\omega\mathfrak{t})\vert_{Q^{\vee}}), \end{align*}$$

so by (7.37), it suffices to show that

(7.42) $$ \begin{align} \begin{aligned} d_{w_y}(\mathfrak{s}_{\omega(J)}\omega\mathfrak{t})&=d_{w_y\omega}(\mathfrak{s}_c\mathfrak{t}),\\ k_{w_y}(\omega c)\kappa_{D\omega}(c)&=k_{w_y\omega}(c). \end{aligned} \end{align} $$

For the first equation of (7.42), note that by Lemma 7.9, we have

$$\begin{align*}d_{w_y}(\mathfrak{s}_{\omega(J)}\omega\mathfrak{t})=d_{w_y}(\omega(\mathfrak{s}_{J}\mathfrak{t}))= (\omega^{-1}d_{w_y})(\mathfrak{s}_{J}\mathfrak{t}). \end{align*}$$

But $\Pi (w_y\omega )=\omega ^{-1}\Pi (w_y)$ since $\omega \in \Omega _\Lambda $ ; hence, $\omega ^{-1}d_{w_y}=d_{w_y\omega }$ in $\mathcal {P}$ . For the second equation, it suffices to show that

(7.43) $$ \begin{align} k_\omega(y)=\kappa_{D\omega}(y). \end{align} $$

Write $\omega =w_{\zeta ,\Lambda }$ ( $\zeta \in \Lambda _{\text {min}}$ ). Then we compute for $y\in E$ ,

$$ \begin{align*} \begin{aligned} k(w_{\zeta,\Lambda}y)&=\prod_{\alpha\in\Phi_0^+}k_\alpha^{\eta(\alpha(v_{\zeta,\Lambda}^{-1}y+\zeta))/2}\\ &=\prod_{\alpha\in\Pi(v_{\zeta,\Lambda})}k_\alpha^{\eta(\alpha(v_{\zeta,\Lambda}^{-1}y)+1)/2} \prod_{\beta\in\Phi_0^+\setminus\Pi(v_{\zeta,\Lambda})}k_\beta^{\eta(\beta(v_{\zeta,\Lambda}^{-1}y))/2}\\ &=\prod_{\alpha\in\Pi(v_{\zeta,\Lambda})}k_\alpha^{-\eta(-\alpha(v_{\zeta,\Lambda}^{-1}y))/2} \prod_{\beta\in\Phi_0^+\setminus\Pi(v_{\zeta,\Lambda}^{-1})}k_\beta^{\eta(\beta(y))/2}\\ &=\prod_{\alpha\in\Pi(v_{\zeta,\Lambda}^{-1})}k_\alpha^{-\eta(\alpha(y))/2} \prod_{\beta\in\Phi_0^+\setminus\Pi(v_{\zeta,\Lambda}^{-1})}k_\beta^{\eta(\beta(y))/2}= k(y)\kappa_{v_{\zeta,\Lambda}^{-1}}(y), \end{aligned} \end{align*} $$

where we have used (7.4) in the second equality, (5.1) and

$$\begin{align*}v_{\zeta,\Lambda}(\Phi_0^+\setminus\Pi(v_{\zeta,\Lambda}))=\Phi_0^+\setminus\Pi(v_{\zeta,\Lambda}^{-1}) \end{align*}$$

in the third equality, and $v_{\zeta ,\Lambda }(\Pi (v_{\zeta ,\Lambda }))=-\Pi (v_{\zeta ,\Lambda }^{-1})$ in the fourth equality. Now (7.43) follows, since $Dw_{\zeta ,\Lambda }=v_{\zeta ,\Lambda }^{-1}$ .

(2) Let $\omega \in \Omega _\Lambda $ and $y\in \mathcal {O}_c$ . Then $\omega w_{y}\omega ^{-1}=w_{\omega y}$ , Proposition 6.13 and part (1) of the proposition give

$$ \begin{align*} \begin{aligned} \pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_\omega^Y)E_y^J(\cdot;\mathfrak{t}\vert_{Q^{\vee}})&= d_{w_y}(\mathfrak{s}_J\mathfrak{t})^{-1} k_{w_y}(c)^{-1}\pi_{c,\mathfrak{t}}^{\Lambda,\Lambda^\prime}(S_{w_{\omega y}\omega}^Y)x^c\\ &=\frac{d_{\omega w_y}(\mathfrak{s}_J\mathfrak{t})k_{\omega w_y}(c)} {d_{w_y}(\mathfrak{s}_J\mathfrak{t})k_{w_y}(c)} E_{\omega y;c}^{\Lambda,\Lambda^\prime}(\cdot;\mathfrak{t}). \end{aligned} \end{align*} $$

Note that $d_{\omega w}=d_w$ ( $\omega \in \Omega _\Lambda $ , $w\in W_\Lambda $ ) in $\mathcal {P}$ since $\Pi (\omega w)=\Pi (w)$ . The result then follows from the observation that

$$\begin{align*}\frac{k_{\omega w_y}(c)}{k_{w_y}(c)}=\frac{k(\omega w_yc)}{k(w_yc)}= \frac{k(\omega y)}{k(y)}=k_\omega(y). \end{align*}$$

7.6 The theory for the $\text {GL}_{r+1}$ root datum

In this subsection, we make the results of the previous two subsections explicit for the root datum associated to $\text {GL}_{r+1}$ .

Let $\{\epsilon _i\}_{i=1}^{r+1}$ be the standard orthonormal basis of $\mathbb {R}^{r+1}$ , and fix a positive integer $\ell \in \mathbb {Z}_{>0}$ . Then

$$\begin{align*}\Phi_0:=\big\{(\epsilon_i-\epsilon_j)/\ell\,\,\,\, | \,\,\,\, 1\leq i\not=j\leq r+1\big\}\subset\mathbb{R}^{r+1} \end{align*}$$

is a root system of type $A_r$ , with roots having squared norm $2/\ell ^2$ (this particular normalisation of the basis will be convenient for the metaplectic theory; see Sections 9 and 10). Take $\alpha _i:=(\epsilon _i-\epsilon _{i+1})/\ell $ ( $1\leq i\leq r$ ) as the simple roots of $\Phi _0$ . Then $\varphi =(\epsilon _1-\epsilon _{r+1})/\ell $ , and hence,

$$\begin{align*}\alpha_0=\big((\epsilon_{r+1}-\epsilon_1)/\ell,1\big). \end{align*}$$

The alcove $\overline {C}_+$ consists of the vectors $y=(y_1,\ldots ,y_{r+1})\in \mathbb {R}^{r+1}$ satisfying

(7.44) $$ \begin{align} y_{r+1}\leq y_r\leq\cdots \leq y_1\leq y_{r+1}+\ell. \end{align} $$

The face $C^J$ consists of the vectors $y\in \overline {C}_+$ such that

  1. 1. $y_{i+1}=y_{i}$ if $i\in J_0$ ,

  2. 2. $y_1=y_{r+1}+\ell $ if $0\in J$ ,

  3. 3. the remaining inequalities in (7.44) are strict.

We now take $(\Lambda , \Lambda ^\prime ):=(\ell \mathbb {Z}^{r+1},\ell \mathbb {Z}^{r+1})\in (\mathcal {L}^{\times 2})_1$ , with $\ell \mathbb {Z}^{r+1}$ the lattice generated by the basis $\{\ell \epsilon _i\}_{i=1}^{r+1}$ of $\mathbb {R}^{r+1}$ . The abelian group $\Omega _{\ell \mathbb {Z}^{r+1}}$ is a free group of rank $1$ generated by

$$\begin{align*}u:=w_{\ell\epsilon_1,\ell\mathbb{Z}^{r+1}}=\tau(\ell\epsilon_1)s_1\cdots s_r. \end{align*}$$

Then $v_{\ell \epsilon _1,\ell \mathbb {Z}^{r+1}}=s_r\cdots s_2s_1$ and $u\alpha _j=\alpha _{j+1}$ ( $0\leq j\leq r$ ) by (2.3), where the indices are read modulo $r+1$ (this will be done throughout this example).

Note that $(\mathbb {R}^{r+1})_{\text {co}}= \mathbb {R}\varpi $ and $(\mathbb {R}^{r+1})_{\text {co}}\cap \ell \mathbb {Z}^{r+1}= \mathbb {Z}\varpi $ with

$$\begin{align*}\varpi:=\ell(\epsilon_1+\cdots+\epsilon_{r+1}). \end{align*}$$

Furthermore, $I_{\ell \mathbb {Z}^{r+1}}=[1,r]$ (see (7.5)). A complete set of representatives of the $(\mathbb {R}^{r+1})_{\text {co}}\cap \ell \mathbb {Z}^{r+1}$ -orbits inside $(\ell \mathbb {Z}^{r+1})_{\text {min}}$ is $\{0\}\cup \{\varpi _{i,\ell \mathbb {Z}^{r+1}}^{\vee }\}_{i=1}^r$ with

$$\begin{align*}\varpi_{i,\ell\mathbb{Z}^{r+1}}^{\vee}:=\ell(\epsilon_1+\cdots+\epsilon_i)\qquad (1\leq i\leq r). \end{align*}$$

For $s,i\in \mathbb {Z}$ , we have

$$\begin{align*}u^{s(r+1)+i}=\tau(s\varpi)u^i \end{align*}$$

and

$$\begin{align*}u^i=w_{\varpi_{i,\ell\mathbb{Z}^{r+1}}^{\vee}}\quad (1\leq i\leq r) \end{align*}$$

(see, for example, [Reference Stokman, Koornwinder and Stokman44, §9.3.5]).

Lemma 7.31. We have

(7.45) $$ \begin{align} \Omega_{\ell\mathbb{Z}^{r+1},c}=\{1\}\qquad \forall\, c\in \overline{C}_+. \end{align} $$

In particular, ${}^{\ell \mathbb {Z}^{r+1}}T_{\ell \mathbb {Z}^{r+1}}^c=T_{\ell \mathbb {Z}^{r+1},J}$ for $c\in C^J$ and

$$\begin{align*}\mathcal{O}_{\ell\mathbb{Z}^{r+1},c}=\bigsqcup_{(s,i)\in\mathbb{Z}\times [0,r]}\mathcal{O}_{u^ic+s\varpi}. \end{align*}$$

Proof. If $u^s\in \Omega _{\ell \mathbb {Z}^{r+1},c}$ , then $c=u^{s(r+1)}c=c+s\varpi $ ; hence, $s=0$ .

For $c\in C^J$ and $\mathfrak {t}\in T_{\ell \mathbb {Z}^{r+1},J}$ , the $A_r$ -type quasi-polynomial representation $\pi _{c,\mathfrak {t}}: \mathbb {H}\rightarrow \text {End}(\mathcal {P}^{(c)})$ can thus be extended to a representation

$$\begin{align*}\widetilde{\pi}_{c,\mathfrak{t}}:=\pi_{c,\mathfrak{t}}^{\ell\mathbb{Z}^{r+1},\ell\mathbb{Z}^{r+1}}:\widetilde{\mathbb{H}}\rightarrow\text{End}(\mathcal{P}_{\ell\mathbb{Z}^{r+1}}^{(c)}) \end{align*}$$

of the double affine Hecke algebra

(7.46) $$ \begin{align} \widetilde{\mathbb{H}}:=\mathbb{H}_{\ell\mathbb{Z}^{r+1},\ell\mathbb{Z}^{r+1}} \end{align} $$

of type $\text {GL}_{r+1}$ . Note that $\widetilde {\mathbb {H}}$ depends on a single multiplicity parameter $k\in \mathbf {F}^\times $ . Algebraic generators of $\widetilde {\mathbb {H}}$ are $x^\lambda $ ( $\lambda \in \ell \mathbb {Z}^{r+1}$ ), $T_i$ ( $1\leq i\leq r$ ) and $u^{\pm 1}$ (note that $T_0=uT_ru^{-1}$ ). The corresponding commuting elements $Y^\mu \in \widetilde {\mathbb {H}}$ for $\mu =\ell \sum _{i=1}^{r+1}m_i\epsilon _i\in \ell \mathbb {Z}^{r+1}$ ( $m_i\in \mathbb {Z}$ ) are explicitly given by

$$\begin{align*}Y^\mu=(Y^{\ell\epsilon_1})^{m_1}\cdots \big(Y^{\ell\epsilon_{r+1}}\big)^{m_{r+1}} \end{align*}$$

with

(7.47) $$ \begin{align} Y^{\ell\epsilon_i}:=T_{i-1}^{-1}\cdots T_2^{-1}T_1^{-1}uT_r\cdots T_{i+1}T_i \qquad (1\leq i\leq r+1). \end{align} $$

For $i=r+1$ , this should be read as $Y^{\ell \epsilon _{r+1}}=T_r^{-1}\cdots T_2^{-1}T_1^{-1}u$ .

For $c\in C^J$ and $\mathfrak {t}\in T_{\ell \mathbb {Z}^{r+1},J}$ , the extended quasi-polynomial representation

$$\begin{align*}\widetilde{\pi}_{c,\mathfrak{t}}: \widetilde{\mathbb{H}}\rightarrow\text{End}(\mathcal{P}_{\ell\mathbb{Z}^{r+1}}^{(c)}) \end{align*}$$

is then characterised by the formulas

$$ \begin{align*} \begin{aligned} \widetilde{\pi}_{c,\mathfrak{t}}(T_i)x^y&=k_i^{\chi_{\mathbb{Z}}(\alpha_i(y))}x^{s_iy}+(k_i-k_i^{-1})\nabla_i(x^y),\\ \widetilde{\pi}_{c,\mathfrak{t}}(u)x^y&=\mathfrak{t}_{y;c}^{-\ell\epsilon_{r+1}}x^{s_1\cdots s_ry},\\ \widetilde{\pi}_{c,\mathfrak{t}}(x^\lambda)x^y&=x^{y+\lambda} \end{aligned} \end{align*} $$

for $i\in [1,r]$ , $\lambda \in \ell \mathbb {Z}^{r+1}$ and $y\in \mathcal {O}_{\ell \mathbb {Z}^{r+1},c}$ . For the corresponding monic quasi-polynomial eigenfunctions

$$\begin{align*}\widetilde{E}_{y;c}(x;\mathfrak{t}):=E_{y;c}^{\ell\mathbb{Z}^{r+1},\ell\mathbb{Z}^{r+1}}(x;\mathfrak{t}) \end{align*}$$

of the $\widetilde {\pi }_{c,\mathfrak {t}}(Y^\mu )$ ( $\mu \in \ell \mathbb {Z}^{r+1}$ ), we then have the following result.

Corollary 7.32. Let $c\in C^J$ and $\mathfrak {t}\in T_{\ell \mathbb {Z}^{r+1},J}$ . Choose a W-orbit $\mathcal {O}$ in $\mathcal {O}_{\ell \mathbb {Z}^{r+1},c}$ and assume that the generic conditions (7.33) hold true for $\Lambda ^\prime =\ell \mathbb {Z}^{r+1}$ .

  1. 1. We have

    (7.48) $$ \begin{align} \widetilde{E}_{y+s\varpi;c}(x;\mathfrak{t})=x^{s\varpi}\widetilde{E}_{y;c}(x;\mathfrak{t}) \end{align} $$
    for $y\in \mathcal {O}$ and $s\in \mathbb {Z}$ .
  2. 2. If $\mathcal {O}=\mathcal {O}_{u^ic}$ ( $i\in \mathbb {Z}$ ) and $\mathfrak {t}$ satisfies the genericity conditions (7.36), then

    (7.49) $$ \begin{align} \widetilde{E}_{y;c}(x;\mathfrak{t})=E_y^{u^i(J)}(x;(u^i\mathfrak{t})\vert_{Q^{\vee}})\qquad\quad \forall\,y\in\mathcal{O}_{u^ic}. \end{align} $$

Proof. (1) This is a special case of Remark 7.29.

(2) This is a special case of Theorem 7.26(3).

Remark 7.33. Writing $i=s(r+1)+j$ in (7.49) with $s\in \mathbb {Z}$ and $j\in [1,r]$ , we have $u^i(J)=u^j(J)$ and $(u^i\mathfrak {t})\vert _{Q^{\vee }}=(q^{s\varpi }u^j\mathfrak {t})\vert _{Q^{\vee }}=(u^j\mathfrak {t})\vert _{Q^{\vee }}$ , and so

$$\begin{align*}\widetilde{E}_{y;c}(x;\mathfrak{t})=E_y^{u^j(J)}(x;(u^j\mathfrak{t})\vert_{Q^{\vee}})\qquad \forall\, y\in\mathcal{O}_{u^ic}. \end{align*}$$

This is in agreement with (7.48) in view of Lemma 6.6(2).

We finish this subsection by giving explicit descriptions of $T_{\ell \mathbb {Z}^{r+1},J}$ and $T_{\ell \mathbb {Z}^{r+1},J}^{\text {red}}$ .

Identifying $T_{\ell \mathbb {Z}^{r+1}}$ with $(\mathbf {F}^\times )^{r+1}$ by

$$\begin{align*}T_{\ell\mathbb{Z}^{r+1}}\overset{\sim}{\longrightarrow} (\mathbf{F}^\times)^{r+1},\qquad \mathfrak{t}\mapsto (\mathfrak{t}_1,\ldots,\mathfrak{t}_{r+1}) \end{align*}$$

with $\mathfrak {t}_i:=\mathfrak {t}^{\ell \epsilon _i}$ ( $1\leq i\leq r+1$ ), we have $\mathfrak {t}^{\alpha _0^{\vee }}=q^{\ell ^2}\mathfrak {t}_{r+1}\mathfrak {t}_1^{-1}$ and $\mathfrak {t}^{\alpha _i^{\vee }}=\mathfrak {t}_i\mathfrak {t}_{i+1}^{-1}$ ( $1\leq i\leq r$ ). Furthermore, u then acts on $(\mathbf {F}^\times )^{r+1}$ by

$$\begin{align*}u\mathfrak{t}=(q^{\ell^2}\mathfrak{t}_{r+1},\mathfrak{t}_1,\ldots,\mathfrak{t}_r). \end{align*}$$

Combined with (7.45), we then conclude that for $c\in C^J$ ,

$$\begin{align*}{}^{\ell\mathbb{Z}^{r+1}}T_{\ell\mathbb{Z}^{r+1}}^c=T_{\ell\mathbb{Z}^{r+1},J} \simeq\{\mathfrak{t}\in (\mathbf{F}^\times)^{r+1}\,\, | \,\, \mathfrak{t}_j=\mathfrak{t}_{j+1}\,\, (j\in J_0)\, \mbox{ and } \mathfrak{t}_1=q^{\ell^2}\mathfrak{t}_{r+1}\, \mbox{ if }\, 0\in J\}. \end{align*}$$

An explicit description of $T_{\ell \mathbb {Z}^{r+1},J}^{\text {red}}$ is as follows. Let $J\subsetneq [0,r]$ , which we regard as subset of $[1,r+1]$ by identifying $0\equiv r+1$ . Then $\mathfrak {t}\in T_{\ell \mathbb {Z}^{r+1}}$ lies in $T_{\ell \mathbb {Z}^{r+1},J}^{\text {red}}$ if and only if

$$\begin{align*}\mathfrak{t}_i=\mathfrak{t}_{i+1}=\cdots=\mathfrak{t}_j=\mathfrak{t}_{j+1} \end{align*}$$

for all subintervals $[i,j]$ of $[1,r+1]$ such that $i-1,j+1\not \in J$ (here, the numbers are read modulo $r+1$ again). In particular, $T_{\ell \mathbb {Z}^{r+1},J}^{\text {red}}$ is a subtorus of $T_{\ell \mathbb {Z}^{r+1}}$ of dimension $\#([0,r]\setminus J)$ ; cf. Remark 6.2 in case of adjoint root datum.

In the case $J=[1,r]$ corresponding to the usual Macdonald polynomials, one thus has

$$\begin{align*}T_{\ell\mathbb{Z}^{r+1},[1,r]}=T_{\ell\mathbb{Z}^{r+1},[1,r]}^{\text{red}}=\{z^\varpi\,\, | \,\, z\in\mathbf{F}^\times\}. \end{align*}$$

8 Uniform quasi-polynomial representations

In this section, we re-parametrise the quasi-polynomial representation in a way that allows us to naturally glue them together to form a uniform quasi-polynomial representation on $\mathbf {F}[E]$ . The part of the re-parametrisation which does not involve twisting will be given by a map

$$\begin{align*}\widehat{\mathcal{G}}\rightarrow\{E\rightarrow T_{\Lambda^\prime}\},\qquad \widehat{\mathbf{g}}\mapsto \{y\mapsto\mathfrak{t}_y(\widehat{\mathbf{g}})\} \end{align*}$$

satisfying $\mathfrak {t}_c(\widehat {\mathbf {g}})\in \widetilde {T}_{\Lambda ^\prime ,\mathbf {J}(c)}$ (see (3.5)) for all $c\in \overline {C}_+$ . The resulting uniform quasi-polynomial representation of $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ will be a gauged version of

$$\begin{align*}\bigoplus_{c\in\overline{C}_+}\pi_{c,\mathfrak{t}_c(\widehat{\mathbf{g}})}^{Q^{\vee},\Lambda^\prime}: \mathbb{H}_{Q^{\vee},\Lambda^\prime}\rightarrow\text{End}(\mathbf{F}[E]), \end{align*}$$

which now depends on q, $\mathbf {k}$ and on the parametrising set $\widehat {\mathcal {G}}$ of the different choices of uniformizations of the induction parameters of the quasi-polynomial representations.

Throughout this section, we fix $e\in \mathbb {Z}_{>0}$ and a pair $(\Lambda ,\Lambda ^\prime )\in (\mathcal {L}^{\times 2})_e$ . We assume throughout this section that $\mathbf {F}$ contains a $\text {lcm}(e,2h)^{\text {th}}$ root $q^{\frac {1}{\text {lcm}(e,2h)}}$ of q, which we fix once and for all.

We often do not record the dependence on $\Lambda ^\prime $ when dealing with multiplicative characters t of $\Lambda ^\prime $ defined as products of co-character values $z^\alpha $ ( $z\in \mathbf {F}^\times $ , $\alpha \in \Phi _0$ ) since its interpretation as multiplicative character on $\Lambda ^\prime $ is unambiguous (for an example of such a multiplicative character, see (3.12)). If it is important to specify the lattice, then we write $t\vert _{\Lambda ^\prime }$ .

8.1 The g-parameters

The re-parametrisation of the quasi-polynomial representation is based on factorisations of multiplicative characters in terms of co-characters. An important example is the following factorisation of the multiplicative character $q^{\text {pr}_{E^\prime }(\lambda )}\in T_{\Lambda ^\prime }$ for $\lambda \in \Lambda $ .

Lemma 8.1. Let $\lambda \in \Lambda $ . Then as multiplicative character of $\Lambda ^\prime $ , we have

$$\begin{align*}q^{\text{pr}_{E^\prime}(\lambda)}\vert_{\Lambda^\prime}=\prod_{\alpha\in\Phi_0^+}\big(q_\alpha^{\alpha(\lambda)/h}\big)^\alpha\vert_{\Lambda^\prime}. \end{align*}$$

Proof. Let $\lambda \in \Lambda $ . Then $\text {pr}_{E^\prime }(\lambda )\in \text {pr}_{E^\prime }(\Lambda )\in \mathcal {L}$ ; hence, $q^{\text {pr}_{E^\prime }(\lambda )}$ is well defined as multiplicative character of $\Lambda ^\prime $ . For $\mu \in \Lambda ^\prime $ , we have

$$ \begin{align*} \begin{aligned} q^{\langle\text{pr}_{E^\prime}(\lambda),\mu\rangle}&=q^{\frac{1}{h}\sum_{\alpha\in\Phi_0^+}\alpha(\lambda)\langle\alpha^{\vee},\mu\rangle}\\ &=\prod_{\alpha\in\Phi_0^+}\big(q_\alpha^{\alpha(\lambda)/h}\big)^{\alpha(\mu)}, \end{aligned} \end{align*} $$

where we used Proposition 2.10 in the first equality and the fact that $\alpha (\lambda )$ and $\alpha (\mu )$ are integers in the second equality.

For general multiplicative characters $\mathfrak {t}\in T_{\Lambda ^\prime }$ , we have the following result. Recall that

$$\begin{align*}\Lambda^\prime\cap E^\prime\subseteq\frac{1}{h}Q^{\vee} \end{align*}$$

by Proposition 2.10.

Lemma 8.2. Let $\mathfrak {t}\in T_{\Lambda ^\prime }$ and suppose that there exists a $\widetilde {\mathfrak {t}}\in T_{\frac {1}{h}Q^{\vee }}$ such that

$$\begin{align*}\widetilde{\mathfrak{t}}\vert_{\Lambda^\prime\cap E^\prime}=\mathfrak{t}\vert_{\Lambda^\prime\cap E^\prime}. \end{align*}$$

For $\alpha \in \Phi _0$ , denote by $\kappa _\alpha (\widetilde {\mathfrak {t}})\in \mathbf {F}^\times $ the value of $\widetilde {\mathfrak {t}}$ at $\alpha ^{\vee }/h$ .

  1. 1. We have

    (8.1) $$ \begin{align} \mathfrak{t}\vert_{\Lambda^\prime\cap E^\prime}=\prod_{\alpha\in\Phi_0^+}\kappa_\alpha(\widetilde{\mathfrak{t}})^\alpha\vert_{\Lambda^\prime\cap E^\prime} \end{align} $$
    as multiplicative characters of $\Lambda ^\prime \cap E^\prime \in \mathcal {L}$ .
  2. 2. We have

    $$\begin{align*}\kappa_{-\alpha}(\widetilde{\mathfrak{t}})=\kappa_\alpha(\widetilde{\mathfrak{t}})^{-1},\qquad \kappa_{v\alpha}(\widetilde{\mathfrak{t}})=\kappa_\alpha(v^{-1}\widetilde{\mathfrak{t}}),\qquad \kappa_\alpha(q^\lambda\widetilde{\mathfrak{t}})=q_\alpha^{\alpha(\lambda)/h}\kappa_\alpha(\widetilde{\mathfrak{t}}) \end{align*}$$
    for $v\in W_0$ and $\lambda \in \Lambda $ , where $q^\lambda $ is viewed here as the multiplicative character of $\frac {1}{h}Q^{\vee }$ mapping $\mu $ to $q^{\langle \lambda ,\mu \rangle }$ for $\mu \in \frac {1}{h}Q^{\vee }$ .

Proof. (1) Let $\lambda \in \Lambda ^\prime \cap E^\prime $ . By Proposition 2.10, we have

$$\begin{align*}\mathfrak{t}^\lambda=\widetilde{\mathfrak{t}}^{\,\text{pr}_{E^\prime}(\lambda)}=\widetilde{\mathfrak{t}}^{\,\sum_{\alpha\in\Phi_0^+}\frac{\alpha(\lambda)}{h}\alpha^{\vee}}. \end{align*}$$

Since $\alpha (\lambda )\in \mathbb {Z}$ , we conclude that

$$\begin{align*}\mathfrak{t}^\lambda=\prod_{\alpha\in\Phi_0^+}\big(\widetilde{\mathfrak{t}}^{\,\alpha^{\vee}/h}\big)^{\alpha(\lambda)}= \prod_{\alpha\in\Phi_0^+}\kappa_\alpha(\widetilde{\mathfrak{t}})^{\alpha(\lambda)}. \end{align*}$$

(2) This follows by a direct check.

Let $\mathcal {G}^{\text {amb}}$ be the group of tuples $\mathbf {f}:=(f_\alpha )_{\alpha \in \Phi _0}$ of functions $f_\alpha : \mathbb {R}\rightarrow \mathbf {F}^\times $ (the group operation $\mathbf {f}\cdot \mathbf {f}^\prime $ is component-wise pointwise multiplication). For fixed $y\in E$ and $\mathbf {f}\in \mathcal {G}^{\text {amb}}$ , consider the multiplicative character

(8.2) $$ \begin{align} \mathfrak{t}_y(\mathbf{f}):=\prod_{\alpha\in\Phi_0^+}f_\alpha(\alpha(y))^{\alpha}. \end{align} $$

Then $\mathbf {f}\mapsto \mathfrak {t}_y(\mathbf {f})\vert _{\Lambda ^\prime }$ defines a group homomorphism $\mathcal {G}^{\text {amb}}\rightarrow T_{\Lambda ^\prime }$ .

Remark 8.3. Suppose that $q^{\frac {1}{\text {lcm}(e,2h)}}$ is part of an injective group homomorphism $\mathbb {R}\hookrightarrow \mathbf {F}^\times $ , $d\mapsto q^{\,d}$ .

  1. 1. For $y\in E$ , write $q^y\vert _{\Lambda ^\prime }$ for the multiplicative character of $\Lambda ^\prime $ , defined by the formula $\mu \mapsto q^{\langle y,\mu \rangle }$ for $\mu \in \Lambda ^\prime $ . Then we have for $c\in C^J$ ,

    $$\begin{align*}q^c\vert_{\Lambda^\prime}\in T_{\Lambda^\prime,J},\qquad\quad q^{\text{pr}_{E^\prime}(c)}\vert_{\Lambda^\prime}\in \widetilde{T}_{\Lambda^\prime,J}. \end{align*}$$
  2. 2. The map $C^J\cap E^\prime \rightarrow T_J$ , $c\mapsto q^c\vert _{Q^{\vee }}$ is an injective map with image the set of characters $t\in T_J$ taking values in the subgroup $\{q^{d}\}_{d\in \mathbb {R}}$ of $\mathbf {F}^\times $ . Indeed, for such t, its pre-image $\text {log}_q(t)\in C^J\cap E^\prime $ is the unique vector such that $t^{\alpha _i^{\vee }}=q^{\langle \alpha _i^{\vee },\text {log}_q(t)\rangle }$ for $i\in [1,r]$ . By (the proof of) Lemma 8.2, we then also have

    $$\begin{align*}t=\mathfrak{t}_{\text{log}_q(t)}(\widehat{\mathbf{p}})\vert_{Q^{\vee}} \end{align*}$$
    in T, with $\widehat {\mathbf {p}}=(\widehat {p}_\alpha )_{\alpha \in \Phi _0}\in \mathcal {G}^{\text {amb}}$ defined by
    (8.3) $$ \begin{align} \widehat{p}_\alpha(d):=q^{2d/h\|\alpha\|^2}\qquad\quad (\alpha\in\Phi_0, d\in\mathbb{R}). \end{align} $$

    In fact, for this choice of $\widehat {\mathbf {p}}\in \mathcal {G}^{\text {amb}}$ , we have

    (8.4) $$ \begin{align} \mathfrak{t}_y(\widehat{\mathbf{p}})\vert_{\Lambda^\prime}=q^{\text{pr}_{E^\prime}(y)}\vert_{\Lambda^\prime} \qquad (y\in E), \end{align} $$
    which immediately follows from (2.29).

We will now consider a subset $\widehat {\mathcal {G}}\subset \mathcal {G}^{\text {amb}}$ such that for all $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ , the map

$$\begin{align*}E\rightarrow T_{\Lambda^\prime}, \qquad y\mapsto \mathfrak{t}_y(\widehat{\mathbf{g}})\vert_{\Lambda^\prime} \end{align*}$$

is W-equivariant. The conditions are motivated by Lemma 8.2(2) and Remark 8.3(2).

Definition 8.4. Denote by $\widehat {\mathcal {G}}$ the tuples $\widehat {\mathbf {g}}=(\widehat {g}_\alpha )_{\alpha \in \Phi _0}\in \mathcal {G}^{\text {amb}}$ satisfying

  1. 1. $\widehat {g}_{v\alpha }(d+\ell )=q_\alpha ^{\ell /h}\widehat {g}_\alpha (d)$ ,

  2. 2. $\widehat {g}_\alpha (-d)=\widehat {g}_\alpha (d)^{-1}$ ,

  3. 3. $\widehat {g}_\alpha (0)=1$

for $v\in W_0$ , $\alpha \in \Phi _0$ , $\ell \in \mathbb {Z}$ and $d\in \mathbb {R}$ .

The restriction of $\widehat {\mathbf {g}}_\alpha $ to $\frac {1}{2}\mathbb {Z}$ is determined by the sign $\epsilon \in \{\pm 1\}$ such that

$$\begin{align*}\widehat{g}_\alpha\big(\frac{1}{2}\big)=\epsilon q_{\alpha}^{\frac{1}{2h}}. \end{align*}$$

Furthermore, the invariance properties from Definition 8.4(1)&(2) may be reformulated as an equivariance property for suitable actions of the infinite dihedral group on $\mathbb {R}$ and $\mathbf {F}^\times $ . The restriction of $\widehat {g}_\alpha $ to $\mathbb {R}\setminus \frac {1}{2}\mathbb {Z}$ is then determined by $\widehat {g}_\alpha \vert _{(0,1/2)}$ , which may be chosen arbitrarily.

Example 8.5. $\widehat {\mathbf {p}}\in \mathcal {G}^{\text {amb}}$ defined in Remark 8.3(2) lies in $\widehat {\mathcal {G}}$ .

Remark 8.6. The subset $\widehat {\mathcal {G}}$ is a $\mathcal {G}$ -coset in $\mathcal {G}^{\text {amb}}$ , with $\mathcal {G}$ the subgroup of $\mathcal {G}^{\text {amb}}$ consisting of tuples $\mathbf {g}=(g_\alpha )_{\alpha \in \Phi _0}$ satisfying

  1. 1. $g_{v\alpha }(d+\ell )=g_\alpha (d)$ ,

  2. 2. $g_\alpha (-d)=g_\alpha (d)^{-1}$ ,

  3. 3. $g_\alpha (0)=1$

for $v\in W_0$ , $\alpha \in \Phi _0$ , $\ell \in \mathbb {Z}$ and $d\in \mathbb {R}$ .

Lemma 8.7. Let $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ . Then

(8.5) $$ \begin{align} \mathfrak{t}_\lambda(\widehat{\mathbf{g}})\vert_{\Lambda^\prime}=q^{\text{pr}_{E^\prime}(\lambda)}\vert_{\Lambda^\prime} \end{align} $$

for all $\lambda \in \Lambda $ .

Proof. For $\lambda \in \Lambda $ and $\alpha \in \Phi _0^+$ , we have $\alpha (\lambda )\in \mathbb {Z}$ ; hence,

$$\begin{align*}\widehat{g}_\alpha(\alpha(\lambda))=q_\alpha^{\alpha(\lambda)/h} \end{align*}$$

by Definition 8.4(1)&(3). The result now follows from Lemma 8.1.

Lemma 8.8. Let $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ , $\mathbf {g}\in \mathcal {G}$ and $y\in E$ . Then

  1. 1. $\mathfrak {t}_y(\widehat {\mathbf {g}})\vert _{\Lambda ^\prime }\in T_{\Lambda ^\prime }$ does not depend on the choice $\Phi _0^+$ of positive roots,

  2. 2. We have $\mathfrak {t}_{wy}(\widehat {\mathbf {g}})\vert _{\Lambda ^\prime }=(w\mathfrak {t}_y(\widehat {\mathbf {g}}))\vert _{\Lambda ^\prime }$ for all $w\in W_{P^{\vee }}$ .

  3. 3. We have $\mathfrak {t}_{wy}(\widehat {\mathbf {g}})\vert _{P^{\vee }}=(w\mathfrak {t}_y(\widehat {\mathbf {g}}))\vert _{P^{\vee }}$ for all $w\in W_\Lambda $ .

  4. 4. We have $\mathfrak {t}_{wy}(\mathbf {g})=(Dw)\mathfrak {t}_y(\mathbf {g})$ for all $w\in W_\Lambda $ .

Proof. (1) This follows from the fact that $\widehat {g}_{-\alpha }(-d)=\widehat {g}_{-\alpha }(d)^{-1}=\widehat {g}_\alpha (d)^{-1}$ .

(2)&(3) For $v\in W_0$ , we have

$$\begin{align*}\mathfrak{t}_{vy}(\widehat{\mathbf{g}})=\prod_{\alpha\in\Phi_0^+}\widehat{g}_{v\alpha}(\alpha(y))^{v\alpha}= \prod_{\alpha\in\Phi_0^+}\widehat{g}_{\alpha}(\alpha(y))^{v\alpha}=v\mathfrak{t}_y(\widehat{\mathbf{g}}) \end{align*}$$

in $T_{\Lambda ^\prime }$ , where we have used (1) in the first equality. For $\lambda \in \Lambda $ , we have in $T_{\Lambda ^\prime }$ ,

(8.6) $$ \begin{align} \mathfrak{t}_{y+\lambda}(\widehat{\mathbf{g}})=\prod_{\alpha\in\Phi_0^+}\widehat{g}_\alpha(\alpha(y)+\alpha(\lambda))^\alpha=\prod_{\alpha\in\Phi_0^+}q_\alpha^{\alpha(\lambda)\alpha/h}\widehat{g}_\alpha(\alpha(y))^\alpha=q^{\text{pr}_{E^\prime}(\lambda)}\mathfrak{t}_y(\widehat{\mathbf{g}}), \end{align} $$

where the last equality is due to Lemma 8.1. When $\lambda \in P^{\vee }$ , we conclude from (8.6) that $\mathfrak {t}_{y+\lambda }(\widehat {\mathbf {g}})=q^\lambda \mathfrak {t}_y(\widehat {\mathbf {g}})=\tau (\lambda )\mathfrak {t}_y(\widehat {\mathbf {g}})$ in $T_{\Lambda ^\prime }$ , which proves (2). For (3), note that for $\lambda \in \Lambda $ ,

$$\begin{align*}q^{\text{pr}_{E^\prime}(\lambda)}\vert_{P^{\vee}}=q^\lambda\vert_{P^{\vee}} \end{align*}$$

since $P^{\vee }\subset E^\prime $ . Hence, (8.6) also implies that $\mathfrak {t}_{y+\lambda }(\widehat {\mathbf {g}})\vert _{P^{\vee }}=(\tau (\lambda )\mathfrak {t}_y(\widehat {\mathbf {g}}))\vert _{P^{\vee }}$ for all $\lambda \in \Lambda $ . This proves (3).

(4) This is immediate for $w=\tau (\lambda )$ with $\lambda \in \Lambda $ since $g_\alpha $ is $1$ -periodic. For $w=v\in W_0$ , it follows as in the proof of (2).

Denote $\widetilde {T}_{\Lambda ,J}^{\text {red}}:=T_{\Lambda ,J}^{\text {red}}\cap \widetilde {T}_\Lambda $ , with $\widetilde {T}_\Lambda $ given by (3.5).

Corollary 8.9. Let $c\in C^J$ , $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ and $\mathbf {g}\in \mathcal {G}$ . We then have $\mathfrak {t}_c(\widehat {\mathbf {g}})\vert _{\Lambda ^\prime }\in \widetilde {T}_{\Lambda ^\prime ,J}$ and $\mathfrak {t}_c(\mathbf {g})\vert _{\Lambda ^\prime }\in \widetilde {T}_{\Lambda ^\prime ,J}^{\text {red}}$ .

Proof. Let $c\in C^J$ and $j\in J$ . We have $s_j\mathfrak {t}_{c}(\widehat {\mathbf {g}})=\mathfrak {t}_{s_jc}(\widehat {\mathbf {g}})=\mathfrak {t}_{c}(\widehat {\mathbf {g}})$ in $T_{P^{\vee }}$ by Lemma 8.8(2). By (2.37), we conclude that $\mathfrak {t}_{c}(\widehat {\mathbf {g}})^{\alpha _j^{\vee }}=1$ . Hence, $\mathfrak {t}_c(\widehat {\mathbf {g}})\in T_{\Lambda ^\prime ,J}$ . Clearly, $\mathfrak {t}_c(\widehat {\mathbf {g}})\vert _{\Lambda ^\prime \cap E_{\text {co}}}\equiv 1$ ; hence, $\mathfrak {t}_c(\widehat {\mathbf {g}})\in \widetilde {T}_{\Lambda ^\prime ,J}$ . The statement about $\mathfrak {t}_c(\mathbf {g})$ follows in the same way, now using Lemma 8.8(3).

We will now add a $T_{\Lambda ^\prime ,[1,r]}$ -factor to $\mathfrak {t}_y(\widehat {\mathbf {g}})$ (cf. Lemma 3.3). For the associated extended quasi-polynomial representations of $\mathbb {H}_{Q^{\vee },\Lambda ^\prime }$ , this corresponds to twisting by an algebra automorphism $\Xi _{\mathfrak {t}}$ ( $\mathfrak {t}\in T_{\Lambda ^\prime ,[1,r]}\simeq \text {Hom}(\Omega _{\Lambda ^\prime },\mathbf {F}^\times )$ ); see (the proof of) Proposition 7.25.

Definition 8.10.

  1. 1. Denote by $\mathcal {C}_{\Lambda ,\Lambda ^\prime }$ the space of functions $\mathfrak {c}: E_{\text {co}}\rightarrow T_{\Lambda ^\prime ,[1,r]}$ such that

    (8.7) $$ \begin{align} \mathfrak{c}(y+\lambda)=q^\lambda\mathfrak{c}(y)\qquad \forall\, y\in E_{\text{co}},\, \forall\, \lambda\in\text{pr}_{E_{\text{co}}}(\Lambda). \end{align} $$
  2. 2. For $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ , $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ and $y\in E$ , set

    $$\begin{align*}\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c}):=\mathfrak{t}_y(\widehat{\mathbf{g}})\mathfrak{c}\big(\text{pr}_{E_{\text{co}}}(y)\big)\in T_{\Lambda^\prime} \end{align*}$$
    (we suppress the dependence on the two lattices $\Lambda ,\Lambda ^\prime $ ).

Since $\mathfrak {c}(y)\vert _{Q^{\vee }}=1_T$ for $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ and $y\in E_{\text {co}}$ , the set $\mathcal {C}_{\Lambda ,\Lambda ^\prime }$ identifies with the set of functions $E_{\text {co}}/\text {pr}_{E_{\text {co}}}(\Lambda )\rightarrow \text {Hom}(\Lambda ^\prime /Q^{\vee },\mathbf {F}^\times )\simeq \text {Hom}(\Omega _{\Lambda ^\prime },\mathbf {F}^\times )$ . In particular, $\mathcal {C}_{\Lambda ,\Lambda ^\prime }$ is nonempty. Note that

(8.8) $$ \begin{align} \mathfrak{c}(\text{pr}_{E_{\text{co}}}(\lambda))=q^{\text{pr}_{E_{\text{co}}}(\lambda)}\mathfrak{c}(0)\qquad\quad (\lambda\in\Lambda) \end{align} $$

as multiplicative characters of $\Lambda ^\prime $ . Furthermore, if $(\Lambda _i,\Lambda _i^\prime )\in (\mathcal {L}^{\times 2})_e$ and $(\Lambda _1,\Lambda _1^\prime )\leq (\Lambda _2,\Lambda _2^\prime )$ , then the assignment $\mathfrak {c}\mapsto \mathfrak {c}(\cdot )\vert _{\Lambda _1^\prime }$ defines a map $\mathcal {C}_{\Lambda _2,\Lambda _2^\prime }\rightarrow \mathcal {C}_{\Lambda _1,\Lambda _1^\prime }$ .

For $\mu \in \Lambda ^\prime $ , we have

(8.9) $$ \begin{align} \mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c})^{\mu}=\mathfrak{t}_y(\widehat{\mathbf{g}})^{\text{pr}_{E^\prime}(\mu)}\mathfrak{c}(\text{pr}_{E_{\text{co}}}(c_y))^\mu. \end{align} $$

Special cases of this formula are

$$\begin{align*}\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c})^\mu=\mathfrak{t}_y(\widehat{\mathbf{g}})^\mu\quad (\mu\in Q^{\vee}),\qquad \mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c})^\nu=\mathfrak{c}(\text{pr}_{E_{\text{co}}}(c_y))^\nu\quad (\nu\in \Lambda^\prime\cap E_{\text{co}}). \end{align*}$$

Example 8.11. Consider the setup of Remark 8.3. Then

$$\begin{align*}\mathfrak{c}(y):=q^y\in T_{\Lambda^\prime}\qquad (y\in E_{\text{co}}) \end{align*}$$

lies $\mathcal {C}_{\Lambda ,\Lambda ^\prime }$ . Taking $\mathbf {p}=(\widehat {p}_\alpha )_{\alpha \in \Phi _0}\in \widehat {\mathcal {G}}$ with $\widehat {p}_\alpha (d)$ defined by (8.3), we then have for any $y\in E$ ,

(8.10) $$ \begin{align} \mathfrak{t}_y(\widehat{\mathbf{p}},\mathfrak{c})=\mathfrak{t}_y(\widehat{\mathbf{p}})q^{\text{pr}_{E_{\text{co}}}(y)}=q^y \end{align} $$

in $T_{\Lambda ^\prime }$ . Here, we used (8.4) for the second equality.

Lemma 8.8 and Corollary 8.9 generalize as follows.

Lemma 8.12. Let $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ , $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ and $y\in E$ .

  1. 1. For $w\in W_\Lambda $ , we have

    $$\begin{align*}\mathfrak{t}_{wy}(\widehat{\mathbf{g}},\mathfrak{c})=w\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c}) \end{align*}$$
    in $T_{\Lambda ^\prime }$ .
  2. 2. $\mathfrak {t}_c(\widehat {\mathbf {g}},\mathfrak {c})\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ for $c\in \overline {C}_+$ . In particular, for $c\in C^J$ , we have $\mathfrak {t}_{c}(\widehat {\mathbf {p}},\mathfrak {c})\in T_{\Lambda ^\prime ,J}$ .

  3. 3. For $y=\lambda \in \Lambda $ , we have

    $$\begin{align*}\mathfrak{t}_\lambda(\widehat{\mathbf{g}},\mathfrak{c})=q^\lambda\mathfrak{c}(0) \end{align*}$$
    as multiplicative characters of $\Lambda ^\prime $ .

Proof. (1) By Lemma 8.8(2) and the fact that $\mathfrak {c}$ takes values in $T_{\Lambda ^\prime }^{W_0}$ , we have $\mathfrak {t}_{vy}(\widehat {\mathbf {g}},\mathfrak {c})=v\mathfrak {t}_y(\widehat {\mathbf {g}},\mathfrak {c})$ in $T_{\Lambda ^\prime }$ for $v\in W_0$ . Let $\lambda \in \Lambda $ . By (8.6), we have

$$\begin{align*}\mathfrak{t}_{y+\lambda}(\widehat{\mathbf{g}})=q^{\text{pr}_{E^\prime}(\lambda)}\mathfrak{t}_y(\widehat{\mathbf{g}}) \end{align*}$$

in $T_{\Lambda ^\prime }$ . By (8.7), we have

$$\begin{align*}\mathfrak{c}(\text{pr}_{E_{\text{co}}}(y+\lambda))=q^{\text{pr}_{E_{\text{co}}}(\lambda)}\mathfrak{c}(\text{pr}_{E_{\text{co}}}(y)) \end{align*}$$

in $T_{\Lambda ^\prime }$ . Combining both formulas, we conclude that

$$\begin{align*}\mathfrak{t}_{y+\lambda}(\widehat{\mathbf{g}},\mathfrak{c})=q^{\lambda}\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c}) \end{align*}$$

in $T_{\Lambda ^\prime }$ , which completes the proof.

(2) Suppose now that $c\in C^J$ . Then $\mathfrak {t}_c(\widehat {\mathbf {g}},\mathfrak {c})\in T_{\Lambda ^\prime ,J}$ by Corollary 8.9 and by the fact that $\mathfrak {c}$ takes values in $T_{\Lambda ^\prime ,[1,r]}$ . Furthermore, for $\omega \in \Omega _{\Lambda ,c}$ ,

$$\begin{align*}\omega\mathfrak{t}_c(\widehat{\mathbf{g}},\mathfrak{c})=\mathfrak{t}_{\omega c}(\widehat{\mathbf{g}},\mathfrak{c})=\mathfrak{t}_c(\widehat{\mathbf{g}},\mathfrak{c}) \end{align*}$$

by part (1) of the lemma. Hence, $\mathfrak {t}_c(\widehat {\mathbf {g}},\mathfrak {c})\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ .

(3) For $\lambda \in \Lambda $ , we have in $T_{\Lambda ^\prime }$ ,

$$\begin{align*}\mathfrak{t}_\lambda(\widehat{\mathbf{g}},\mathfrak{c})=\mathfrak{t}_\lambda(\widehat{\mathbf{g}})\mathfrak{c}(\text{pr}_{E_{\text{co}}}(\lambda))= q^{\text{pr}_{E^\prime}(\lambda)}q^{\text{pr}_{E_{\text{co}}}(\lambda)}\mathfrak{c}(0)=q^\lambda\mathfrak{c}(0), \end{align*}$$

where we used Lemma 8.7 and (8.9) for the second equality.

Remark 8.13. Let $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ , $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ and $c\in \overline {C}_+$ . Applying Definition 7.16 to the multiplicative character $\mathfrak {t}_{c}(\widehat {\mathbf {g}},\mathfrak {c})\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ gives rise to the multiplicative characters

$$\begin{align*}\mathfrak{t}_{c}(\widehat{\mathbf{g}},\mathfrak{c})_{wc;c}=w\mathfrak{t}_{c}(\widehat{\mathbf{g}},\mathfrak{c})\in T_{\Lambda^\prime}\qquad\quad (w\in W_{\Lambda}). \end{align*}$$

It then follows from Lemma 8.12 that

(8.11) $$ \begin{align} \mathfrak{t}_{c}(\widehat{\mathbf{g}},\mathfrak{c})_{wc;c}=\mathfrak{t}_{wc}(\widehat{\mathbf{g}},\mathfrak{c})\in T_{\Lambda^\prime}\qquad\quad (w\in W_\Lambda) \end{align} $$

in $T_{\Lambda ^\prime }$ .

The following extension of Lemma 7.17 is a direct consequence of Lemma 8.12(2).

Lemma 8.14. Let $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ and $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ . The formulas

$$\begin{align*}w\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y:=w_{\mathfrak{t}_{c_y}(\widehat{\mathbf{g}},\mathfrak{c})}x^y\qquad \quad (w\in W_{\Lambda^\prime},\, y\in E) \end{align*}$$

define a linear left $W_{\Lambda ^\prime }$ -action on $\mathbf {F}[E]$ .

Concretely, by Remark 8.13, we have the formulas

$$ \begin{align*} \begin{aligned} v\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y&:=x^{vy}\qquad\qquad\quad\,\,\, (v\in W_0),\\ \tau(\mu)\cdot_{\widehat{\mathbf{g}}, \mathfrak{c}}x^y&:=\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c})^{-\mu}x^y\qquad (\mu\in\Lambda^\prime), \end{aligned} \end{align*} $$

for $y\in E$ . Furthermore,

(8.12) $$ \begin{align} s_0\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y=\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c})^{\varphi^{\vee}}x^{s_\varphi y},\qquad (w_{\zeta,\Lambda^\prime})\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y=\mathfrak{t}_{v_{\zeta,\Lambda^\prime}^{-1}y}(\widehat{\mathbf{g}},\mathfrak{c})^{-\zeta}x^{v_{\zeta,\Lambda^\prime}^{-1}y} \end{align} $$

for $\zeta \in \Lambda _{\text {min}}^\prime $ , which follows from Remark 8.13 and (7.14). Since $\mathfrak {c}$ takes values in $T_{\Lambda ^\prime ,[1,r]}$ , we have that

$$\begin{align*}w\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y=\Theta_{\mathfrak{c}(\text{pr}_{E_{\text{co}}}(y))}(w)\big(w_{\mathfrak{t}_{c_y}(\widehat{\mathbf{g}})}(x^y)\big) \end{align*}$$

for $w\in W_{\Lambda ^\prime }$ and $y\in E$ by (7.32). In particular, the $\mathfrak {c}$ -dependence of the $W_{\Lambda ^\prime }$ -action $\cdot _{\widehat {\mathbf {g}},\mathfrak {c}}$ on $\mathbf {F}[E]$ (and later on, in Theorem 8.19, the $\mathfrak {c}$ -dependence of the corresponding extended double affine Hecke algebra representation), corresponds to twisting the representations by multiplicative characters of $\Omega _{\Lambda ^\prime }$ ; cf. Proposition 7.25.

Example 8.15. Consider the setup of Example 8.11. Then we have

(8.13) $$ \begin{align} \tau(\mu)\cdot_{\widehat{\mathbf{p}},\mathfrak{c}} x^y=q^{-\langle\mu,y\rangle}x^y\qquad (\mu\in\Lambda^\prime,\, y\in E), \end{align} $$

so in this case, $\cdot _{\widehat {\mathbf {p}},\mathfrak {c}}$ is the standard $W_{\Lambda ^\prime }$ -action on $\mathbf {F}[E]$ by q-translations and reflections.

8.2 The uniform quasi-polynomial representation

For $\widehat {\mathbf {g}}\in \widehat {\mathcal {G}}$ , $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ and $c\in \overline {C}_+$ , we have the extended quasi-polynomial representation

$$\begin{align*}\pi_{c,\mathfrak{t}_c(\widehat{\mathbf{g}},\mathfrak{c})}^{\Lambda,\Lambda^\prime}: \mathbb{H}_{\Lambda,\Lambda^\prime}\rightarrow\text{End}(\mathcal{P}^{(c)}_\Lambda) \end{align*}$$

from Theorem 7.18, since $\mathfrak {t}_c(\widehat {\mathbf {g}},\mathfrak {c})\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ by Lemma 8.12(2). Their direct sum turns $\mathbf {F}[E]$ into an $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -module, with an explicit $(\widehat {\mathbf {g}},\mathfrak {c})$ -dependent $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -action in terms of truncated Demazure-Lusztig operators.

To connect it to metaplectic representation theory, in particular to the $\text {GL}_{r+1}$ -type double affine Hecke algebra representation from [Reference Sahi, Stokman and Venkateswaran38, §5], it is convenient to fix a representative $\widehat {\mathbf {p}}$ of $\widehat {\mathcal {G}}$ and twist the $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -actions $\pi _{c,\mathfrak {t}_c(\widehat {\mathbf {p}}\cdot \mathbf {g},\mathfrak {c})}^{\Lambda ,\Lambda ^\prime }$ on $\mathcal {P}_\Lambda ^{(c)}$ by a linear automorphism of $\mathcal {P}_\Lambda ^{(c)}$ that separates the parameters $\mathbf {g}\in \mathcal {G}$ from $\widehat {\mathbf {p}}\cdot \mathbf {g}\in \widehat {\mathcal {G}}$ (the explicit connection to the metaplectic representation theory will be the subject of Section 10). The linear automorphism, depending on $\mathbf {g}\in \mathcal {G}$ , is defined as follows.

Lemma 8.16. Let $c\in \overline {C}_+$ and $\mathbf {g}\in \mathcal {G}$ . There exists a unique $\mathcal {P}_\Lambda $ -linear automorphism $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ of $\mathcal {P}_\Lambda ^{(c)}$ satisfying

(8.14) $$ \begin{align} \Gamma_{\Lambda,\mathbf{g}}^{(c)}(x^{wc})=\left(\prod_{\alpha\in\Pi(Dw)}g_\alpha(\alpha(c))\right)x^{wc}\qquad\quad (w\in W_\Lambda). \end{align} $$

Proof. Let $\mathbf {F}_{\text {cl}}$ be the algebraic closure of $\mathbf {F}$ . Write $\mathcal {G}_{\text {cl}}$ for the space of $\mathbf {g}$ -parameters $\mathcal {G}$ over the field $\mathbf {F}_{\text {cl}}$ (see Remark 8.6). For $\alpha \in \Phi _0$ , choose $\widetilde {\mathbf {g}}=(\widetilde {g}_\alpha (d))_{\alpha \in \Phi _0}\in \mathcal {G}_{\text {cl}}$ satisfying $\widetilde {g}_\alpha (d)^2=g_\alpha (d)$ for $\alpha \in \Phi _0$ and $d\in \mathbb {R}$ . Define $\gamma _{\widetilde {\mathbf {g}}}: E\rightarrow \mathbf {F}_{\text {cl}}^\times $ by

(8.15) $$ \begin{align} \gamma_{\widetilde{\mathbf{g}}}(y):=\prod_{\alpha\in\Phi_0^+}\widetilde{g}_\alpha(\alpha(y)). \end{align} $$

Then

(8.16) $$ \begin{align} \begin{aligned} \gamma_{\widetilde{\mathbf{g}}}(y+\mu)&=\gamma_{\widetilde{\mathbf{g}}}(y)\qquad\qquad\qquad\qquad\quad\quad\,\,\, (\mu\in\Lambda),\\ \gamma_{\widetilde{\mathbf{g}}}(vy)&=\left(\prod_{\alpha\in\Pi(v)}g_\alpha(\alpha(y))^{-1}\right)\gamma_{\widetilde{\mathbf{g}}}(y)\qquad (v\in W_0). \end{aligned} \end{align} $$

Let $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ be the $\mathbf {F}_{\text {cl}}$ -linear automorphism of $\bigoplus _{y\in \mathcal {O}_{\Lambda ,c}}\mathbf {F}_{\text {cl}}x^y$ defined by

(8.17) $$ \begin{align} \Gamma_{\Lambda,\mathbf{g}}^{(c)}(x^y):=\frac{\gamma_{\widetilde{\mathbf{g}}}(c)}{\gamma_{\widetilde{\mathbf{g}}}(y)}x^y\qquad\quad (y\in\mathcal{O}_{\Lambda,c}). \end{align} $$

Then

$$\begin{align*}\Gamma_{\Lambda,\mathbf{g}}^{(c)}(x^{wc})=\left(\prod_{\alpha\in\Pi(Dw)}g_\alpha(\alpha(c))\right)x^{wc}\qquad\quad (w\in W_\Lambda) \end{align*}$$

by (8.16). Hence, $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ does not depend on the choice of $\widetilde {\mathbf {g}}$ , and it restricts to a $\mathbf {F}$ -linear automorphism of $\mathcal {P}_\Lambda ^{(c)}$ satisfying (8.14). Note that $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ is $\mathcal {P}_\Lambda $ -linear due to $\Lambda $ -translation invariance of $\gamma _{\widetilde {\mathbf {g}}}$ .

Remark 8.17. If $c^\prime \in \mathcal {O}_{\Lambda ,c}\cap \overline {C}_+$ , then $c^\prime =\omega c$ for some $\omega \in \Omega _{\Lambda }$ and

$$\begin{align*}\Gamma_{\Lambda,\mathbf{g}}^{(c)}=\left(\prod_{\alpha\in \Pi(D\omega)}g_\alpha(\alpha(c))\right)\Gamma_{\Lambda,\mathbf{g}}^{(\omega c)} \end{align*}$$

in view of (8.17) and (8.16).

The $W_{\Lambda ^\prime }$ -action $\cdot _{\widehat {\mathbf {g}},\mathfrak {c}}$ on $\mathbf {F}[E]$ from Lemma 8.14 preserves $\mathcal {P}_{\Lambda }^{(c)}$ . The following lemma shows that $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ intertwines the $W_{\Lambda ^\prime }$ -action $\cdot _{\widehat {\mathbf {p}},\mathfrak {c}}$ on $\mathcal {P}_\Lambda ^{(c)}$ with a twisted version of the $W_{\Lambda ^\prime }$ -action $\cdot _{\widehat {\mathbf {g}},\mathfrak {c}}$ on $\mathcal {P}_\Lambda ^{(c)}$ , where $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}$ .

Lemma 8.18. Let $c\in \overline {C}_+$ , $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ , $\mathbf {g}\in \mathcal {G}$ and $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ . Set $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}\in \widehat {\mathcal {G}}$ . Then

$$\begin{align*}\Gamma_{\Lambda,\mathbf{g}}^{(c)}\big(w\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}\Gamma_{\Lambda,\mathbf{g}}^{(c)\,-1}(x^y)\big)= \left(\prod_{a\in\Pi(w)}g_{Da}(a(y))\right)w\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y \end{align*}$$

for $w\in W_{\Lambda ^\prime }$ and $y\in \mathcal {O}_{\Lambda ,c}$ .

Proof. We use the notations introduced in the proof of Lemma 8.16. Fix $c\in \overline {C}_+$ . By a standard argument using (2.12) and Remark 8.6(1), it suffices to prove that

(8.18) $$ \begin{align} \Gamma_{\Lambda,\mathbf{g}}^{(c)}\big(s_j\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}\Gamma_{\Lambda,\mathbf{g}}^{(c)\,-1}(x^y)\big)= g_{D\alpha_j}(\alpha_j(y))\,s_{j}\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y \end{align} $$

for $j\in [0,r]$ and $y\in \mathcal {O}_{\Lambda ,c}$ and

(8.19) $$ \begin{align} \Gamma_{\Lambda,\mathbf{g}}^{(c)}\big(\omega\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}\Gamma_{\Lambda,\mathbf{g}}^{(c)\,-1}(x^y)\big)=\omega\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y \end{align} $$

for $\omega \in \Omega _{\Lambda ^\prime }$ and $y\in \mathcal {O}_{\Lambda ,c}$ .

For $i\in [1,r]$ and $y\in \mathcal {O}_{\Lambda ,c}$ , we have

$$\begin{align*}\Gamma_{\Lambda,\mathbf{g}}^{(c)}\big(s_i\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}\Gamma_{\Lambda,\mathbf{g}}^{(c)\,-1}(x^y)\big)=\frac{\gamma_{\widetilde{\mathbf{g}}}(y)}{\gamma_{\widetilde{\mathbf{g}}}(s_iy)}x^{s_iy}= g_{\alpha_i}(\alpha_i(y))\,s_{i}\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y, \end{align*}$$

which is (8.18) for $j=i\in [1,r]$ . By (8.12), we have for $y\in \mathcal {O}_{\Lambda ,c}$ ,

$$ \begin{align*} \Gamma_{\Lambda,\mathbf{g}}^{(c)}\big(s_0\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}\Gamma_{\Lambda,\mathbf{g}}^{(c)\,-1}(x^y)\big)=\frac{\gamma_{\widetilde{\mathbf{g}}}(y)}{\gamma_{\widetilde{\mathbf{g}}}(s_\varphi y)} \mathfrak{t}_y(\widehat{\mathbf{p}},\mathfrak{c})^{\varphi^{\vee}}x^{s_\varphi y}. \end{align*} $$

Since $\mathfrak {t}_y(\widehat {\mathbf {g}},\mathfrak {c})=\mathfrak {t}_y(\widehat {\mathbf {p}},\mathfrak {c})\mathfrak {t}_y(\mathbf {g})$ , formula (8.18) for $j=0$ then follows from the fact that

$$\begin{align*}\mathfrak{t}_y(\mathbf{g})^{\varphi^{\vee}}=g_{\varphi}(\varphi(y))\prod_{\alpha\in\Pi(s_\varphi)}g_\alpha(\alpha(y))=\frac{\gamma_{\widetilde{\mathbf{g}}}(y)}{g_{D\alpha_0}(\alpha_0(y))\gamma_{\widetilde{\mathbf{g}}}(s_\varphi y)} \end{align*}$$

by (2.4), (2.18) and (8.16).

It remains to prove (8.19). By (8.12), we have for $y\in E$ and $\zeta \in \Lambda ^\prime _{\text {min}}$ ,

$$\begin{align*}w_{\zeta,\Lambda^\prime}\cdot_{\widehat{\mathbf{g}},\mathfrak{c}}x^y=\mathfrak{t}_{v_{\zeta,\Lambda^\prime}^{-1}y}(\mathbf{g})^{-\zeta}(w_{\zeta,\Lambda^\prime}\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}x^y). \end{align*}$$

Formula (8.19) will thus follow from the equality

(8.20) $$ \begin{align} \frac{\gamma_{\widetilde{\mathbf{g}}}(y)}{\gamma_{\widetilde{\mathbf{g}}}(v_{\zeta,\Lambda^\prime}^{-1}y)}=\mathfrak{t}_{v_{\zeta,\Lambda^\prime}^{\,-1}y}(\mathbf{g})^{-\zeta} \qquad (\zeta\in\Lambda^\prime_{\text{min}}). \end{align} $$

This in turn follows from a straightforward computation using (7.4), (8.16) and Remark 8.6.

We now have the following uniform variant of the quasi-polynomial representation $\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }$ of $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ from Theorem 7.18.

Theorem 8.19. Let $\mathbf {g}\in \mathcal {G}$ , $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ and $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ .

  1. 1. The formulas

    (8.21) $$ \begin{align} \begin{aligned} \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(T_j)x^y&:=k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}g_{D\alpha_j}(\alpha_j(y))^{-1}s_j\cdot_{\widehat{\mathbf{p}},\mathfrak{c}} x^y+(k_j-k_j^{-1})\nabla_j(x^y),\\ \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(x^\lambda)x^y&:=x^{y+\lambda},\\ \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(\omega)x^y&:=\omega\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}x^y \end{aligned} \end{align} $$
    for $j\in [0,r]$ , $\lambda \in \Lambda $ , $\omega \in \Omega _{\Lambda ^\prime }$ and $y\in E$ define a representation
    $$\begin{align*}\pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}: \mathbb{H}_{\Lambda,\Lambda^\prime}\rightarrow\text{End}(\mathbf{F}[E]). \end{align*}$$
  2. 2. Let $c\in \overline {C}_+$ . Then $\mathcal {P}_\Lambda ^{(c)}$ is a $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -submodule of $(\mathbf {F}[E],\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ,\Lambda ^\prime })$ , and the $\mathcal {P}_\Lambda $ -linear automorphism $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ of $\mathcal {P}_\Lambda ^{(c)}$ realises an isomorphism

    $$\begin{align*}\Gamma_{\Lambda,\mathbf{g}}^{(c)}: \big(\mathcal{P}_\Lambda^{(c)},\pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(\cdot)\vert_{\mathcal{P}_\Lambda^{(c)}}\big)\overset{\sim}{\longrightarrow} \mathcal{P}_{\Lambda,\mathfrak{t}_c(\widehat{\mathbf{g}},\mathfrak{c})}^{(c)} \end{align*}$$
    of $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -modules, where $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}\in \widehat {\mathcal {G}}$ .

Proof. Fix $c\in \overline {C}_+$ . The linear operators on $\mathbf {F}[E]$ defined by (8.21) preserve $\mathcal {P}_\Lambda ^{(c)}$ , and $\mathfrak {t}_{c}(\widehat {\mathbf {g}},\mathfrak {c})\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ by Corollary 8.9. In view of Theorem 7.18, it thus suffices to show that

(8.22) $$ \begin{align} \Gamma_{\Lambda,\mathbf{g}}^{(c)}\big(\pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(h)\Gamma_{\Lambda,\mathbf{g}}^{(c)\,-1}(x^y)\big)= \pi_{c,\mathfrak{t}_c(\widehat{\mathbf{g}},\mathfrak{c})}^{\Lambda,\Lambda^\prime}(h)x^y\qquad\quad (y\in\mathcal{O}_{\Lambda,c}) \end{align} $$

for $h=T_j$ ( $0\leq j\leq r$ ), $h=x^\lambda $ ( $\lambda \in \Lambda $ ) and $h=\omega $ ( $\omega \in \Omega _{\Lambda ^\prime }$ ).

For $h=x^\lambda $ ( $\lambda \in \Lambda $ ), this is trivial. For $h=T_j$ , note that $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ commutes with the truncated divided difference operator $\nabla _j$ since $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ is $\mathcal {P}$ -linear. Hence, in this case, (8.22) follows from Lemma 8.18.

Finally, for $h=\omega $ ( $\omega \in \Omega _{\Lambda ^\prime }$ ), formula (8.22) follows from Lemma 8.14 and (8.19).

Remark 8.20. Recall that both the definition of the extended quasi-polynomial representation $\pi _{c,\mathfrak {t}_c(\widehat {\mathbf {g}},\mathfrak {c})}^{\Lambda ,\Lambda ^\prime }$ of $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ and of $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}\in \text {End}_{\mathcal {P}_\Lambda }(\mathcal {P}_\Lambda ^{(c)})$ depend on the choice of a representative c of the $W_\Lambda $ -orbit $\mathcal {O}_{\Lambda ,c}$ inside $\overline {C}_+$ (see Remark 7.19 and Remark 8.17). This is no longer the case for $\big (\mathcal {P}_\Lambda ^{(c)},\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ,\Lambda ^\prime }(\cdot )\vert _{\mathcal {P}_\Lambda ^{(c)}}\big )$ , but now it depends on a choice of a base-point $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ . Note that for all $c\in \overline {C}_+$ ,

$$\begin{align*}\pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(\cdot)\vert_{\mathcal{P}_\Lambda^{(c)}}\simeq\pi_{\mathbf{g}^\prime,\widehat{\mathbf{p}}^\prime,\mathfrak{c}}^{\Lambda,\Lambda^\prime}(\cdot)\vert_{\mathcal{P}_\Lambda^{(c)}} \end{align*}$$

when $\widehat {\mathbf {p}}\cdot \mathbf {g}=\widehat {\mathbf {p}}^\prime \cdot \mathbf {g}^\prime $ , in view of Theorem 8.19(3).

The uniform quasi-polynomial representations for different root data are related as follows.

Lemma 8.21. Let $(\Lambda _i,\Lambda _i^\prime )\in (\mathcal {L}^{\times 2})_e$ such that $(\Lambda _1,\Lambda _1^\prime )\leq (\Lambda _2,\Lambda _2^\prime )$ . Then

$$\begin{align*}\pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda_2,\Lambda_2^\prime}\vert_{\mathbb{H}_{\Lambda_1,\Lambda_1^\prime}}= \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}(\cdot)\vert_{\Lambda_1^\prime}}^{\Lambda_1,\Lambda_1^\prime}. \end{align*}$$

Proof. This is immediate (cf. Proposition 7.20).

8.3 Uniform quasi-polynomial eigenfunctions

In the following theorem, we introduce the uniform versions of the monic quasi-polynomial eigenfunctions $E_{y;c}^{\Lambda ,\Lambda ^\prime }(x;\mathfrak {t})$ from Theorem 7.26.

Proposition 8.22. Let $\mathcal {O}\subset E$ be a W-orbit, $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ , $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ and $\mathbf {g}\in \mathcal {G}$ such that

(8.23) $$ \begin{align} \mathfrak{s}_y\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c})\not=\mathfrak{s}_{y^\prime}\mathfrak{t}_{y^\prime}(\widehat{\mathbf{g}},\mathfrak{c})\,\, \text{ in }\, T_{\Lambda^\prime}\, \text{ when }\, y,y^\prime\in \mathcal{O}\, \text{ and } y\not=y^\prime, \end{align} $$

where $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}\in \widehat {\mathcal {G}}$ . For each $y\in \mathcal {O}$ , the following holds true.

  1. 1. There exists a unique joint eigenfunction $\mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})\in \mathbf {F}[E]$ of the commuting operators $\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ,\Lambda ^\prime }(Y^\mu )$ ( $\mu \in \Lambda ^\prime $ ) satisfying

    (8.24) $$ \begin{align} \mathcal{E}_y^{\Lambda,\Lambda^\prime}(x;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})= x^y+\text{l.o.t.} \end{align} $$
  2. 2. We have

    (8.25) $$ \begin{align} \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(Y^\mu)\mathcal{E}_y^{\Lambda,\Lambda^\prime}(\cdot;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})=(\mathfrak{s}_y\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c}))^{-\mu} \mathcal{E}_y^{\Lambda,\Lambda^\prime}(\cdot;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})\qquad\forall\, \mu\in\Lambda^{\prime}. \end{align} $$
  3. 3. We have

    (8.26) $$ \begin{align} \Gamma_{\Lambda,\mathbf{g}}^{(c_y)}\left(\mathcal{E}_y^{\Lambda,\Lambda^\prime}(x;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})\right)= \left(\prod_{\alpha\in\Pi(v_y^{-1})}g_\alpha(\alpha(c_y))\right) E_{y;c_y}^{\Lambda,\Lambda^\prime}(x;\mathfrak{t}_{c_y}(\widehat{\mathbf{g}},\mathfrak{c})). \end{align} $$

Proof. Write $c=c_y\in \overline {C}_+$ . Note that (8.23) implies the genericity condition (7.33) for $\mathfrak {t}=\mathfrak {t}_c(\widehat {\mathbf {g}},\mathfrak {c})\in {}^{\Lambda }T_{\Lambda ^\prime }^c$ , in view of (8.11). Hence, the right-hand side of (8.26) is well defined. The result now follows immediately from Theorem 8.19(2), Theorem 7.26(1)&(2) and Lemma 8.16.

Note that

$$\begin{align*}\mathcal{E}_y^{\Lambda,\Lambda^\prime}(x;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})=x^y\,\,\text{ when }\,\, y\in\overline{C}_+ \end{align*}$$

by (8.24). In particular, this holds true when $y\in E_{\text {co}}=C^{[1,r]}$ .

The explicit formula for the Y-weight of $\mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})\in \mathbf {F}[E]$ is

(8.27) $$ \begin{align} \mathfrak{s}_y\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c})=\mathfrak{c}(\text{pr}_{E_{\text{co}}}(y))\prod_{\alpha\in\Phi_0^+}\big(k_\alpha^{\eta(\alpha(y))}\widehat{g}_\alpha(\alpha(y))\big)^\alpha, \end{align} $$

where $\widehat {\mathbf {g}}=\widehat {\mathbf {p}}\cdot \mathbf {g}$ . Note that (8.25) in particular implies that

(8.28) $$ \begin{align} \pi^{\Lambda,\Lambda^\prime}_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}(Y^\zeta)\mathcal{E}_y^{\Lambda,\Lambda^\prime}(\cdot;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})=\mathfrak{c}(\text{pr}_{E_{\text{co}}}(c_y))^{-\zeta}\mathcal{E}_y^{\Lambda,\Lambda^\prime}(\cdot;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}) \qquad\quad(\zeta\in\Lambda^\prime\cap E_{\text{co}}) \end{align} $$

for all $y\in \mathcal {O}$ .

Proposition 8.23. Let $(\Lambda _i,\Lambda _i^\prime )\in (\mathcal {L}^{\times 2})_e$ such that $(\Lambda _1,\Lambda _1^\prime )\leq (\Lambda _2,\Lambda _2^\prime )$ . Let $\mathcal {O}\subset E$ be a W-orbit, $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ , $\mathfrak {c}\in \mathcal {C}_{\Lambda _2,\Lambda ^\prime _2}$ and $\mathbf {g}\in \mathcal {G}$ such that

(8.29) $$ \begin{align} \mathfrak{s}_y\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c}(\cdot)\vert_{\Lambda_1^\prime})\not=\mathfrak{s}_{y^\prime}\mathfrak{t}_{y^\prime}(\widehat{\mathbf{g}}, \mathfrak{c}(\cdot)\vert_{\Lambda_1^\prime})\,\, \text{ in }\, T_{\Lambda_1^\prime}\, \text{ when }\, y,y^\prime\in \mathcal{O}\, \text{ and } y\not=y^\prime, \end{align} $$

where $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}\in \widehat {\mathcal {G}}$ . Then

(8.30) $$ \begin{align} \mathcal{E}_y^{\Lambda_2,\Lambda_2^\prime}(x;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})=\mathcal{E}_y^{\Lambda_1,\Lambda_1^\prime}(x;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}(\cdot)\vert_{\Lambda_1^\prime}) \qquad \forall\, y\in\mathcal{O}. \end{align} $$

Proof. Note that (8.29) implies that (8.23) holds true for $\Lambda ^\prime =\Lambda _2^\prime $ . Hence, both sides of (8.30) are well defined. Furthermore, both sides of (8.30) are joint eigenfunctions of $\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ,\Lambda ^\prime }(Y^\mu )$ ( $\mu \in \Lambda _1^\prime $ ) by Lemma 8.21. The result now follows from Proposition 8.22(1) applied to the pair of lattices $(\Lambda ,\Lambda ^\prime )=(\Lambda _1,\Lambda _1^\prime )$ .

The relation between $ \mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})$ and the quasi-polynomial eigenfunctions relative to the adjoint root datum is as follows.

Corollary 8.24. Let $\mathcal {O}\subset E$ be a W-orbit, $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ , $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ and $\mathbf {g}\in \mathcal {G}$ . Set $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}\in \widehat {\mathcal {G}}$ and suppose that the genericity conditions (8.29) hold true for $\Lambda _1^\prime =Q^{\vee }$ . For $y\in \mathcal {O}$ , we then have

(8.31) $$ \begin{align} \Gamma_{Q^{\vee},\mathbf{g}}^{(c_y)}\big(\mathcal{E}_y^{\Lambda,\Lambda^\prime}(x;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})\big)= \left(\prod_{\alpha\in\Pi(v_y^{-1})}g_\alpha(\alpha(c_y))\right)E_y^{\mathbf{J}(c_y)}(x;\mathfrak{t}_{c_y}(\widehat{\mathbf{g}})\vert_{Q^{\vee}}). \end{align} $$

Under these assumptions, $\mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})$ does not depend on $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ .

Proof. We have $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}\vert _{\mathcal {P}^{(c)}}=\Gamma _{Q^{\vee },\mathbf {g}}^{(c)}$ ( $c\in \overline {C}_+$ ) and $\mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})\in \mathcal {P}^{(c_y)}$ ; hence, the left-hand side of (8.31) is well defined.

Since $(Q^{\vee },Q^{\vee })\leq (\Lambda ,\Lambda ^\prime )$ , formula (8.31) then follows from Proposition 8.23, (8.26) and the fact that

$$\begin{align*}\mathcal{E}_{y;c_y}^{Q^{\vee},Q^{\vee}}(x;\mathfrak{t}_{c_y}(\widehat{\mathbf{g}},\mathfrak{c}(\cdot)\vert_{Q^{\vee}}))=E_y^{\mathbf{J}(c_y)}(x;\mathfrak{t}_{c_y}(\widehat{\mathbf{g}})). \end{align*}$$

Here, we have also used that $\mathfrak {c}(\cdot )\vert _{Q^{\vee }}\equiv 1_T$ .

Changing the perspective by extending the eigenvalue equations from adjoint root datum to extended root data leads to the following main result of this subsection.

Theorem 8.25. Let $\mathcal {O}\subset E$ be a W-orbit, $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ and $\mathbf {g}\in \mathcal {G}$ . Suppose that

(8.32) $$ \begin{align} \mathfrak{s}_y\mathfrak{t}_y(\widehat{\mathbf{g}})\not=\mathfrak{s}_{y^\prime}\mathfrak{t}_{y^\prime}(\widehat{\mathbf{g}})\,\, \text{ in }\, T\, \text{ when }\, y,y^\prime\in \mathcal{O}\, \text{ and } y\not=y^\prime, \end{align} $$

where $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}\in \widehat {\mathcal {G}}$ . Then

$$\begin{align*}\mathcal{E}_y(x;\mathbf{g},\widehat{\mathbf{p}}):=\mathcal{E}_y^{Q^{\vee},Q^{\vee}}(x;\mathbf{g},\widehat{\mathbf{p}},1_T)\qquad (y\in\mathcal{O}) \end{align*}$$

satisfies for all $(\Lambda ,\Lambda ^\prime )\in (\mathcal {L}^{\times 2})_e$ and all $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ ,

(8.33) $$ \begin{align} \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(Y^\mu)\mathcal{E}_y(\cdot;\mathbf{g},\widehat{\mathbf{p}})=(\mathfrak{s}_y\mathfrak{t}_y(\widehat{\mathbf{g}},\mathfrak{c}))^{-\mu} \mathcal{E}_y(\cdot;\mathbf{g},\widehat{\mathbf{p}})\qquad\forall\, \mu\in\Lambda^{\prime}. \end{align} $$

Furthermore,

(8.34) $$ \begin{align} \Gamma_{Q^{\vee},\mathbf{g}}^{(c_y)}\big(\mathcal{E}_y(x;\mathbf{g},\widehat{\mathbf{p}})\big)= \left(\prod_{\alpha\in\Pi(v_y^{-1})}g_\alpha(\alpha(c_y))\right)E_y^{\mathbf{J}(c_y)}(x;\mathfrak{t}_{c_y}(\widehat{\mathbf{g}})\vert_{Q^{\vee}}). \end{align} $$

Proof. Fix $(\Lambda ,\Lambda ^\prime )\in (\mathcal {L}^{\times 2})_e$ and $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ . By (8.9), the parameter conditions (8.32) imply (8.23). Furthermore, $\mathfrak {c}(\cdot )\vert _{Q^{\vee }}=1_T$ ; hence, (8.30) shows that

$$\begin{align*}\mathcal{E}_y^{\Lambda,\Lambda^\prime}(x;\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c})=\mathcal{E}_y(x;\mathbf{g},\widehat{\mathbf{p}}) \end{align*}$$

for all $y\in \mathcal {O}$ . Then (8.33) follows from (8.25), and (8.34) from (8.31).

Remark 8.26. Under special conditions on the parameters $\mathbf {g}$ and $\widehat {\mathbf {p}}$ , the quasi-polynomials $\mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})$ ( $y\in \mathcal {O}$ ) may depend on $\mathfrak {c}\in \mathcal {C}_{\Lambda ,\Lambda ^\prime }$ . For instance, this is the case when $(\Lambda ,\Lambda ^\prime )=(Q^{\vee },P^{\vee })$ with $Q^{\vee }\subsetneq P^{\vee }$ if the Y-spectrum satisfies (8.23) but not (8.32).

Further results for $\mathcal {E}_y(x;\mathbf {g},\widehat {\mathbf {p}})$ and $\mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})$ , such as orthogonality and symmetrisation, can be directly obtained from the corresponding results for $E_{y;c}^{\Lambda ,\Lambda ^\prime }(x;\mathfrak {t})$ and $E_y^J(x;\mathfrak {t})$ using (8.26) and (8.31).

By Lemma 8.12(1), the $W_{\Lambda ^\prime }$ -action $\cdot _{\widehat {\mathbf {p}},\mathfrak {c}}$ on $\mathbf {F}[E]$ extends to a $W_{\Lambda ^\prime }\ltimes \mathcal {P}_\Lambda $ -action, with $\mathcal {P}_\Lambda $ acting by multiplication operators (cf. Lemma 4.2). This extends to a $W_{\Lambda ^\prime }\ltimes \mathcal {Q}_\Lambda $ -action on $\mathbf {F}_{\mathcal {Q}_\Lambda }[E]:=\mathcal {Q}_\Lambda \otimes _{\mathcal {P}_\Lambda }\mathbf {F}[E]$ ; cf. Subsection 4.5.

Theorem 8.27. For $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}$ , $\mathbf {g}\in \mathcal {G}$ and $\mathfrak {c}\in \mathcal {C}$ , the formulas

$$ \begin{align*} \begin{aligned} \sigma_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(s_j)(x^y)&:= \frac{k_j^{\chi_{\mathbb{Z}}(\alpha_j(y))}(x^{\alpha_j^{\vee}}-1)}{g_{D\alpha_j}(\alpha_j(y))(k_jx^{\alpha_j^{\vee}}-k_j^{-1})} s_j\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}x^y+\left(\frac{k_j-k_j^{-1}}{k_jx^{\alpha_j^{\vee}}-k_j^{-1}}\right)x^{y-\lfloor D\alpha_j(y)\rfloor\alpha_j^{\vee}},\\ \sigma_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(\omega)(x^y)&:=\omega\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}x^y, \qquad\qquad\sigma_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}(f)x^y:=fx^y \end{aligned} \end{align*} $$

for $y\in E$ , $0\leq j\leq r$ , $\omega \in \Omega _{\Lambda ^\prime }$ and $f\in \mathcal {Q}_\Lambda $ define a representation

$$\begin{align*}\sigma_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda,\Lambda^\prime}:W_{\Lambda^\prime}\ltimes\mathcal{Q}_\Lambda\rightarrow\text{End}_{\mathbf{F}}(\mathbf{F}_{\mathcal{Q}_\Lambda}[E]). \end{align*}$$

Proof. One can derive the result from Lemma 7.11 by similar arguments as in the proof of Theorem 8.19, using the extension of $\Gamma _{\Lambda ,\mathbf {g}}^{(c)}$ to a $\mathcal {Q}_\Lambda $ -linear automorphism of $\mathcal {Q}_\Lambda \otimes _{\mathcal {P}_\Lambda }\mathcal {P}_\Lambda ^{(c)}$ . Alternatively, one can repeat the proof of Theorem 4.9, now using the $\mathbb {H}_{\Lambda ,\Lambda ^\prime }$ -representation $\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ,\Lambda ^\prime }$ .

9 Metaplectic representations and metaplectic polynomials

In this subsection, we generalise the $\text {GL}_{r+1}$ -type metaplectic double affine Hecke algebra representation and the associated metaplectic polynomials from [Reference Sahi, Stokman and Venkateswaran38, §5] to arbitrary root systems using the uniform quasi-polynomial representation for extended double affine Hecke algebras (Theorem 8.19) and the corresponding uniform quasi-polynomials (Theorem 8.25). This section provides full proofs for the results announced in [Reference Sahi, Stokman and Venkateswaran38, §5].

The metaplectic double affine Hecke algebra action is an action on spaces $\mathcal {P}_\Lambda $ of polynomials on a torus $T_\Lambda $ , with $\Lambda \in \mathcal {L}$ . It depends on a metaplectic datum, which is a pair $(n,\mathbf {Q})$ with $n\in \mathbb {Z}_{>0}$ and $\mathbf {Q}: \Lambda \rightarrow \mathbb {Q}$ a nonzero $W_0$ -invariant quadratic form that restricts to an integer-valued quadratic form on $Q^{\vee }$ . The metaplectic datum leads to a finite root system $\Phi _0^m$ , the so-called metaplectic root system, which is isomorphic to either $\Phi _0$ or $\Phi _0^{\vee }$ (see Subsection 9.1). Its associated co-root lattice $Q^{m\vee }$ is contained in  $Q^{\vee }$ .

The metaplectic representation will be realised as a sub-representation of the uniform quasi-polynomial representation relative to the metaplectic root system $\Phi _0^m$ , with the parameters $(\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c})$ fixed in such a way that the twisted $W_0\ltimes Q^{m\vee }$ -action $\cdot _{\widehat {\mathbf {p}},\mathfrak {c}}$ on $\mathcal {P}_\Lambda $ reduces to the standard action by q-dilation and reflection operators.

The associated uniform quasi-polynomials $\mathcal {E}_y(x;\mathbf {g},\widehat {\mathbf {p}})\in \mathcal {P}_\Lambda $ ( $y\in \Lambda $ ) will provide the metaplectic polynomials.

The $\mathbf {g}$ -dependence reduces to a dependence on so-called metaplectic parameters $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ which, in the context of representation theory of metaplectic covers of reductive groups over a non-archimedean local field, are certain Gauss sums (see, for example, [Reference Chinta, Gunnells and Puskas15, Reference Chinta and Gunnells14]).

In the remainder of the section, we fix a metaplectic datum $(n,\mathbf {Q})$ , and we assume that $\mathbf {Q}$ takes positive values on co-roots. We write $\kappa _\ell \in \mathbb {Z}_{>0}$ for the natural number

(9.1) $$ \begin{align} \kappa_\ell:=\mathbf{Q}(\alpha^{\vee})\qquad\quad (\alpha\in{}^\ell\Phi_0) \end{align} $$

for $\ell \in \{\text {sh},\text {lg}\}$ .

9.1 The metaplectic parameters

Write $\mathbf {B}: \Lambda \times \Lambda \rightarrow \mathbb {Q}$ for the $W_0$ -invariant symmetric bilinear pairing

$$\begin{align*}\mathbf{B}(\lambda,\mu):=\mathbf{Q}(\lambda+\mu)-\mathbf{Q}(\lambda )-\mathbf{Q}(\mu)\qquad (\lambda,\mu\in \Lambda). \end{align*}$$

Then

(9.2) $$ \begin{align} \mathbf{B}(\lambda,\alpha^{\vee})=\mathbf{Q}(\alpha^{\vee})\alpha(\lambda)\qquad (\lambda\in \Lambda,\, \alpha\in \Phi_0) \end{align} $$

and

(9.3) $$ \begin{align} m(\alpha):=\frac{n}{\text{gcd}(n,\mathbf{Q}(\alpha^{\vee}))}=\frac{\text{lcm} (n,\mathbf{Q}(\alpha^{\vee}))}{\mathbf{Q}(\alpha^{\vee})} \qquad (\alpha\in\Phi_0) \end{align} $$

defines a $W_0$ -invariant $\mathbb {Z}_{>0}$ -valued function on $\Phi _0$ . The associated metaplectic root system is defined by

$$\begin{align*}\Phi_0^m:=\left\{\alpha^m\,\, | \,\, \alpha\in\Phi_0\right\}\,\,\text{ with }\,\, \alpha^m:=m(\alpha)^{-1}\alpha. \end{align*}$$

Then $\Phi _0^m\simeq \Phi _0$ if m is constant, and $\Phi _0^m\simeq \Phi _0^{\vee }$ otherwise (see [Reference Sahi, Stokman and Venkateswaran38]). In particular, the Weyl group of $\Phi _0^m$ is still $W_0$ . The fixed basis $\{\alpha _1,\ldots ,\alpha _r\}$ of $\Phi _0$ determines a basis $\{\alpha _1^m,\ldots ,\alpha _r^m\}$ of $\Phi _0^m$ . Note that $\Phi _0^{m\vee }=\{\alpha ^{m\vee }\}_{\alpha \in \Phi _0}$ , with $\alpha ^{m\vee }=m(\alpha )\alpha ^{\vee }$ . Write

$$\begin{align*}Q^{m\vee}:=\mathbb{Z}\Phi_0^{m\vee}\subseteq Q^{\vee} \end{align*}$$

for the coroot lattice of $\Phi _0^m$ .

Let $\Phi ^m:=\Phi _0^m\times \mathbb {Z}$ be the affine root system of $\Phi _0^m$ , with corresponding extended basis $\{\alpha _0^m,\alpha _1^m,\ldots ,\alpha _r^m\}$ . Then $\alpha _0^m=(-\vartheta ^m,1)$ with $\vartheta ^m\in \Phi _0^{m+}$ the highest root in $\Phi _0^{m+}$ . For the underlying root $\vartheta \in \Phi _0^+$ , this means that

$$ \begin{align*} \vartheta= \begin{cases} \varphi\quad &\mbox{ if } m \mbox{ is constant},\\ \theta\quad &\mbox{ otherwise}, \end{cases} \end{align*} $$

where $\theta \in \Phi _0^+$ is the highest short root.

Consider the affine Weyl group $W^m=W_0\ltimes Q^{m\vee }$ corresponding to the affine root system $\Phi ^m$ . The associated simple reflections are denoted by

$$\begin{align*}s_j^m:=s_{\alpha_j^m}\in W^m\qquad\quad (j\in [0,r]). \end{align*}$$

Then $s_i^m=s_i$ for $i\in [1,r]$ , and $s_0^m=\tau (\vartheta ^{m\vee })s_\vartheta $ . The fundamental alcove is now given by

$$\begin{align*}C_+^m:=\{y\in E \,\,\, | \,\,\, \alpha_i(y)>0\,\,\, (1\leq i\leq r)\,\,\, \& \,\,\, \vartheta(y)<m(\vartheta)\}. \end{align*}$$

Its faces are denoted by $C^{mJ}$ for $J\subsetneq [0,r]$ .

Let $\mathbf {k}: \Phi ^m\rightarrow \mathbf {F}^\times $ be a multiplicity function on $\Phi ^m$ . Its values at simple roots are denoted by $k_j:=k_{\alpha _j^m}$ for $j\in [0,r]$ . Write $\mathcal {G}^m$ for the g-parameter space $\mathcal {G}$ with respect to the root system $\Phi _0^m$ (see Remark 8.6).

Definition 9.1. We write $\mathcal {M}=\mathcal {M}_{(n,\mathbf {Q})}$ for the set of tuples

$$\begin{align*}\underline{h}:=(h_s(\alpha))_{s\in\mathbf{Q}(\alpha^{\vee})\mathbb{Z},\,\alpha\in\Phi_0} \end{align*}$$

with $h_s(\alpha )\in \mathbf {F}^\times $ satisfying

(9.4) $$ \begin{align} \begin{aligned} h_{s+\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))}(v\alpha)&=h_s(\alpha)\qquad\quad (v\in W_0,\, s\in\mathbf{Q}(\alpha^{\vee})\mathbb{Z}),\\ h_0(\alpha)&=-1,\\ h_s(\alpha)h_{-s}(\alpha)&=k_{\alpha^m}^{-2}\qquad\qquad (s\in\mathbf{Q}(\alpha^{\vee})\mathbb{Z}\setminus\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}) \end{aligned} \end{align} $$

for all $\alpha \in \Phi _0$ .

We call $\mathcal {M}_{(n,\mathbf {Q})}$ the set of metaplectic parameters relative to the metaplectic datum $(n,\mathbf {Q})$ . Note that for $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ ,

(9.5) $$ \begin{align} h_{\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))/2}(\alpha)\in\{\pm k_{\alpha^m}^{-1}\}\quad \mbox{ if }\,\, m(\alpha)\in 2\mathbb{Z} \end{align} $$

in view of (9.3).

Lemma 9.2. We have a surjective map

$$\begin{align*}\mathcal{G}^m\twoheadrightarrow\mathcal{M}_{(n,\mathbf{Q})},\qquad \mathbf{g} \mapsto\underline{h}^{\mathbf{g}} =(h_s^{\mathbf{g}}(\alpha))_{s\in\mathbf{Q}(\alpha^{\vee})\mathbb{Z},\,\alpha\in\Phi_0} \end{align*}$$

with $h_s^{\mathbf {g}}(\alpha )\in \mathbf {F}^\times $ defined by

(9.6) $$ \begin{align} h_s^{\mathbf{g}}(\alpha):=-k_{\alpha^m}^{-\chi_{\mathbb{Z}\setminus\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}}(s)} g_{\alpha^m}(s/\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))). \end{align} $$

Proof. This is a straightforward check, which is left to the reader.

Remark 9.3. For the metaplectic representation of the affine Hecke algebra defined in [Reference Sahi, Stokman and Venkateswaran38, Thm. 3.7], parameters $h_s(\alpha )$ ( $s\in \mathbb {Z}$ , $\alpha \in \Phi _0$ ) are used which satisfy

(9.7) $$ \begin{align} \begin{aligned} h_{s+n}(v\alpha)&=h_s(\alpha)\qquad\quad (v\in W_0, s\in\mathbb{Z}),\\ h_0(\alpha)&=-1,\\ h_s(\alpha)h_{-s}(\alpha)&=k_{\alpha^m}^{-2}\qquad\qquad\,\,\,\, (s\in\mathbb{Z}\setminus n\mathbb{Z}) \end{aligned} \end{align} $$

for all $\alpha \in \Phi _0$ (see [Reference Sahi, Stokman and Venkateswaran38, Def. 3.5]), although the metaplectic representation only depends on the parameters $h_s(\alpha )$ with $s\in \mathbf {Q}(\alpha ^{\vee })\mathbb {Z}$ and $\alpha \in \Phi _0$ , which naturally form a tuple of metaplectic parameters in $\mathcal {M}_{(n,\mathbf {Q})}$ . Conversely, note that a tuple $\underline {h}=(h_s(\alpha ))_{s\in \mathbf {Q}(\alpha ^{\vee })\mathbb {Z},\alpha \in \Phi _0}\in \mathcal {M}_{(n,\mathbf {Q})}$ of metaplectic parameters can be extended to a set of parameters $(h_s(\alpha ))_{s\in \mathbb {Z},\alpha \in \Phi _0}$ in such a way that (9.7) holds true. For instance, one can take the extension

$$ \begin{align*} \begin{aligned} h_{\text{gcd}(n,\mathbf{Q}(\alpha^{\vee}))s}(\alpha)&:=h_{\mathbf{Q}(\alpha^{\vee})s\ell_\alpha}(\alpha)\qquad\,\,\,\,\, (s\in\mathbb{Z}),\\ h_s(\alpha)&:=k_{\alpha^m}^{-1}\qquad\qquad\qquad\,\,\, (s\in\mathbb{Z}\setminus\text{gcd}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}) \end{aligned} \end{align*} $$

for $\alpha \in \Phi _0$ , where $\ell _\alpha \in \mathbb {Z}$ is such that $\text {gcd}(n,\mathbf {Q}(\alpha ^{\vee }))-\ell _\alpha \mathbf {Q}(\alpha ^{\vee })\in n\mathbb {Z}$ .

9.2 The metaplectic basic representation

In [Reference Sahi, Stokman and Venkateswaran38, Thm. 5.4], we introduced the metaplectic basic representation of the $\text {GL}_{r+1}$ double affine Hecke algebra. In this subsection, we introduce the metaplectic basic representation for extended double affine Hecke algebras for all types.

The metaplectic representation depends on q, $\mathbf {k}$ and on metaplectic parameters $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ . To keep the notations manageable, we will suppress the dependence on $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ from the notations.

Let $\mathbb {H}^m:=\mathbb {H}(\mathbf {k},q)$ be the double affine Hecke algebra associated to the root system $\Phi _0^m$ (see Definition 2.12), which we call the metaplectic double affine Hecke algebra. We denote by $\mathcal {L}^m$ the set of finitely generated abelian subgroups $\Lambda ^m\subset E$ such that

(9.8) $$ \begin{align} Q^{m\vee}\subseteq\Lambda^m\quad \&\quad \alpha(\Lambda^m)\subseteq m(\alpha)\mathbb{Z}\quad \forall\, \alpha\in\Phi_0. \end{align} $$

Note that it is the set $\mathcal {L}$ from Definition 2.8, with the role of $\Phi _0$ taken over by $\Phi _0^m$ .

We have the following supplement to Lemma 9.2.

Lemma 9.4. Let $\Lambda \in \mathcal {L}$ and $\mathbf {g}\in \mathcal {G}^m$ . For $\alpha \in \Phi _0$ and $\lambda \in \Lambda $ , we have

(9.9) $$ \begin{align} g_{\alpha^m}(\alpha^m(\lambda))=-k_{\alpha^m}^{1-\chi_{\mathbb{Z}}(\alpha(\lambda)/m(\alpha))}h_{\mathbf{Q}(\alpha^{\vee})\alpha(\lambda)}^{\mathbf{g}}(\alpha). \end{align} $$

Proof. By (9.3), we have

$$\begin{align*}\alpha^m(\lambda)= \frac{\mathbf{Q}(\alpha^{\vee})\alpha(\lambda)}{\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))}, \end{align*}$$

and $\alpha (\lambda )\in \mathbb {Z}$ since $\lambda \in \Lambda $ . Hence, we can apply (9.6) to obtain

$$\begin{align*}g_{\alpha^m}(\alpha^m(\lambda))=-k_{\alpha^m}^{\chi_{\mathbb{Z}\setminus\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}}(\mathbf{Q}(\alpha^{\vee})\alpha(\lambda))} h_{\mathbf{Q}(\alpha^{\vee})\alpha(\lambda)}^{\mathbf{g}}(\alpha). \end{align*}$$

The result now follows from the identity

$$\begin{align*}\chi_{\mathbb{Z}\setminus\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}}(\mathbf{Q}(\alpha^{\vee})\alpha(\lambda))=1- \chi_{\mathbb{Z}}(\alpha(\lambda)/m(\alpha)), \end{align*}$$

which follows from an elementary computation using (9.3).

Let $\Lambda ^{m\prime }\in \mathcal {L}^m$ such that $\langle \Lambda ,\Lambda ^{m\prime }\rangle \subseteq \mathbb {Z}$ . One can, for instance, take $\Lambda ^{m\prime }=Q^{m\vee }$ , since the inclusions $Q^{m\vee }\subseteq Q^{\vee }\subseteq Q$ imply that

$$\begin{align*}\langle \Lambda,Q^{m\vee}\rangle\subseteq\langle P^{\vee},Q^{m\vee}\rangle\subseteq\mathbb{Z}. \end{align*}$$

The metaplectic extended affine Weyl group $W_{\Lambda ^{m\prime }}=W_0\ltimes \Lambda ^{m\prime }$ acts on $\mathcal {P}_\Lambda $ by q-dilations and reflections,

$$\begin{align*}v(x^\lambda):=x^{v\lambda} \quad (v\in W_0),\qquad\quad \tau(\mu)(x^{\lambda}):=q^{-\langle\lambda,\mu\rangle}x^\lambda\quad (\mu\in \Lambda^{m\prime}) \end{align*}$$

for all $\lambda \in \Lambda $ (cf. Subsection 2.6 and Subsection 7.3). Note that $s_0^m(x^\lambda )=q_\vartheta ^{m(\vartheta )\vartheta (\lambda )}x^{s_\vartheta \lambda }$ .

Let $\ell \in \mathbb {Z}_{>0}$ and write $r_\ell (s)\in \{0,\ldots ,\ell -1\}$ for the remainder of $s\in \mathbb {Z}$ modulo $\ell $ . Define linear maps $\nabla _j^m: \mathcal {P}_\Lambda \rightarrow \mathcal {P}_\Lambda $ for $j\in [0,r]$ by

$$ \begin{align*} \begin{aligned} \nabla_0^m(x^\lambda)&:=\left(\frac{1-(q_\vartheta^{m(\vartheta)}x^{-\vartheta^{\vee}})^{(r_{m(\vartheta)}(-\vartheta(\lambda))+\vartheta(\lambda))}}{1-(q_\vartheta^{m(\vartheta)}x^{-\vartheta^{\vee}})^{m(\vartheta)}}\right)x^\lambda,\\ \nabla_i^m(x^\lambda)&:=\left(\frac{1-x^{(r_{m(\alpha_i)}(\alpha_i(\lambda))-\alpha_i(\lambda))\alpha_i^{\vee}}}{1-x^{m(\alpha_i)\alpha_i^{\vee}}}\right)x^\lambda \end{aligned} \end{align*} $$

for $\lambda \in \Lambda $ and $i\in [1,r]$ . We now have the following action of the metaplectic double affine Hecke algebra on the space $\mathcal {P}_\Lambda $ of polynomials on $T_\Lambda $ .

Theorem 9.5. Fix $\Lambda \in \mathcal {L}$ and $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ .

For $\Lambda ^m,\Lambda ^{m\prime }\in \mathcal {L}^m$ such that $\Lambda ^m\subseteq \Lambda $ and $\langle \Lambda ,\Lambda ^{m\prime }\rangle \subseteq \mathbb {Z}$ , the formulas

(9.10) $$ \begin{align} \begin{aligned} \pi^{m}_\Lambda(T_0)x^\lambda&:=-k_0h_{\mathbf{B}(\lambda,\vartheta^{\vee})}(\vartheta)q_{\vartheta}^{m(\vartheta)\vartheta(\lambda)}x^{s_\vartheta\lambda}+(k_0-k_0^{-1})\nabla_0^m(x^\lambda),\\ \pi^{m}_\Lambda(T_i)x^\lambda&:=-k_ih_{-\mathbf{B}(\lambda,\alpha_i^{\vee})}(\alpha_i)x^{s_i\lambda}+(k_i-k_i^{-1})\nabla_i^m(x^\lambda),\\ \pi^{m}_\Lambda(x^\mu)x^\lambda&:=x^{\lambda+\mu},\\ \pi^{m}_\Lambda(\omega)x^\lambda&:=\omega(x^\lambda) \end{aligned} \end{align} $$

for $\lambda \in \Lambda $ , $i\in [1,r]$ , $\mu \in \Lambda ^m$ and $\omega \in \Omega _{\Lambda ^{m\prime }}$ define a representation

$$\begin{align*}\pi^m_\Lambda: \mathbb{H}_{\Lambda^m,\Lambda^{m\prime}}\rightarrow\text{End}(\mathcal{P}_\Lambda), \end{align*}$$

where $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ is the extension of the metaplectic double affine Hecke algebra $\mathbb {H}^m$ by the lattices $(\Lambda ^m,\Lambda ^{m\prime })$ .

Proof. Without loss of generality, we may assume that $\mathbf {F}$ contains a $(2h)^{\text {th}}$ -root $q^{\frac {1}{2h}}$ of q. Write $\widehat {\mathcal {G}}^m$ for the set $\widehat {\mathcal {G}}$ of $\widehat {\mathbf {g}}$ -parameters relative to the root system $\Phi _0^m$ (Definition 8.4). For $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}^m$ , $\mathbf {g}\in \mathcal {G}^m$ and $\mathfrak {c}\in \mathcal {C}_{\Lambda ^m,\Lambda ^{m\prime }}$ , the uniform quasi-polynomial representation $\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ^m,\Lambda ^{\prime m}}: \mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}\rightarrow \text {End}(\mathbf {F}[E])$ relative to the finite root system $\Phi _0^m$ (see Theorem 8.19) contains $\mathcal {P}_\Lambda $ as a subrepresentation, because $\Lambda ^m\subseteq \Lambda $ . Denote

$$\begin{align*}\pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c};\Lambda}:=\pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}^{\Lambda^m,\Lambda^{m\prime}}(\cdot)\vert_{\mathcal{P}_\Lambda}: \mathbb{H}_{\Lambda^m,\Lambda^{m\prime}}\rightarrow\text{End}(\mathcal{P}_\Lambda) \end{align*}$$

for its representation map. Its defining formulas are

(9.11) $$ \begin{align} \begin{aligned} \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c};\Lambda}(T_j)x^\lambda&:=k_j^{\chi_{\mathbb{Z}}(\alpha_j^m(\lambda))}g_{D\alpha_j^m}(\alpha_j^m(\lambda))^{-1}s_j^m\cdot_{\widehat{\mathbf{p}},\mathfrak{c}} x^\lambda+(k_j-k_j^{-1})\nabla_j(x^\lambda),\\ \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c};\Lambda}(x^\mu)x^\lambda&:=x^{\lambda+\mu},\\ \pi_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c};\Lambda}(\omega)x^\lambda&:=\omega\cdot_{\widehat{\mathbf{p}},\mathfrak{c}}x^\lambda \end{aligned} \end{align} $$

for $\lambda \in \Lambda $ , $j\in [0,r]$ , $\mu \in \Lambda ^m$ and $\omega \in \Omega _{\Lambda ^{m\prime }}$ , with $\nabla _j$ the truncated divided difference operator (4.5) relative to the root system $\Phi _0^m$ .

Fix $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}^m$ and $\mathfrak {c}\in \mathcal {C}_{\Lambda ^m,\Lambda ^{\prime m}}$ such that

(9.12) $$ \begin{align} \widehat{p}_{\alpha^m}(z):=q_{\alpha}^{m(\alpha)^2z/h}\quad (z\in n^{-1}\mathbb{Z}),\qquad\quad \mathfrak{c}(\lambda):=q^\lambda\quad (\lambda\in\text{pr}_{E_{\text{co}}}(\Lambda)). \end{align} $$

For this specific choice, we have

(9.13) $$ \begin{align} q^\lambda=\mathfrak{t}_\lambda(\widehat{\mathbf{p}},\mathfrak{c})\qquad (\lambda\in\Lambda) \end{align} $$

in $T_{\Lambda ^{\prime m}}$ (compare with Example 8.11). In particular, $w\cdot _{\widehat {\mathbf {p}},\mathfrak {c}}(x^\lambda )=w(x^\lambda )$ for $w\in W_{\Lambda ^{m\prime }}$ and $\lambda \in \Lambda $ . The defining formulas (9.11) for the corresponding representation $\pi _{\mathbf {g};\Lambda }:=\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c};\Lambda }$ of $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ reduce to

(9.14) $$ \begin{align} \begin{aligned} \pi_{\mathbf{g};\Lambda}(T_j)x^\lambda&=k_j^{\chi_{\mathbb{Z}}(\alpha_j^m(\lambda))}g_{D\alpha_j^m}(\alpha_j^m(\lambda))^{-1} s_j^m(x^\lambda)+(k_j-k_j^{-1})\nabla_j(x^\lambda),\\ \pi_{\mathbf{g};\Lambda}(x^\mu)x^\lambda&=x^{\lambda+\mu},\\ \pi_{\mathbf{g};\Lambda}(\omega)x^\lambda&=\omega(x^\lambda) \end{aligned} \end{align} $$

for $j\in [0,r]$ , $\omega \in \Omega _{\Lambda ^{m\prime }}$ , $\mu \in \Lambda ^m$ and $\lambda \in \Lambda $ . We now show that (9.14) matches with the defining formulas for $\pi ^{m}_\Lambda $ under the correspondence (9.6) between $\mathbf {g}$ -parameters and metaplectic parameters.

Let $\mathbf {g}\in \mathcal {G}^m$ and set $\underline {h}:=\underline {h}^{\mathbf {g}}\in \mathcal {M}_{(n,\mathbf {Q})}$ in the defining formulas (9.10) for $\pi ^m_\Lambda $ . It is then clear that $\pi _{\mathbf {g};\Lambda }(p)=\pi ^{m}_\Lambda (p)$ for $p\in \mathcal {P}_{\Lambda ^m}$ , and $\pi _{\mathbf {g};\Lambda }(\omega )=\pi ^{m}_\Lambda (\omega )$ for $\omega \in \Omega _{\Lambda ^{m\prime }}$ . So it suffices to show that

(9.15) $$ \begin{align} \pi_{\mathbf{g};\Lambda}(T_j)=\pi^{m}_\Lambda(T_j)\qquad\quad (j\in [0,r]). \end{align} $$

To show this, first note that

$$\begin{align*}\nabla_j(\cdot)\vert_{\mathcal{P}_\Lambda}=\nabla_j^m \end{align*}$$

as endomorphisms of $\mathcal {P}_\Lambda $ , since $r_\ell (s)=s-\ell \lfloor \frac {s}{\ell }\rfloor $ for $s\in \mathbb {Z}$ and $\ell \in \mathbb {Z}_{>0}$ and

$$\begin{align*}\mathbf{B}(\lambda,\alpha^{\vee})=\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\alpha^m(\lambda) \end{align*}$$

for $\lambda \in \Lambda $ and $\alpha \in \Phi _0$ by (9.2) and (9.3) (compare with the proof of Lemma 9.4). Then (9.15) follows from

(9.16) $$ \begin{align} -k_{\alpha^m}h_{-\mathbf{B}(\lambda,\alpha^{\vee})}^{\mathbf{g}}(\alpha)=k_{\alpha^m}^{\chi_{\mathbb{Z}}(\alpha^m(\lambda))}g_{\alpha^m}(\alpha^m(\lambda))^{-1} \qquad (\alpha\in\Phi_0,\,\lambda\in\Lambda), \end{align} $$

which in turn is a consequence of (9.9) and (9.4).

Definition 9.6. We write $\mathcal {P}_\Lambda ^{m}$ for $\mathcal {P}_\Lambda $ endowed with the $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ -actions $\pi ^{m}_\Lambda $ from Theorem 9.5. We call $\pi ^m_\Lambda $ the metaplectic basic representation.

Remark 9.7.

  1. 1. The condition $\langle \Lambda ,\Lambda ^{m\prime }\rangle \subseteq \mathbb {Z}$ in Theorem 9.5 may be relaxed to $\langle \Lambda ,\Lambda ^{m\vee }\rangle \subseteq e^{-1}\mathbb {Z}$ when q has an $e^{th}$ root in $\mathbf {F}$ .

  2. 2. Suppose that $\Lambda _1,\Lambda _2\in \mathcal {L}$ with $\Lambda _1\subseteq \Lambda _2$ and that $\Lambda ^m,\Lambda ^{m\prime }\in \mathcal {L}^m$ such that $\Lambda ^m\subseteq \Lambda _1$ and $\langle \Lambda _2,\Lambda ^{m\prime }\rangle \subseteq \mathbb {Z}$ . Then

    $$\begin{align*}\pi_{\Lambda_2}^m(\cdot)\vert_{\mathcal{P}_{\Lambda_1}}=\pi_{\Lambda_1}^m \end{align*}$$
    as representations of $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ .

Corollary 9.8. Let $\Lambda \in \mathcal {L}$ and $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ . Let $\Lambda ^m,\Lambda ^{m\prime }\in \mathcal {L}^m$ such that $\Lambda ^m\subseteq \Lambda $ and $\langle \Lambda ,\Lambda ^{m\prime }\rangle \subseteq \mathbb {Z}$ . The inclusion map $\mathcal {P}_{\Lambda ^m}\hookrightarrow \mathcal {P}_{\Lambda }$ is a morphism

$$\begin{align*}\mathcal{P}_{\Lambda^m,1_{T_{\Lambda^{m\prime}}}}^{(0)}\hookrightarrow \mathcal{P}_{\Lambda}^m \end{align*}$$

of $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ -modules, (i.e., the $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ -submodule $(\mathcal {P}_{\Lambda ^m},\pi _\Lambda ^m(\cdot )\vert _{\mathcal {P}_{\Lambda ^m}})$ of $\mathcal {P}_\Lambda ^m$ is Cherednik’s polynomial representation of $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ ).

Proof. Note that $\nabla _j^m(\cdot )\vert _{\mathcal {P}_{\Lambda ^m}}$ is the standard divided difference operator associated to the simple root $\alpha _j^m\in \Phi ^m$ . Furthermore, for $\lambda \in \Lambda ^m$ and $\alpha \in \Phi _0$ , we have

$$\begin{align*}\mathbf{B}(\lambda,\alpha^{\vee})=\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\alpha^m(\lambda)\in\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z} \end{align*}$$

by (9.2) and (9.3); hence, $h_{\mathbf {B}(\lambda ,\alpha ^{\vee })}(\alpha )=-1$ (see Definition 9.1). Hence, the defining formulas (9.10) for $\pi _\Lambda ^m$ , restricted to $\mathcal {P}_{\Lambda ^m}$ , reduce to the defining formulas for $\pi _{0,1_{T_{\Lambda ^{m\prime }}}}^{\Lambda ^m,\Lambda ^{m\prime }}$ .

Remark 9.9.

  1. (1) The restriction of $\pi ^{m}_\Lambda $ to the affine Hecke algebra generated by $T_1, \dots , T_r$ and $x^{\nu }$ for $\nu \in Q^{m\vee }$ coincides with the metaplectic representation of the affine Hecke algebra from [Reference Sahi, Stokman and Venkateswaran38, Thm. 3.7] (be aware that the finite root system in [Reference Sahi, Stokman and Venkateswaran38], denoted by $\Phi $ , is $\Phi _0^{\vee }$ in the present paper).

  2. (2) We show in this remark that for the $\text {GL}_{r+1}$ root datum, Theorem 9.5 reduces to [Reference Sahi, Stokman and Venkateswaran38, Thm. 5.4]. Recall the notations from Subsection 7.6. Let $\Phi _0=\{\epsilon _i-\epsilon _j\}_{1\leq i\not =\leq j\leq r+1}$ be the root system of type $A_r$ with $\ell =1$ (i.e., the roots have squared length equal to $2$ ). As simple roots, we take $\alpha _i:=\epsilon _i-\epsilon _{i+1}$ for $i=1,\ldots ,r$ . Let $(n,\mathbf {Q})$ be an associated metaplectic datum. Write $\kappa =\kappa _{\text {lg}}\in \mathbb {Z}_{>0}$ for the value of $\mathbf {Q}$ at co-roots. The resulting multiplicity function m is identically equal to $n/\text {gcd}(n,\kappa )$ ; hence, the metaplectic root system $\Phi _0^m=m^{-1}\Phi _0$ is the root system of type $A_{r+1}$ from Subsection 7.6 with $\ell =m$ (i.e., the squared length of the roots is equal to $2/m^2$ ).

    Take $\Lambda =\bigoplus _{i=1}^{r+1}\mathbb {Z}\epsilon _i$ , which lies in $\mathcal {L}$ , and

    $$\begin{align*}(\Lambda^m,\Lambda^{m\prime})=(m\mathbb{Z}^{r+1},m\mathbb{Z}^{r+1})\in (\mathcal{L}^m)^{\times 2}. \end{align*}$$

    Then

    $$\begin{align*}\widetilde{\mathbb{H}}^m=\mathbb{H}_{m\mathbb{Z}^{r+1},m\mathbb{Z}^{r+1}} \end{align*}$$
    is the $\text {GL}_{r+1}$ -type double affine Hecke algebra of type $\text {GL}_{r+1}$ relative to the $A_r$ root system $\Phi _0^m$ . The resulting representation
    $$\begin{align*}\pi^{m}_{\mathbb{Z}^{r+1}}: \widetilde{\mathbb{H}}^m\rightarrow\text{End}(\mathcal{P}_{\mathbb{Z}^{r+1}}) \end{align*}$$
    is the metaplectic basic representation $\widehat {\pi }^{(n,\kappa )}$ from [Reference Sahi, Stokman and Venkateswaran38, Thm. 5.4].

The proof of Theorem 9.5 implies that the metaplectic basic representation $\pi ^m_\Lambda $ is a sub-representation of a uniform quasi-polynomial representation; hence, it is isomorphic to a direct sum of extended quasi-polynomial representations (see Theorem 8.19(2)). We now describe this isomorphism in more detail.

Let $\Lambda \in \mathcal {L}$ . The set

$$\begin{align*}X^m(\Lambda):=\Lambda\cap\overline{C}_+^m \end{align*}$$

parametrizes the $W^m$ -orbits in $\Lambda $ ,

(9.17) $$ \begin{align} \Lambda=\bigsqcup_{c\in X^m(\Lambda)}\mathcal{O}_c. \end{align} $$

Note that $X^m(\Lambda )=\Lambda _{\text {min}}$ when $\Lambda \in \mathcal {L}^m$ . Consider the associated decomposition

(9.18) $$ \begin{align} \mathcal{P}_\Lambda=\bigoplus_{c\in X^m(\Lambda)}\mathcal{P}^{(c)} \end{align} $$

as $\mathcal {P}_{Q^{m\vee }}$ -modules. In fact, (9.18) is a decomposition as $\mathbb {W}^m$ -modules, with $\mathbb {W}^m$ the double affine Weyl group associated to the metaplectic root system $\Phi _0^m$ ; see Proposition 2.11. For $\mathbf {g}\in \mathcal {G}^m$ , consider the $\mathcal {P}_{Q^{m\vee }}$ -linear automorphism

$$\begin{align*}\bigoplus_{c\in X^m(\Lambda)}\Gamma^{(c)}_{Q^{m\vee},\mathbf{g}} \end{align*}$$

of $\mathcal {P}_\Lambda $ . By Lemma 9.4 and Lemma 9.2, we can express it entirely in terms of the metaplectic parameters $\underline {h}^{\mathbf {g}}$ associated to $\mathbf {g}$ , leading to the following lemma.

In the following lemma, the elements $c_y\in \overline {C}_+^m$ and $v_y\in W_0$ (see Definition 2.3) are relative to the metaplectic root system $\Phi _0^m$ .

Lemma 9.10. For fixed $\underline {h}=(h_s(\alpha ))_{s,\alpha }\in \mathcal {M}_{(n,\mathbf {Q})}$ , the formulas

$$\begin{align*}\Gamma^m(x^{\lambda}):=\left(\prod_{\alpha\in\Pi(v_\lambda^{-1})}\big(-k_{\alpha^m}^{1-\chi_{\mathbb{Z}}(\alpha(c_\lambda)/m(\alpha))}h_{\mathbf{Q}(\alpha^{\vee})\alpha(c_\lambda)}(\alpha)\big) \right)x^{\lambda}\qquad\quad (\lambda\in\Lambda) \end{align*}$$

define a $\mathcal {P}_{Q^{m\vee }}$ -linear automorphism $\Gamma ^m=\Gamma ^m_{\underline {h}}$ of $\mathcal {P}_\Lambda $ satisfying

$$\begin{align*}\Gamma^m_{\underline{h}^{\mathbf{g}}}=\bigoplus_{c\in X^m(\Lambda)}\Gamma_{Q^{m\vee},\mathbf{g}}^{(c)}\qquad\quad (\mathbf{g}\in\mathcal{G}^m). \end{align*}$$

Proof. This follows from Lemma 8.16, Lemma 9.2 and Lemma 9.4.

For $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ and $\lambda \in \Lambda $ , write

(9.19) $$ \begin{align} \mathfrak{t}_\lambda(\underline{h}):=\prod_{\alpha\in\Phi_0^+}\left(-k_{\alpha^m}^{1-\chi_{\mathbb{Z}}(\alpha(\lambda)/m(\alpha))}h_{\mathbf{Q}(\alpha^{\vee})\alpha(\lambda)}(\alpha)\right)^{\frac{\alpha}{m(\alpha)}}, \end{align} $$

viewed as multiplicative character of $\Lambda ^{m\prime }$ for any $\Lambda ^{m\prime }\in \mathcal {L}^m$ . Note that

(9.20) $$ \begin{align} \mathfrak{t}_\lambda(\underline{h}^{\mathbf{g}})= \mathfrak{t}_\lambda(\mathbf{g})\qquad\quad(\mathbf{g}\in\mathcal{G}^m), \end{align} $$

where we use the definition (8.2) of $\mathfrak {t}_\lambda (\mathbf {f})$ relative to the metaplectic root system $\Phi _0^m$ . By Lemma 8.12(2) and Lemma 9.2, we conclude that for $c\in X^m(\Lambda )$ and $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ ,

(9.21) $$ \begin{align} \mathfrak{t}_c(\underline{h})\in T_{\mathbf{J}(c),\Lambda^{m\prime}}^{\text{red}}\,\,\mbox{ and }\,\, q^c, q^c\mathfrak{t}_c(\underline{h})\in {}^{\Lambda^m}T_{\Lambda^{m\prime}}^c. \end{align} $$

Corollary 9.11. Let $\Lambda \in \mathcal {L}$ and $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ .

  1. 1. The $\mathcal {P}_{Q^{m\vee }}$ -linear automorphism $\Gamma ^m$ of $\mathcal {P}_\Lambda $ defines an isomorphism

    $$\begin{align*}\mathcal{P}^{m}_\Lambda\overset{\sim}{\longrightarrow} \bigoplus_{c\in X^m(\Lambda)}\mathcal{P}_{q^c\mathfrak{t}_{c}(\underline{h})}^{(c)} \end{align*}$$
    of $\mathbb {H}^m$ -modules.
  2. 2. For $\Lambda ^m,\Lambda ^{m\prime }\in \mathcal {L}^m$ such that $\Lambda ^m\subseteq \Lambda $ and $\langle \Lambda ,\Lambda ^{m\prime }\rangle \subseteq \mathbb {Z}$ , we have

    (9.22) $$ \begin{align} \mathcal{P}^{m}_\Lambda\simeq \bigoplus_{c\in X^m(\Lambda,\Lambda^m)}\mathcal{P}_{\Lambda^m,q^c\mathfrak{t}_c(\underline{h})}^{(c)} \end{align} $$
    as $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ -modules, where $X^m(\Lambda ,\Lambda ^m)$ is a complete set of representatives of the $\Omega _{\Lambda ^{m}}$ -orbits in $X^m(\Lambda )$ .

Proof. Without loss of generality, we may assume that $\mathbf {F}$ contains a $(2h)^{th}$ root of q.

Let $\mathbf {g}\in \mathcal {G}^m$ such that $\underline {h}=\underline {h}^{\mathbf {g}}$ . Choose $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}^m$ and $\mathfrak {c}\in \mathcal {C}_{\Lambda ^m,\Lambda ^{m\prime }}$ satisfying (9.12). Then

(9.23) $$ \begin{align} \mathcal{P}_\Lambda^m=(\mathcal{P}_\Lambda,\pi^{\Lambda^m,\Lambda^{m\prime}}_{\mathbf{g},\widehat{\mathbf{p}},\mathfrak{c}}(\cdot)\vert_{\mathcal{P}_\Lambda}) \end{align} $$

as $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ -modules, by the proof of Theorem 9.5. Then Lemma 9.10 and Theorem 8.19(2) show that the automorphism $\Gamma ^m$ of $\mathcal {P}_\Lambda $ is an isomorphism

$$\begin{align*}\mathcal{P}_\Lambda^m\overset{\sim}{\longrightarrow}\bigoplus_{c\in X^m(\Lambda)}\mathcal{P}^{(c)}_{\mathfrak{t}_c(\widehat{\mathbf{g}},\mathfrak{c})} \end{align*}$$

of $\mathbb {H}^m$ -modules, where $\widehat {\mathbf {g}}:=\widehat {\mathbf {p}}\cdot \mathbf {g}$ . Then (1) follows from the fact that

(9.24) $$ \begin{align} \mathfrak{t}_\lambda(\widehat{\mathbf{g}},\mathfrak{c})=q^\lambda\mathfrak{t}_\lambda(\mathbf{g})=q^\lambda\mathfrak{t}_\lambda(\underline{h}^{\mathbf{g}})\qquad\quad (\lambda\in\Lambda) \end{align} $$

by (9.13) and (9.20).

Part (2) follows in a similar manner, now using that $X^m(\Lambda ,\Lambda ^m)$ parametrizes the $W_{\Lambda ^m}$ -orbits in $\Lambda $ , leading to the decomposition

$$\begin{align*}\mathcal{P}_\Lambda=\bigoplus_{c\in X^m(\Lambda,\Lambda^m)}\mathcal{P}_{\Lambda^m}^{(c)} \end{align*}$$

of $\mathcal {P}_\Lambda $ as $\mathcal {P}_{\Lambda ^m}$ -modules. The isomorphism

$$\begin{align*}\mathcal{P}^{m}_\Lambda\overset{\sim}{\longrightarrow} \bigoplus_{c\in X^m(\Lambda,\Lambda^m)}\mathcal{P}_{\Lambda^m,q^c\mathfrak{t}_c(\underline{h})}^{(c)} \end{align*}$$

of $\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}$ -modules is then realised by $\bigoplus _{c\in X^m(\Lambda ,\Lambda ^m)}\Gamma ^{(c)}_{\Lambda ^m,\mathbf {g}}$ , where $\Gamma _{\Lambda ^m,\mathbf {g}}^{(c)}$ is the $\mathcal {P}_{\Lambda ^m}$ -linear automorphism of $\mathcal {P}_{\Lambda ^m}^{(c)}$ from Lemma 8.16 and $\mathbf {g}\in \mathcal {G}^m$ is such that $\underline {h}=\underline {h}^{\mathbf {g}}$ .

Remark 9.12. Remark 7.19 and the fact that $\Gamma ^m$ restricts to the identity on $\mathcal {P}_{\Lambda ^m}$ show that Corollary 9.11(2) is consistent with Corollary 9.8.

Corollary 9.11(2) and Proposition 7.25 allow to add twist parameters to $\pi ^m_\Lambda : \mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}\rightarrow \text {End}(\mathcal {P}_\Lambda )$ , with the twist parameters encoding a choice of multiplicative character of $\Omega _{\Lambda ^{m\prime }}$ for each $W^m$ -orbit in $\mathcal {O}_\Lambda $ . We do not give full details here, since the metaplectic polynomials defined in Subsection 9.4 do not depend on these twist parameters.

We now look for the number of free $h_s(\alpha )$ -parameters in the metaplectic basic representation $\pi ^{m}_\Lambda : \mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}\rightarrow \text {End}(\mathcal {P}_\Lambda )$ . Consider the subgroup

$$\begin{align*}Z_{\Lambda,\text{lg}}^m:=\{\varphi(\lambda)\,\, \text{mod}\,\,m(\varphi)\mathbb{Z}\,\,\, | \,\,\, \lambda\in\Lambda\} \end{align*}$$

of $\mathbb {Z}/m(\varphi )\mathbb {Z}$ and the subgroup

$$ \begin{align*} Z_{\Lambda,\text{sh}}^m:= \begin{cases} \{\theta(\lambda)\,\, \text{mod}\,\,m(\theta)\mathbb{Z}\,\,\, | \,\,\, \lambda\in\Lambda\}\qquad &\mbox{ if }\, \theta\not=\varphi,\\ \{\overline{0}\}\qquad &\mbox{ if }\, \theta=\varphi \end{cases} \end{align*} $$

of $\mathbb {Z}/m(\theta )\mathbb {Z}$ , where $\overline {0}$ is the neutral element of $\mathbb {Z}/m(\theta )\mathbb {Z}$ . Both subgroups are independent of the lattices $\Lambda ^m$ and $\Lambda ^{m\prime }$ .

For $s\in \mathbb {Z}_{>0}$ , the $2$ -group $\mathbb {Z}_2$ acts on $\mathbb {Z}/s\mathbb {Z}$ by group automorphisms via sign changes. Any subgroup Z of $\mathbb {Z}/s\mathbb {Z}$ is $\mathbb {Z}_2$ -stable, and its subgroup $Z^{\text {inv}}$ of $\mathbb {Z}_2$ -fixed elements is of order $1$ if $[\mathbb {Z}/s\mathbb {Z}:Z]$ is odd and of order $2$ otherwise. This in particular applies to the subgroups $Z_{\Lambda ,\text {lg}}^m\subseteq \mathbb {Z}/m(\varphi )\mathbb {Z}$ and $Z_{\Lambda ,\text {sh}}^m\subseteq \mathbb {Z}/m(\theta )\mathbb {Z}$ .

Write

$$\begin{align*}Z_{\Lambda,\ell}^{m,\text{reg}}:=Z_{\Lambda,\ell}^m\setminus Z_{\Lambda,\ell}^{m,\text{inv}}\subseteq Z_{\Lambda,\ell}^m. \end{align*}$$

Then $Z_{\Lambda ,\text {sh}}^{m,\text {reg}}=\emptyset $ if $\Phi _0=\Phi _0^{\text {lg}}$ is simply laced. Let

$$\begin{align*}\widetilde{Z}_{\Lambda,\ell}^{m,\text{reg}}\subseteq Z_{\Lambda,\ell}^{m,\text{reg}} \end{align*}$$

be a complete set of representatives of the $\mathbb {Z}_2$ -orbits in $Z_\ell ^{m,\text {reg}}$ .

Proposition 9.13. The metaplectic basic representation $\pi ^m_\Lambda : \mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }}\rightarrow \text {End}(\mathcal {P}_\Lambda )$ depends on

$$\begin{align*}\sum_{\ell\in\{\text{sh},\text{lg}\}}\#\widetilde{Z}_{\Lambda,\ell}^{m,\text{reg}} \end{align*}$$

free parameters $\{h_{\kappa _\ell z}(\varphi _\ell ) \,\, | \,\, z\in \widetilde {Z}_{\Lambda ,\ell }^{m,\text {reg}}, \ell \in \{\text {sh},\text {lg}\}\}$ and on

$$\begin{align*}\sum_{\ell\in\{\text{sh},\text{lg}\}}\left(\#Z_{\Lambda,\ell}^{m,\text{inv}}-1\right) \end{align*}$$

choices of signs $\{ h_{\kappa _\ell z}(\varphi _\ell ) \,\, | \,\, z\in Z_{\Lambda ,\ell }^{m,\text {inv}}\setminus \{\overline {0}\}, \ell \in \{\text {sh},\text {lg}\}\}$ .

Proof. This follows from Definition 9.1, (9.5), Theorem 9.5, (9.2) and (9.3).

Note that the maximum number of signs in $\pi _\Lambda ^m$ is $2$ .

Remark 9.14. By Proposition 9.13, the metaplectic representation $\pi _{\Lambda }^{m}$ can be defined over the field $\mathbf {F}=\mathbb {Q}\big (q,\mathbf {k}^{\frac {1}{2}},\check {h}\big )$ , where $\check {h}$ stands for the parameters $h_{\kappa _\ell z}(\varphi _\ell )$ ( $z\in \widetilde {Z}_{\Lambda ,\ell }^{m,\text {reg}}$ , $\ell \in \{\text {sh},\text {lg}\}$ ) viewed as indeterminates.

Remark 9.15. Consider the $\text {GL}_{r+1}$ metaplectic basic representation $\pi _{\mathbb {Z}^{r+1}}^m: \widetilde {\mathbb {H}}\rightarrow \text {End}(\mathcal {P}_{\mathbb {Z}^{r+1}})$ (see Remark 9.9(2)). In this case, the multiplicity function $\mathbf {k}$ is a single scalar $k\in \mathbf {F}^\times $ , $Z_{\text {sh}}^m=\{\overline {0}\}$ and

$$\begin{align*}Z_{\text{lg}}^m=\mathbb{Z}/m\mathbb{Z}; \end{align*}$$

hence, $Z_{\text {lg}}^{m,\text {reg}}=Z_{\text {lg}}^m\setminus \{\overline {0}\}$ if m is odd and $Z_{\text {lg}}^{m,\text {reg}}=Z_{\text {lg}}^m\setminus \{\overline {0}, \overline {m/2}\}$ if m is even. As representatives for the $\mathbb {Z}_2$ -orbits in $Z_\ell ^{m,\text {reg}}$ , we can take

$$ \begin{align*} \widetilde{Z}_{\text{lg}}^m=\{\overline{\ell}\,\, | \,\, \ell\in\mathbb{Z}:\,\, 1\leq \ell<\tfrac{m}{2}\}\subset \mathbb{Z}/m\mathbb{Z}, \end{align*} $$

which is empty if $m\leq 2$ . We conclude that the representation $\pi _{\mathbb {Z}^{r+1}}^{m}$ does not have free metaplectic parameters if $m\leq 2$ , and it has $\lfloor \frac {m-1}{2}\rfloor $ free metaplectic parameters if $m>2$ . Furthermore, $\pi _{\mathbb {Z}^{r+1}}^m$ depends on a sign $\epsilon \in \{\pm 1\}$ when m is even. This is in accordance with [Reference Sahi, Stokman and Venkateswaran38, §5].

9.3 A metaplectic affine Weyl group action on rational functions

Denote by $\mathcal {Q}^m$ the quotient field of $\mathcal {P}_{Q^{m\vee }}$ . For a finitely generated abelian subgroup $L\subset E$ , let $\mathbf {F}_{\mathcal {Q}^m}[L]$ be the $\mathcal {Q}^m$ -subalgebra of $\mathbf {F}_{\mathcal {Q}^m}[E]$ generated by $x^\lambda $ ( $\lambda \in L$ ). In view of (9.16), we can now formulate the following metaplectic version of Theorem 8.27.

Corollary 9.16. For $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ , the formulas

$$ \begin{align*} \begin{aligned} \sigma^{m}_\Lambda(s_0^m)(x^\lambda)&:=-k_0h_{\mathbf{B}(\lambda,\vartheta^{\vee})}(\vartheta) \left(\frac{x^{\alpha_0^{m\vee}}-1}{k_0x^{\alpha_0^{m\vee}}-k_0^{-1}}\right)q_\vartheta^{m(\vartheta)\vartheta(\lambda)}x^{s_\vartheta\lambda}\\&\qquad\qquad\qquad\qquad +\left(\frac{k_0-k_0^{-1}}{k_0x^{\alpha_0^{m\vee}}-k_0^{-1}}\right) x^{\lambda-\lfloor -\vartheta^m(\lambda)\rfloor \alpha_0^{m\vee}},\\ \sigma^{m}_\Lambda(s_i^m)(x^\lambda)&:=-k_ih_{-\mathbf{B}(\lambda,\alpha_i^{\vee})}(\alpha_i) \left(\frac{x^{\alpha_i^{m\vee}}-1}{k_ix^{\alpha_i^{m\vee}}-k_i^{-1}}\right)x^{s_i\lambda}+ \left(\frac{k_i-k_i^{-1}}{k_ix^{\alpha_i^{m\vee}}-k_i^{-1}}\right) x^{\lambda-\lfloor \alpha_i^m(\lambda)\rfloor \alpha_i^{m\vee}},\\ \sigma_\Lambda^{m}(\omega)(x^\lambda)&:=\omega(x^\lambda),\qquad \sigma_\Lambda^{m}(f)x^\lambda:=fx^\lambda \end{aligned} \end{align*} $$

for $\lambda \in \Lambda $ , $1\leq i\leq r$ , $\omega \in \Omega _{\Lambda ^{m\prime }}$ and $f\in \mathbf {F}_{\mathcal {Q}^m}[\Lambda ^m]$ define a representation

$$\begin{align*}\sigma^m_\Lambda: W_{\Lambda^{m\prime}}\ltimes\mathbf{F}_{\mathcal{Q}^m}[\Lambda^m]\rightarrow \text{End}_{\mathbf{F}}(\mathbf{F}_{\mathcal{Q}^m}[\Lambda]). \end{align*}$$

Note that $\sigma _\Lambda ^m=\pi _\Lambda ^{m,X-\text {loc}}\circ \beth $ ; cf. Remark 4.10.

The restriction of $\sigma _\Lambda ^{m}$ to $W_0\ltimes \mathbf {F}_{\mathcal {Q}^m}[\Lambda ^m]$ reduces to [Reference Sahi, Stokman and Venkateswaran38, Prop. 3.19(iii)] (with the role of the root system replaced by the dual root system). Let $\rho ^{\vee }$ and $\rho ^{m\vee }$ be the half-sum of positive co-roots for $\Phi _0$ and $\Phi _0^m$ , respectively. Then the conjugated $W_0$ -action

(9.25) $$ \begin{align} \sigma^{\text{CG}}_\Lambda(v)f:=x^{\rho^{\vee}-\rho^{m\vee}}\sigma_{\Lambda+P^{\vee}}^m(v)\big(x^{-\rho^{\vee}+\rho^{m\vee}}f\big)\qquad (v\in W_0,\, f\in\mathbf{F}_{\mathcal{Q}^m}[\Lambda]) \end{align} $$

reduces to the Chinta-Gunnells’ [Reference Chinta and Gunnells13, Reference Chinta and Gunnells14] action when the multiplicity functions $\mathbf {k}$ and $h_s$ are constant; see [Reference Sahi, Stokman and Venkateswaran38, Thm. 3.21 & Cor. 3.22] (the right-hand side of (9.25) is well defined, since $\mathbf {Q}$ has a unique extension to a $W_0$ -invariant rational quadratic form on $\Lambda +P^{\vee }$ ). This follows from the explicit formula

(9.26) $$ \begin{align} \begin{aligned} \sigma^{\text{CG}}_\Lambda(s_i)(fx^\lambda)&=k_i^2h_{\mathbf{Q}(\alpha_i^{\vee})-\mathbf{B}(\lambda,\alpha_i^{\vee})}(\alpha_i) \left(\frac{1-x^{-\alpha_i^{m\vee}}}{1-k_i^2x^{\alpha_i^{m\vee}}}\right)(s_if)x^{\alpha_i^{\vee}+s_i\lambda}\\ &\quad+\left(\frac{1-k_i^2}{1-k_i^2x^{\alpha_i^{m\vee}}}\right)(s_if)x^{\lambda+\lfloor -(\lambda,\alpha_i^{m\vee})\rfloor\alpha_i^{m\vee}} \end{aligned} \end{align} $$

for $i\in [1,r]$ , $f\in \mathbf {F}_{\mathcal {Q}^m}[\Lambda ^m]$ and $\lambda \in \Lambda $ , which easily follows from Corollary 9.16 (see [Reference Sahi, Stokman and Venkateswaran38, Cor. 3.22]). Hence,

(9.27) $$ \begin{align} \begin{aligned} \mathcal{T}_i^m(f):=&-f+\left(\frac{1-k_i^2x^{\alpha_i^{m\vee}}}{1-x^{\alpha_i^{m\vee}}}\right)\big(f-x^{\alpha_i^{m\vee}}\sigma_\Lambda^{\text{CG}}(s_i)(f)\big)\\ =&-f+\left(\frac{1-k_i^2x^{\alpha_i^{m\vee}}}{1-x^{\alpha_i^{m\vee}}}\right)\big(f-x^{\rho^{\vee}}\sigma_\Lambda^m(s_i)(fx^{-\rho^{\vee}})\big)\\ =&-k_ix^{\rho^{\vee}}\pi_{\Lambda}^{m,X-\text{loc}}(T_i^{-1})(fx^{-\rho^{\vee}}) \end{aligned} \end{align} $$

for $i\in [1,r]$ and $f\in \mathbf {F}_{\mathcal {Q}^m}[\Lambda ]$ are the metaplectic Demazure-Lusztig operators from [Reference Chinta, Gunnells and Puskas15, (11)] when the multiplicity functions $\mathbf {k}$ and $h_s$ are constant (see Remark 4.11 for the third equality in (9.27)). By the third equality in (9.27), the ( $W_0,\{s_1,\ldots ,s_r\})$ -braid relations of the metaplectic Demazure-Lusztig operators (see, for example, [Reference Chinta, Gunnells and Puskas15, Prop. 7]) are a consequence of Theorem 9.5. In fact, Corollary 9.16 and Theorem 9.5 extend the Chinta-Gunnells $W_0$ -action and the $H_0$ -action by metaplectic Demazure-Lusztig operators to actions of the affine Weyl group W and the affine Hecke algebra H, respectively. Including the action of $\mathcal {P}$ by multiplication operators, they promote to actions of the double affine Weyl group $\mathbb {W}$ and the double affine Hecke algebra $\mathbb {H}$ , respectively.

Remark 9.17. Recently, the Chinta-Gunnells $W_0$ -action (9.25) and the associated $H_0$ -action by metaplectic Demazure-Lusztig operators have been generalised to arbitrary Coxeter groups in [Reference Patnaik and Puskas35, §3.3]. In case of the affine Weyl group $(W,\{s_0,\ldots ,s_r\})$ , we expect that the resulting representations of W and H (see [Reference Patnaik and Puskas35, Prop. 3.3.4]) are isomorphic to $\sigma _\Lambda ^m\vert _W$ and $\pi _\Lambda ^m\vert _{H}$ with appropriate choices of $\mathbf {k}$ and $\underline {h}$ .

9.4 Metaplectic polynomials

In this subsection, we use the quasi-polynomial eigenfunctions $E_y^J(x;\mathfrak {t})$ from Theorem 6.3 to construct the root system generalisations of the metaplectic polynomials from [Reference Sahi, Stokman and Venkateswaran38, Thm. 5.7].

Let $\Lambda \in \mathcal {L}$ . The torus elements $\mathfrak {s}_\lambda $ ( $\lambda \in \Lambda $ ) relative to the metaplectic root system $\Phi _0^m$ are given by

(9.28) $$ \begin{align} \mathfrak{s}_\lambda:=\prod_{\alpha\in\Phi_0^+}k_{\alpha^m}^{\eta(\alpha^m(\lambda))\alpha^m}= \prod_{\alpha\in\Phi_0^+}\left(k_{\alpha^m}^{\chi_{\mathbb{Z}_{>0}}(\alpha(\lambda)/m(\alpha))-\chi_{\mathbb{Z}_{\leq 0}}(\alpha(\lambda)/m(\alpha))}\right)^{\frac{\alpha}{m(\alpha)}}; \end{align} $$

see (5.2). We will view $\mathfrak {s}_\lambda $ and $q^\lambda \mathfrak {s}_\lambda \mathfrak {t}_\lambda (\underline {h})$ as multiplicative character of $\Lambda ^{m\prime }$ for any lattice $\Lambda ^{m\prime }\in \mathcal {L}^m$ . Note that by Proposition 5.4, Lemma 8.8(4) and (9.20), we have for $c\in X^m(\Lambda )$ ,

(9.29) $$ \begin{align} w(q^c\mathfrak{s}_c\mathfrak{t}_c(\underline{h}))= q^{wc}\mathfrak{s}_{wc}\mathfrak{t}_{wc}(\underline{h})\qquad\quad (w\in W^m). \end{align} $$

Theorem 9.18. Let $\Lambda \in \mathcal {L}$ and let $\mathcal {O}$ be a $W^m$ -orbit in $\Lambda $ . Choose metaplectic parameters $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ such that in $T_{Q^{m\vee }}$ ,

(9.30) $$ \begin{align} q^\lambda\mathfrak{s}_\lambda\mathfrak{t}_\lambda(\underline{h})\not=q^{\lambda^\prime}\mathfrak{s}_{\lambda^\prime}\mathfrak{t}_{\lambda^\prime}(\underline{h})\qquad\mbox{ if }\,\,\lambda,\lambda^\prime\in\mathcal{O}\,\, \mbox{ and }\,\, \lambda\not=\lambda^\prime. \end{align} $$

For all $\lambda \in \mathcal {O}$ , we then have

  1. 1. There exists a unique simultaneous eigenfunction $E_\lambda ^m(x)\in \mathcal {P}_\Lambda $ of the commuting operators $\pi ^m_\Lambda (Y^\mu )$ ( $\mu \in Q^{m\vee }$ ) satisfying

    $$\begin{align*}E^m_\lambda(x)=x^\lambda+\text{l.o.t.} \end{align*}$$
  2. 2. For all $\Lambda ^{m\prime }\in \mathcal {L}^m$ satisfying $\langle \Lambda ,\Lambda ^{m\prime }\rangle \subseteq \mathbb {Z}$ , we have

    $$\begin{align*}\pi_\Lambda^m(Y^\mu)E^m_\lambda=\big(q^\lambda\mathfrak{s}_\lambda\mathfrak{t}_\lambda(\underline{h})\big)^{-\mu}E^m_\lambda\qquad \forall\, \mu\in\Lambda^{m\prime}. \end{align*}$$
  3. 3. We have

    $$\begin{align*}\Gamma^m\big(E^m_\lambda(x)\big)=\left(\prod_{\alpha\in\Pi(v_\lambda^{-1})}\big(-k_{\alpha^m}^{1-\chi_{\mathbb{Z}}(\alpha(c_\lambda)/m(\alpha))} h_{\mathbf{Q}(\alpha^{\vee})\alpha(c_\lambda)}(\alpha)\big) \right)E_\lambda^{\mathbf{J}(c_\lambda)}(x; q^{c_\lambda}\mathfrak{t}_{c_\lambda}(\underline{h})\vert_{Q^{m\vee}}), \end{align*}$$
    where $c_\lambda \in \overline {C}_+^m$ and $v_\lambda \in W_0$ (see Definition 2.3) are relative to the metaplectic root system $\Phi _0^m$ .

Proof. Note that under the assumption on $\Lambda ^{m\prime }\in \mathcal {L}^m$ as stated in part (2), the metaplectic representation $\pi _\Lambda ^m: \mathbb {H}_{Q^{m\vee },\Lambda ^{m\prime }}\rightarrow \text {End}(\mathcal {P}_\Lambda )$ is well defined since $Q^{m\vee }\subseteq \Lambda $ .

(1) This follows from Corollary 9.11(1) and Theorem 6.3.

(2) This is a direct consequence of part (1), Theorem 7.26(2)&(3) and (9.29).

(3) This follows from Corollary 9.11(1). The normalisation factor can be derived from Lemma 9.10.

Remark 9.19. Suppose that $\mathbf {F}$ contains a $(2h)^{th}$ root $q^{\frac {1}{2h}}$ of q. Let $\mathbf {g}\in \mathcal {G}^m$ such that $\underline {h}=\underline {h}^{\mathbf {g}}$ , and choose $\widehat {\mathbf {p}}\in \widehat {\mathcal {G}}^m$ satisfying (9.12). Then

$$\begin{align*}E^m_\lambda(x)=\mathcal{E}_\lambda(x;\mathbf{g},\widehat{\mathbf{p}}), \end{align*}$$

which follows from the proof of Corollary 9.11.

We call $E_\lambda ^m\in \mathcal {P}_\Lambda $ the monic metaplectic polynomial of degree $\lambda \in \Lambda $ relative to the metaplectic datum $(n,\mathbf {Q})$ and the choice of lattice $\Lambda \in \mathcal {L}$ , which we suppress from the notations.

It follows from Theorem 9.18 that the metaplectic basic representation $\pi ^m: \mathbb {H}\rightarrow \text {End}(\mathcal {P}_\Lambda )$ is Y-semisimple if (9.30) holds true for all W-orbits $\mathcal {O}=\mathcal {O}_c$ ( $c\in X^m(\Lambda )$ ) in $\Lambda $ . Its Y-spectrum then consists of the multiplicative characters

(9.31) $$ \begin{align} q^{\lambda}\mathfrak{s}_\lambda\mathfrak{t}_\lambda(\underline{h})=q^\lambda\prod_{\alpha\in\Phi_0^+}\left( -k_{\alpha^m}^{1-2\chi_{\mathbb{Z}_{\leq 0}}(\alpha(\lambda)/m(\alpha))}h_{\mathbf{Q}(\alpha^{\vee})\alpha(\lambda)}(\alpha)\right)^{\frac{\alpha}{m(\alpha)}} \qquad\quad (\lambda\in\Lambda) \end{align} $$

(the explicit formula follows from (9.28) and (9.19)). Note that by (9.21), we have for $c\in X^m(\Lambda )$ ,

$$\begin{align*}q^c\mathfrak{s}_c\mathfrak{t}_c(\underline{h})\in L_{\mathbf{J}(c)},\qquad\quad q^c,q^c\mathfrak{t}_c(\underline{h})\in T_{\mathbf{J}(c)},\qquad\quad \mathfrak{t}_c(\underline{h})\in T_{\mathbf{J}(c)}^{\text{red}}. \end{align*}$$

Remark 9.20. To compare (9.31) with the spectrum of the $\text {GL}_{r+1}$ -type metaplectic polynomials in [Reference Sahi, Stokman and Venkateswaran38, §5.4], note that

$$\begin{align*}q^{\lambda}\mathfrak{s}_\lambda\mathfrak{t}_\lambda(\underline{h})=q^\lambda\prod_{\alpha\in\Phi_0^+}\left(\sigma_{\mathbf{Q}(\alpha^{\vee})\alpha(\lambda)}^{(\underline{h})}(\alpha)\right)^{-\frac{\alpha}{m(\alpha)}} \end{align*}$$

with $\sigma _s^{(\underline {h})}(\alpha )$ defined by

$$ \begin{align*} \sigma_s^{(\underline{h})}(\alpha):= \begin{cases} k_{\alpha^m}^{-1}\qquad &\mbox{if }\,\, s\in\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}_{>0},\\ -k_{\alpha^m}h_{-s}(\alpha)\qquad &\mbox{if }\,\, s\in\mathbf{Q}(\alpha^{\vee})\mathbb{Z}\setminus\text{lcm}(n,\mathbf{Q}(\alpha^{\vee}))\mathbb{Z}_{>0}. \end{cases} \end{align*} $$

This follows by a direct computation using (9.3) and (9.4).

We conclude this section by considering the type $A_r$ metaplectic polynomials in a bit more detail. We use the notations from Remark 9.9(2). We thus have the explicit realisation of the $A_r$ -type root system $\Phi _0$ in $\mathbb {R}^{r+1}$ as $\{\epsilon _i-\epsilon _j\}_{i\not =j}$ with $\{\epsilon _i\}_{i=1}^{r+1}$ the standard orthonormal basis of $\mathbb {R}^{r+1}$ , and we have a metaplectic datum $(n,\mathbf {Q})$ giving rise to the constant metaplectic multiplicity function $m=n/\text {gcd}(n,\kappa )\in \mathbb {Z}_{>0}$ .

Take $\Lambda =\mathbb {Z}^{r+1}\in \mathcal {L}$ , so that

$$\begin{align*}\mathcal{P}_{\mathbb{Z}^{r+1}}=\mathbf{F}[x_1^{\pm 1},\ldots,x_{r+1}^{\pm 1}] \end{align*}$$

with $x_i:=x^{\epsilon _i}$ for $i\in [1,r+1]$ . The resulting $A_{r}$ -type monic metaplectic polynomials

$$\begin{align*}E_\lambda^m(x)\in\mathcal{P}_{\mathbb{Z}^{r+1}}=\mathbf{F}[x_1^{\pm 1},\ldots,x_{r+1}^{\pm 1}] \end{align*}$$

of degree $\lambda \in \mathbb {Z}^{r+1}$ then satisfies

(9.32) $$ \begin{align} \pi_{\mathbb{Z}^{r+1}}^m(Y^\mu)E_\lambda^m=\big(q^\lambda\mathfrak{s}_\lambda\mathfrak{t}_\lambda(\underline{h})\big)^{-\mu}E_\lambda^m\qquad\quad\forall\, \mu\in m\mathbb{Z}^{r+1}, \end{align} $$

with $\pi _{\mathbb {Z}^{r+1}}^m: \widehat {\mathbb {H}}=\mathbb {H}_{m\mathbb {Z}^{r+1},m\mathbb {Z}^{r+1}}\rightarrow \text {End}(\mathcal {P}_{\mathbb {Z}^{r+1}})$ the $\text {GL}_{r+1}$ metaplectic basic representation. The Coxeter-type expressions for the Y-elements are determined by the formulas

$$\begin{align*}Y^{m\epsilon_i}=T_{i-1}^{-1}\cdots T_2^{-1}T_1^{-1}uT_r\cdots T_{i+1}T_i\in\widetilde{\mathbb{H}}\qquad\quad (1\leq i\leq r+1), \end{align*}$$

which follows from (7.47) with $\ell =m$ .

Remark 9.21. When all the free parameters are formal (i.e., when $\mathbf {F}=\mathbb {Q}(q,\mathbf {k}^{\frac {1}{2}},\check {h})$ (see Remark 9.14)), then the generic condition (9.30) is valid for all W-orbits $\mathcal {O}$ in $\mathbb {Z}^{r+1}$ , and hence, the commuting operators $\pi _{\mathbb {Z}^{r+1}}^m(Y^\mu )$ ( $\mu \in m\mathbb {Z}^{r+1}$ ) are simultaneously diagonalisable. By inspection of the spectrum (see (9.31)), their common eigenspaces are one-dimensional. In this situation, $E_\lambda ^{m}(x)\in \mathcal {P}_{\mathbb {Z}^{r+1}}$ can be characterised as the unique Laurent polynomial $\sum _{\lambda ^\prime \in \mathbb {Z}^{r+1}}d_{\lambda ^\prime } x^{\lambda ^\prime }$ satisfying the eigenvalue equations (9.32) and satisfying $d_\lambda =1$ . This proves [Reference Sahi, Stokman and Venkateswaran38, Thm. 5.7] and shows that $E_\lambda ^m(x)$ is the $\text {GL}_{r+1}$ metaplectic polynomial $E_\lambda ^{(m)}(x)$ from [Reference Sahi, Stokman and Venkateswaran38, §5.4].

10 Limits to metaplectic Iwahori-Whittaker functions

In this section, we show that the Whittaker limits of the metaplectic polynomials $E_\lambda ^m(x)$ and their anti-symmetric versions produce metaplectic Iwahori-Whittaker and metaplectic spherical functions, respectively. The key point is to relate formula (6.52) for the Whittaker limit $\overline {E}_y^J(x)$ of $E_y^J(x;\mathfrak {t})$ with the expression of the metaplectic Iwahori-Whittaker function in terms of metaplectic Demazure-Lusztig operators obtained by Patnaik and Puskas in [Reference Patnaik and Puskas34, (6.48)].

We freely use the notations from the previous section and from Subsection 6.8. In particular, we assume that the ground field $\mathbf {F}$ is $\mathbf {K}(q^{\frac {1}{2h}})$ with $\mathbf {K}$ of characteristic zero and that the multiplicity functions $\mathbf {k}$ and $h_s$ take value in $\mathbf {K}^\times $ . We write $\overline {\mathcal {M}}_{(n,\mathbf {Q})}$ for the subset of parameters $\underline {h}\in \mathcal {M}_{(n,\mathbf {Q})}$ with multiplicity functions $h_s$ taking values in $\mathbf {K}^\times $ . We furthermore assume that $\Lambda \in \mathcal {L}$ contains $P^{\vee }$ .

In this section, we work in the metaplectic context. We thus have a fixed metaplectic datum $(n,\mathbf {Q})$ , and the quasi-polynomials $E_y^J(x;\mathfrak {t})$ and $E_y^{J,-}(x;\mathfrak {t})$ and their Whittaker limits $\overline {E}_y^J(x)$ and $\overline {E}_y^{J,-}(x)$ occurring in this section will be relative to the metaplectic affine root system $\Phi ^m$ and a choice of a multiplicity function $\mathbf {k}$ on $\Phi ^m$ . Recall that the Whittaker limit boils down to specialising coefficients at $q^{-\frac {1}{2h}}=0$ (see Subsection 6.8).

Remark 10.1. The link to representation metaplectic Whittaker functions for metaplectic covers of reductive groups requires the field $\mathbf {K}$ to be a non-archimedean local field containing the $n^{\text {th}}$ roots of unity, and with the cardinality of its residue field equal to $1$ modulo $2n$ (here, n is the positive integer from the metaplectic datum $(n,\mathbf {Q})$ ). In this case, the multiplicity function $\mathbf {k}$ takes on the constant value k with $k^{-2}$ equal to the cardinality of the residue field. The multiplicity functions $h_s$ of the metaplectic paramaters $\underline {h}\in \overline {\mathcal {M}}_{(n,\mathbf {Q})}$ are also constant and are equal to certain Gauss sums. For further details, see [Reference Chinta, Gunnells and Puskas15, Reference Patnaik and Puskas34].

Let $\overline {H}_0^m$ be the finite Hecke algebra over $\mathbf {K}$ viewed as subalgebra of $\mathbb {H}^m$ in the natural way. Consider the right $\overline {H}_0^m$ -action on $\mathbf {K}_{\mathcal {Q}^m}[\Lambda ]$ , defined by

$$ \begin{align*} (fx^\lambda)\blacktriangleleft^m T_v:=\pi_\Lambda^{m,X-\text{loc}}(T_{v^{-1}})(fx^\lambda) \end{align*} $$

for $v\in W_0$ , $f\in \mathcal {Q}^m$ and $\lambda \in \Lambda $ . This is the metaplectic analog of the $\overline {H}_0$ -action defined in Lemma 6.46. In fact, by Corollary 9.11(1), we have

(10.1) $$ \begin{align} \Gamma^m\big(fx^\lambda\blacktriangleleft^mT_v\big)=\Gamma^m(fx^\lambda)\blacktriangleleft T_v\qquad\quad (v\in W_0), \end{align} $$

where $\blacktriangleleft $ now denotes the right $\overline {H}_0^m$ -action (6.40) on $\mathbf {K}_{\mathcal {Q}^m}[\Lambda ]$ .

Define $\gamma _\mu =\gamma _\mu (\underline {h})\in \mathbf {K}^\times $ for $\mu \in \Lambda $ by

$$\begin{align*}\gamma_\mu:=\prod_{\alpha\in\Pi(v_\mu^{-1})}\left(-k_{\alpha^m}^{1-\chi_{\mathbb{Z}}(\alpha(c_\mu)/m(\alpha))}h_{\mathbf{Q}(\alpha^{\vee})\alpha(c_\mu)}(\alpha)\right), \end{align*}$$

where $c_\lambda \in \overline {C}_+^m$ and $v_\lambda \in W_0$ (see Definition 2.3) are defined relative to the metaplectic root system $\Phi _0^m$ . Then $\Gamma ^m(x^\mu )=\gamma _\mu x^\mu $ for all $\mu \in \Lambda $ ; see Lemma 9.10. We furthermore write

$$\begin{align*}\kappa_v^m(y):=\prod_{\alpha\in\Phi_0^m}k_{\alpha^m}^{-\eta(\alpha^m(y))}\qquad\qquad (v\in W_0,\, y\in E); \end{align*}$$

cf. (4.10).

Lemma 10.2. For $\lambda \in \Lambda $ and $\underline {h}\in \overline {\mathcal {M}}_{(n,\mathbf {Q})}$ , the Whittaker limit $\overline {E}_\lambda ^m(x)$ of the metaplectic polynomial $E_\lambda ^m(x)\in \mathcal {P}_\Lambda $ is well defined. Then

(10.2) $$ \begin{align} \Gamma^m(\overline{E}_\lambda^m(x))=\gamma_\lambda\,\overline{E}_\lambda^{\mathbf{J}(c_\lambda)}(x)\qquad\qquad \forall\,\lambda\in\Lambda \end{align} $$

and

(10.3) $$ \begin{align} \begin{aligned} \overline{E}_\lambda^m(x)&=x^\lambda\qquad\qquad\qquad\qquad\qquad\quad\, (\lambda\in\Lambda\cap\overline{E}_-),\\ \overline{E}_{v\lambda}^m(x)&= \frac{\gamma_{v\lambda}}{\kappa_v^m(\lambda)\gamma_\lambda}\, x^\lambda\blacktriangleleft^m T_v^{-1}\qquad\quad\,\,\, (v\in W_0,\,\lambda\in\Lambda\cap E_-). \end{aligned} \end{align} $$

Proof. By Theorem 9.18, we have

(10.4) $$ \begin{align} \Gamma^m(E_\lambda^m(x))=\gamma_\lambda\,E_\lambda^{\mathbf{J}(c_\lambda)}\big(x;q^{c_\lambda}\mathfrak{t}_{c_\lambda}(\underline{h})\vert_{Q^{m\vee}}\big) \end{align} $$

for $\lambda \in \Lambda $ . Since $\gamma _\mu $ does not depend on $q^{\frac {1}{2h}}$ , it follows from (10.4) and Proposition 6.47 that the Whittaker limit $\overline {E}_\lambda ^m(x)$ of $E_\lambda ^m(x)$ exists, and that (10.2) holds true.

By (6.49), we have $\overline {E}_\lambda ^m(x)=x^\lambda $ for $\lambda \in \Lambda \cap \overline {E}_-$ . For $\lambda \in \Lambda \cap E_-$ and $v\in W_0$ , we have by (6.52) and (10.1),

$$ \begin{align*} \begin{aligned} \Gamma^m\big(\overline{E}_{v\lambda}^m(x)\big)&=\frac{\gamma_{v\lambda}}{\kappa_v^m(\lambda)}\,x^\lambda\blacktriangleleft T_v^{-1}\\ &=\frac{\gamma_{v\lambda}}{\kappa_v^m(\lambda)\gamma_\lambda}\, \Gamma^m(x^\lambda)\blacktriangleleft T_v^{-1}\\ &=\frac{\gamma_{v\lambda}}{\kappa_v^m(\lambda)\gamma_\lambda}\,\Gamma^m\big(x^\lambda\blacktriangleleft^mT_v^{-1}\big), \end{aligned} \end{align*} $$

which concludes the proof of the lemma.

For $\lambda \in \Lambda $ , define the antisymmetric variant of the metaplectic polynomial $E_\lambda ^m(x)$ by

$$\begin{align*}E_\lambda^{m,-}(x):=\pi_\Lambda^m(\mathbf{1}_-)E_\lambda^m(x), \end{align*}$$

with $\mathbf {1}_-\in \overline {H}_0^m$ given by (6.33).

Corollary 10.3. For $\underline {h}\in \overline {\mathcal {M}}_{(n,\mathbf {Q})}$ and $\lambda \in \Lambda $ , the Whittaker limit $\overline {E}_\lambda ^{m,-}(x)$ of $E_\lambda ^{m,-}(x)$ is well defined. Furthermore,

(10.5) $$ \begin{align} \Gamma^m\big(\overline{E}_\lambda^{m,-}(x)\big)=\gamma_\lambda\,\overline{E}_\lambda^{\mathbf{J}(c_\lambda),-}(x)\qquad\forall\,\lambda\in\Lambda, \end{align} $$

and for $\lambda \in \Lambda \cap E_-$ ,

(10.6) $$ \begin{align} \overline{E}_\lambda^{m,-}(x)=\kappa_{w_0}^m(0)^{-2}\sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v^m(0) \frac{\kappa_v^m(\lambda)\gamma_\lambda}{\gamma_{v\lambda}}\,\overline{E}_{v\lambda}^{m}(x). \end{align} $$

Proof. By Corollary 9.11(1), we have

$$ \begin{align*} \begin{aligned} \Gamma^m\big(E_\lambda^{m,-}(x)\big)&=\gamma_\lambda\,\pi_{c_\lambda,q^{c_\lambda}\mathfrak{t}_{c_\lambda}(\underline{h})}(\mathbf{1}_-) E_\lambda^{\mathbf{J}(c_\lambda)}\big(x;q^{c_\lambda}\mathfrak{t}_\lambda(\underline{h})\vert_{Q^{m\vee}}\big)\\ &=\gamma_\lambda\,E_\lambda^{\mathbf{J}(c_\lambda),-}\big(x;q^{c_\lambda}\mathfrak{t}_\lambda(\underline{h})\vert_{Q^{m\vee}}\big). \end{aligned} \end{align*} $$

Hence, the Whittaker limit $\overline {E}_\lambda ^{m,-}(x)$ of $E_\lambda ^{m,-}(x)$ exists, and (10.5) holds true. For $\lambda \in \Lambda \cap E_-$ , we have by (10.5), Corollary 6.52 and (10.2),

$$ \begin{align*} \begin{aligned} \Gamma^m\big(\overline{E}_\lambda^{m,-}(x)\big)&=\kappa_{w_0}^m(0)^{-2}\sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v^m(0)\kappa_v^m(\lambda) \gamma_\lambda\,\overline{E}_{v\lambda}^{\mathbf{J}(c_\lambda)}(x)\\ &=\kappa_{w_0}^m(0)^{-2}\sum_{v\in W_0}(-1)^{\ell(v)}\kappa_v^m(0) \frac{\kappa_v^m(\lambda)\gamma_\lambda}{\gamma_{v\lambda}}\,\overline{E}_{v\lambda}^{m}(x), \end{aligned} \end{align*} $$

which completes the proof.

To relate (10.3) to formula [Reference Patnaik and Puskas34, (6.48)] of the metaplectic Iwahori-Whittaker function, we need to compare the $W_0$ -action $\sigma _\Lambda ^{\text {CG}}$ and the metaplectic Demazure-Lusztig operator $\mathcal {T}_i^m$ with [Reference Patnaik and Puskas34, (4.7)] and [Reference Patnaik and Puskas34, (4.10)]. The operator $w_{\alpha _i}\ast $ in [Reference Patnaik and Puskas34, (4.7)] corresponds to $I\circ \sigma _\Lambda ^{\text {CG}}(s_i)\circ I$ , where I is the $\mathbf {F}$ -algebra automorphism of $\mathbf {F}_{\mathcal {Q}^m}[\Lambda ]$ mapping $x^\lambda $ to $x^{-\lambda }$ for all $\lambda \in \Lambda $ . In this identification, the parameters $(n(\alpha _i^{\vee }),v)$ in [Reference Patnaik and Puskas34, (4.10)] correspond to $(m(\alpha _i),\mathbf {k}^2)$ . Similarly, the operator $\mathbf {T}_{\alpha _i}$ in [Reference Patnaik and Puskas34, (4.10)] corresponds to

(10.7) $$ \begin{align} \widetilde{\mathcal{T}}_i^m:= I\circ \mathcal{T}_i^m\circ I, \end{align} $$

with $\mathcal {T}_i^m$ the metaplectic Demazure-Lusztig operator (9.27).

The $\widetilde {\mathcal {T}}_i^m$ ( $i\in [1,r]$ ) satisfy the braid relations by the last formula of (9.27) (this was originally observed in [Reference Chinta, Gunnells and Puskas15, Prop. 7]), and hence, we can define for $v\in W_0$ the linear operator

$$\begin{align*}\widetilde{\mathcal{T}}_v^m:=\widetilde{\mathcal{T}}_{i_1}^m\cdots\widetilde{\mathcal{T}}_{i_\ell}^m \end{align*}$$

acting on $\mathbf {K}_{\mathcal {Q}^m}[\Lambda ]$ , where $v=s_{i_1}\cdots s_{i_\ell }$ is a reduced expression of $v\in W_0$ .

Definition 10.4. For $\lambda \in \Lambda \cap \overline {E}_+$ and $v\in W_0$ , set

$$ \begin{align*} \mathcal{W}_{v,\lambda}:=\widetilde{\mathcal{T}}_v^m\big(x^\lambda\big),\qquad\quad \mathcal{W}_\lambda:=\sum_{v\in W_0}\mathcal{W}_{v,\lambda} \end{align*} $$

in $\mathcal {P}_\Lambda $ .

The expressions [Reference Patnaik and Puskas34, Cor. 5.4] and [Reference Patnaik and Puskas34, (5.8)] of the metaplectic Iwahori-Whittaker function and the spherical Whittaker function in terms of metaplectic Demazure-Lusztig operators show that in the realm of representation theory of metaplectic covers of reductive groups (see Remark 10.1), the polynomial $\mathcal {W}_{v,\lambda }$ is a metaplectic Iwahori-Whittaker function and $\mathcal {W}_\lambda $ is a metaplectic spherical Whittaker function (the formula expressing the metaplectic spherical Whittaker function in terms of metaplectic Demazure-Lusztig operators was originally obtained in [Reference Chinta, Gunnells and Puskas15, Thm. 16]).

In the following theorem, we express $\mathcal {W}_{v,\lambda }$ and $\mathcal {W}_\lambda $ in terms of the Whittaker limits $\overline {E}_\lambda ^m(x)$ and $\overline {E}_\lambda ^{m,-}(x)$ of the metaplectic polynomials $E_\lambda ^m(x)$ and $E_\lambda ^{m,-}(x)$ .

Theorem 10.5. Let $\underline {h}\in \overline {\mathcal {M}}_{(n,\mathbf {Q})}$ and $\lambda \in \Lambda \cap \overline {E}_+$ .

  1. 1. For $v\in W_0$ , we have

    $$\begin{align*}\mathcal{W}_{v,\lambda}=(-1)^{\ell(v)}\kappa_v^m(0)\frac{\kappa_v^m(-\lambda-\rho^{\vee})\gamma_{-\lambda-\rho^{\vee}}}{\gamma_{-v(\lambda+\rho^{\vee})}}\,\, x^{-\rho^{\vee}} I\big(\overline{E}^{m}_{-v(\lambda+\rho^{\vee})}(x)\big). \end{align*}$$
  2. 2. We have

    (10.8) $$ \begin{align} \mathcal{W}_\lambda=\kappa_{w_0}^m(0)^2\,x^{-\rho^{\vee}}I\big(\overline{E}_{-\lambda-\rho^{\vee}}^{m,-}(x)\big). \end{align} $$

Proof. (1) Substituting (10.7) and the last equality of (9.27), we have

(10.9) $$ \begin{align} \begin{aligned} \mathcal{W}_{v,\lambda}=\widetilde{\mathcal{T}}_v^m\big(x^\lambda\big)&=(-1)^{\ell(v)}\kappa_v^m(0)\,x^{-\rho^{\vee}} I\left(\pi_\Lambda^{m,X-\text{loc}}\big(T_{v^{-1}}^{-1}\big)(x^{-\lambda-\rho^{\vee}})\right)\\ &=(-1)^{\ell(v)}\kappa_v^m(0)\,x^{-\rho^{\vee}}I\big(x^{-\lambda-\rho^{\vee}}\blacktriangleleft^m T_v^{-1}\big). \end{aligned} \end{align} $$

Note that $-\lambda -\rho ^{\vee }\in \Lambda \cap E_-$ , so the result follows from (10.3).

(2) This follows directly by combining part (1) and (10.6).

Remark 10.6. Formulas (10.2) and (10.5) now also allow to express $\mathcal {W}_{v,\lambda }$ and $\mathcal {W}_\lambda $ in terms of $\overline {E}_y^J(x)$ and $\overline {E}_y^{J,-}(x)$ .

Finally, we shortly explain how McNamara’s [Reference McNamara30, Thm. 15.2] Casselman-Shalika type formula for the metaplectic spherical Whittaker function can be derived from (6.56). First, note that $\Gamma ^m$ has a unique extension to a $\mathcal {Q}^m$ -linear automorphism of $\mathbf {K}_{\mathcal {Q}^m}[\Lambda ]$ . Define a linear $W_0$ -action on $\mathbf {K}_{\mathcal {Q}^m}[\Lambda ]$ by

$$\begin{align*}v\blacktriangleright^m (fx^\lambda):=\sigma_\Lambda^m(v)(fx^\lambda) \end{align*}$$

for $v\in W_0$ , $f\in \mathcal {Q}^m$ and $\lambda \in \Lambda $ . Then Corollary 9.16, (4.19) and Corollary 9.11(1) imply that

(10.10) $$ \begin{align} \Gamma^m\big(v\blacktriangleright^m (fx^\lambda)\big)=v\blacktriangleright \Gamma^m(fx^\lambda)\qquad\quad (v\in W_0), \end{align} $$

where $v\blacktriangleright (fx^\lambda )=\sigma _{c_\lambda ,q^{c_\lambda }\mathfrak {t}_\lambda (\underline {h})}(v)(fx^\lambda )$ (see (6.54)).

Proposition 10.7. Let $\lambda \in \Lambda \cap E_-$ . Then

(10.11) $$ \begin{align} \overline{E}_\lambda^{m,-}(x)=\left(\prod_{\alpha\in\Phi_0^+}\frac{1-k_{\alpha^m}^{-2}x^{-\alpha^{m\vee}}}{1-x^{-\alpha^{m\vee}}}\right) \sum_{v\in W_0}(-1)^{\ell(v)}v\blacktriangleright^m x^\lambda \end{align} $$

and

(10.12) $$ \begin{align} I(\mathcal{W}_\lambda)=\kappa_{w_0}^m(0)^2\left(\prod_{\alpha\in\Phi_0^+}\frac{1-k_{\alpha^m}^{-2}x^{-\alpha^{m\vee}}}{1-x^{-\alpha^{m\vee}}}\right) \sum_{v\in W_0}(-1)^{\ell(v)}\big(\prod_{\alpha\in\Pi(v^{-1})}x^{\alpha^{m\vee}}\big) \sigma_\Lambda^{\text{CG}}(v)\big(x^{-\lambda}\big). \end{align} $$

Proof. The proof of (10.11) follows by combining (10.5), (6.56) and (10.10). Substituting (10.11) into (10.8) and using that

$$\begin{align*}x^{\rho^{\vee}}\,\big(v\blacktriangleright^mx^{-\lambda-\rho^{\vee}}\big)=x^{\rho^{m\vee}}\sigma_\Lambda^{CG}(v)\big(x^{-\lambda-\rho^{m\vee}}\big) \end{align*}$$

by (9.25), we get

$$\begin{align*}I(\mathcal{W}_\lambda)=\kappa_{w_0}^m(0)^2\left(\prod_{\alpha\in\Phi_0^+} \frac{1-k_{\alpha^m}^{-2}x^{-\alpha^{m\vee}}}{1-x^{-\alpha^{m\vee}}}\right) \sum_{v\in W_0}(-1)^{\ell(v)}x^{\rho^{m\vee}} \sigma_\Lambda^{\text{CG}}(v)\big(x^{-\lambda-\rho^{m\vee}}\big). \end{align*}$$

The result now follows from the fact that

$$\begin{align*}\rho^{m\vee}-v\rho^{m\vee}=\sum_{\alpha\in\Pi(v^{-1})}\alpha^{m\vee}. \end{align*}$$

In the context of representation theory of metaplectic reductive groups over non-archimedean local fields (see Remark 10.1), formula (10.12) reduces to McNamara’s [Reference McNamara30, Thm. 15.2] Casselman-Shalika type formula for the metaplectic spherical Whittaker function; see also [Reference Sahi, Stokman and Venkateswaran38, Thm. 4.9].

11 Index of symbols

In the first column of the following table, we list spaces, groups and sets that will frequently occur in the paper. In the second column, we list the notations that we try to use throughout the paper for their elements.

§1.7 Conventions

$\mathbf {F},q,\text {Hom}(G,H),T_\Lambda ,R[A],\mathcal {P}_\Lambda ,x^\lambda ,t^\lambda ,[\ell ,m]$ .

§2.1 Root systems and Weyl groups

$E,\langle \cdot ,\cdot \rangle , \|\cdot \|, E_{\text {co}}, E^{\text {reg}}, E_{\pm },y_{\pm },E^\prime , E_{\pm }^\prime ,\text {pr}_{E^\prime },\text {pr}_{E_{\text {co}}},m, \Phi _0,\Phi _0^{\pm },\Delta _0=\{\alpha _i\}_{i=1}^r,\varphi , Q,$

$\Phi _0^{\vee },\Delta _0^{\vee }=\{\alpha _i^{\vee }\}_{i=1}^r,W_0,w_0, s_\alpha ,s_i,Q^{\vee },P^{\vee },P^{\vee ,\pm },\varpi _i^{\vee }$ .

§2.2 Affine root systems and affine Weyl groups

$E^{a,\text {reg}},C_{\pm }, \mathcal {O}_y,E^*\oplus \mathbb {R}\overset {D}{\twoheadrightarrow } E^*,\Phi ,\Phi ^{\pm },\Delta =\{\alpha _j\}_{j=0}^r,W,W_y, s_a,s_0,\tau (\mu ),W\overset {D}{\twoheadrightarrow } W_0$ .

§2.3 Length function and parabolic subgroups

$\ell (w),\Pi (w),c_y,w_y,\mu _y,v_y,\mathbf {J}(c),\Phi _J,\Phi _J^{\pm },W_J, W^J$ .

§2.4 The Coxeter complex of W

$\Sigma (\Phi ),C^J,C_J,\mathcal {F}_\Sigma (E,X)$ .

§2.5 The double affine Weyl group

$\Phi ^{\vee },a^{\vee },K,\widehat {Q}^{\vee },\mathbb {W},\mathcal {L},\Lambda ,\widehat {\Lambda },j_\Lambda ,\mathbf {q}$ .

§2.6 Algebras of q -difference reflection operators

$\mathbf {F},q_\alpha ,T_\Lambda ,T,t^\lambda ,\varpi ,z^\xi ,z^\mu ,\mathcal {P}_\Lambda ,\mathcal {P},\mathcal {Q}_\Lambda ,\mathcal {Q}, W\ltimes \mathcal {P},W\ltimes \mathcal {Q}$ .

§2.7 The double affine Hecke algebra

${}^{\text {sh}}\Phi _0,{}^{\text {lg}}\Phi _0,{}^{\text {sh}}k,{}^{\text {lg}}k,k,\mathbf {k},k_a,k_j,\chi _{\pm },\chi , \mathbb {H}=\mathbb {H}(\mathbf {k},q),H=H(\mathbf {k}),H_0=H_0(\mathbf {k})$ ,

$\mathcal {P}_Y,T_w,Y^\mu ,U_0,p(Y^{\pm 1}),\delta $ .

§2.8 Intertwiners

$S_j^X,S_w^X,S_j^Y,S_w^Y,\beth ,\mathbb {H}^{X-\text {loc}},\mathbb {H}^{Y-\text {loc}},\widetilde {S}_j^X,\widetilde {S}_w^X,\widetilde {S}_j^Y,\widetilde {S}_w^Y$ .

§3.1 Induction parameters

$T_{\Lambda ,J},T_{\Lambda ,J}^{\text {red}},T_J,T_J^{\text {red}},T_\Lambda ^G,\widetilde {T}_\Lambda , \widetilde {T}_{\Lambda ,J},L_{\Lambda ,J},L_J,L^J,1_{T_\Lambda },W_J^{\text {red}},J_0,\Phi _0^J,\Phi _{0,J_0},\Phi _0^{+,J}$ , $m(\varphi ),h,\eta _J(\alpha ),n_i(\alpha ),\mathfrak {s}_J$ .

§3.2 The $\mathbb {H}$ -module $\mathbb {M}_t^J$

$\mathbb {H}_J^X,\mathbb {H}_J^Y,\chi _{J,t}^X,\chi _{J,t},H_J,\mathbf {F}_{J,t},1_{J,t},\mathbb {M}_t^J,m_t^J,J(t),L^J,\mathbf {F}_{J,t}^X,{}^X\mathbb {M}_t^J,\pi ,t_{\text {sph}},\mathcal {M}(T),\mathcal {E},m_{w;t}^J,\mu _w,v_w$ .

§4.1 Spaces of quasi-polynomials

$R,R[E],x^y,x^{(y,m)},\mathcal {P}^{(c)},\mathfrak {t}_y,w_{\mathfrak {t}},T_{\mathbb {R}}^{\vee },\mathcal {P}_\gamma ,[c]$ .

§4.2 Truncated divided difference operators

$\nabla _a,\nabla _j$ .

§4.3 The quasi-polynomial realisation of $\mathbb {M}_t^J$

$\chi _B,\eta ,\kappa _v(y),\pi _{c,\mathfrak {t}},\mathcal {P}_{\mathfrak {t}}^{(c)},\phi _{c,\mathfrak {t}}$ .

§4.4 The face dependence of the quasi-polynomial representation

$e_{w,w^\prime ;\mathfrak {t}}^{J,h}, \text {pr}_{c,c^\prime }^{\mathfrak {t}^\prime }$ .

§4.5 Affine Weyl group actions on quasi-rational functions

$\mathbf {F}_{\mathcal {Q}}[E],\mathcal {Q}_\gamma ,\sigma _{c,\mathfrak {t}},\check {s}_j,\pi _{c,\mathfrak {t}}^{X-\text {loc}},\varphi ,\xi _y$ .

§5.1 Properties of the base point $\mathfrak {s}_J$

$\mathfrak {s}_y,k(y),k_w(y)$ .

§5.2 The $\mathbb {H}$ -action on $\mathcal {P}^{(c)}$

$\psi _{c,\mathfrak {t}},H_{i,m}$ .

§5.4 The parabolic Bruhat order on E

$\leq _B,(E,\leq ),\prec ,\prec _\alpha ,\text {sgn}$ .

§5.5 Triangularity properties

$G_{\mathfrak {t}}(a)$ .

§6.1 Genericity conditions for the multiplicity function

$T_J^\prime ,\lambda _J,q^y,P,\varpi _i,I^{\text {co}},\mathfrak {t}(\mathbf {z})$ , $W_0^I$ .

§6.2 The monic Y -eigenbasis of $\mathcal {P}_{\mathfrak {t}}^{(c)}$

$m_y^J(x;\mathfrak {t}),d_{y,y^\prime ;\mathfrak {t}}^J,d_{w,w^\prime ;\mathfrak {t}}^J,\mathcal {S}(M),M[s],E_y^J(x;\mathfrak {t}),e_{y,y^\prime ;\mathfrak {t}}^J,e_{w,w^\prime ;\mathfrak {t}}^J$ .

§6.3 Creation operators and irreducibility conditions

$d_w(x),\epsilon _{w,u},\epsilon _{w,u}^\prime ,\mathcal {P}^{(c)}_{<w},n_w$ .

§6.4 Closure relations

$e_{w,u^\prime ;\mathfrak {t}^\prime }^{J,J^\prime }$ .

§6.5 The normalised Y-eigenbasis and pseudo-duality

$r_w(x),P_y^J(x;\mathfrak {t}),t_y,\rho _w$ .

§6.6 (Anti)symmetrisation

$\mathbf {1}_{\pm },\mathcal {Q}_{\mathfrak {t}}^{(c),+},\mathcal {P}_{\mathfrak {t}}^{(c),+},E_y^{J,\pm }(x;\mathfrak {t}),P_y^{J,+}(x;\mathfrak {t}),\mathcal {O}_c^+,W_{0,y_{\pm }},W_0^{y_\pm },g_y,C_y^{\pm }(y^\prime )$ .

§6.7 Pseudo-unitarity and orthogonality relations

$\ast ,\mathbf {F}_u^\times ,\mathbf {F}_r,T_u,T_r,N_w(x), (\cdot ,\cdot )_{J,\mathfrak {t}},p^*,\text {ct},\mathcal {W}$ .

§6.8 The Whittaker limit

$\mathbf {K},\mathbf {F}_{\text {reg}},\overline {\mathcal {P}}^{(c)},\overline {\mathcal {P}},\mathcal {P}_{\text {reg}}^{(c)},\overline {\mathcal {Q}}^{(c)},\overline {\mathcal {Q}},\overline {d},\overline {f},\overline {H}_0,\overline {H},\overline {T},\overline {T}_J^{\text {red}},\overline {E}_y^J(x), \overline {E}_y^{J,\pm }(x),\blacktriangleleft ,\blacktriangleright , s_\mu $ .

§7.1 Extended root datum and extended affine Weyl groups

$e,(\mathcal {L}^{\times 2})_e,(\Lambda _1,\Lambda _1^\prime )\leq (\Lambda _2,\Lambda _2^\prime ),q^{\ell /e},q_\alpha ^{\ell /e},W_{\Lambda },\Omega _{\Lambda },w_{\lambda ,\Lambda },v_{\lambda ,\Lambda },\Lambda _{\text {min}},\zeta ^\prime ,\omega (j),I_{\Lambda ^\prime },\varpi _{i,\Lambda ^\prime }^{\vee }$ .

§7.2 The extended double affine Hecke algebra

$\mathbb {H}_{\Lambda ,\Lambda ^\prime },H_{\Lambda ^\prime },\Lambda ^{\prime +}$ .

§7.3 The extended quasi-polynomial representation

$\mathcal {P}_\Lambda ^{(c)},\pi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime },\mathcal {P}_{\Lambda ,\mathfrak {t}}^{(c)},\sigma _{c,\mathfrak {t}}^{Q^{\vee },\Lambda ^\prime },V_{c,\mathfrak {t}},V_{c,\mathfrak {t}}(\omega ),\phi _\omega ,W_{\Lambda ,c},\Omega _{\Lambda ,J},\Omega _{\Lambda ,c},\Omega _{\Lambda }^{c},{}^\Lambda T_{\Lambda ^\prime }^c,\mathfrak {t}_{y;c},w_{\mathfrak {t};c},\psi _{c,\mathfrak {t}}^{\Lambda ,\Lambda ^\prime }$ .

§7.4 Twist parameters

$\Xi _{\mathfrak {t}},\Theta _{\mathfrak {t}}$ .

§7.5 The extended eigenvalue equations

$E_{y;c}^{\Lambda ,\Lambda ^\prime }(x;\mathfrak {t})$ .

§7.6 The theory for the $\text {GL}_{r+1}$ root datum

$\epsilon _i,u,\varpi ,\varpi _{i,\ell \mathbb {Z}^{r+1}}^{\vee },\widetilde {\mathbb {H}},\widetilde {\pi }_{c,\mathfrak {t}},Y_i,\widetilde {E}_{y;c}(x;\mathfrak {t})$ .

§8.1 The g -parameters

$\widetilde {\mathfrak {t}}$ , $\kappa _\alpha (\widetilde {\mathfrak {t}}),\mathcal {G}^{\text {amb}},\mathbf {f},f_\alpha ,\widehat {\mathcal {G}},q^y,\widehat {\mathbf {g}},\widehat {g}_\alpha ,\mathcal {G},\mathbf {g},g_\alpha ,\mathcal {C}_{\Lambda ,\Lambda ^\prime },\mathfrak {c},\mathfrak {t}_y(\widehat {\mathbf {g}},\mathfrak {c}),w\cdot _{\widehat {\mathbf {p}},\mathfrak {c}}x^y$ .

§8.2 The uniform quasi-polynomial representation

$\Gamma _{\Lambda ,\mathbf {g}}^{(c)},\mathbf {F}_{\text {cl}},\widetilde {\mathbf {g}},\gamma _{\widetilde {\mathbf {g}}},\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ,\Lambda ^\prime }$ .

§8.3 Uniform quasi-polynomial eigenfunctions

$\mathcal {E}_y^{\Lambda ,\Lambda ^\prime }(x;\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}),\mathcal {E}_y(x;\mathbf {g},\widehat {\mathbf {p}}),\mathbf {F}_{\mathcal {Q}_\Lambda }[E],\sigma _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c}}^{\Lambda ,\Lambda ^\prime }$ .

§9 Metaplectic representations and metaplectic polynomials

$n,\mathbf {Q},\kappa _\ell $ .

§9.1 The metaplectic parameters

$\mathbf {B},m(\alpha ),\Phi _0^m,\alpha ^m,Q^{m\vee },\Phi ^m,\vartheta ,\theta ,s_j^m,C_+^m,C^{mJ},\mathcal {G}^m,\mathcal {M}=\mathcal {M}_{(n,\mathbf {Q})}$ , $\underline {h},h_s(\alpha ),\underline {h}^{\mathbf {g}}, h_s^{\mathbf {g}}(\alpha ),\ell _\alpha $ .

§9.2 The metaplectic basic representation

$\mathbb {H}^m,\mathcal {L}^m,\Lambda ^m,\Lambda ^{m\prime },r_\ell (s),\nabla _j^m,\pi _\Lambda ^{m},\mathbb {H}_{\Lambda ^m,\Lambda ^{m\prime }},\widehat {\mathcal {G}}^m,\pi _{\mathbf {g},\widehat {\mathbf {p}},\mathfrak {c};\Lambda }, \pi _{\mathbf {g};\Lambda }, \mathcal {P}_\Lambda ^{m},X^m(\Lambda )$ ,

$\Gamma ^m=\Gamma ^m_{\underline {h}},\mathfrak {t}_\lambda (\underline {h}),X^m(\Lambda ,\Lambda ^m), \kappa ,\widetilde {\mathbb {H}},Z_{\Lambda ,\ell }^m,\overline {0},Z_{\Lambda ,\ell }^{m,\text {inv}}, Z_{\Lambda ,\ell }^{m,\text {reg}},\widetilde {Z}_{\Lambda ,\ell }^{m,\text {reg}},\check {h}$ .

§9.3 A metaplectic affine Weyl group action on rational functions

$\mathcal {Q}^m,\mathbf {F}_{\mathcal {Q}^m}[L],\sigma _\Lambda ^m,\rho ^{\vee },\rho ^{m\vee },\sigma _\Lambda ^{\text {CG}},\mathcal {T}_i^m$ .

§9.4 The metaplectic polynomials

$E_\lambda ^m,\sigma _s^{(\underline {h})}(\alpha )$ .

§10 Limits to metaplectic Iwahori-Whittaker functions

$\overline {\mathcal {M}}_{(n,\mathbf {Q})},\overline {H}_0^m,\blacktriangleleft ^m,\gamma _\mu =\gamma _\mu (\underline {h}),\kappa _v^m(y), \overline {E}_\lambda ^m(x),E_\lambda ^{m,-}(x),\overline {E}_\lambda ^{m,-},\widetilde {\mathcal {T}}_i^m,\widetilde {\mathcal {T}}_v^m,\mathcal {W}_{v,\lambda }, \mathcal {W}_\lambda ,\blacktriangleright ^m$ .

Acknowledgements

JS thanks Ivan Cherednik, Oleg Chalykh and Valentin Buciumas, and VV thanks Alexei Borodin and Eric Rains, for interesting discussions. We thank the referees for their valuable comments and pointing out several typos.

Competing interest

The authors have no competing interests to declare.

Financial support

The research of SS was partially supported by NSF grants DMS-1939600 and 2001537, and Simons Foundation grants 509766 and 00006698.

References

Berest, Y., Etingof, P. and Ginzburg, V., ‘Cherednik operators and differential operators on quasi-invariants’, Duke Math. J. 118 (2003), 279337.CrossRefGoogle Scholar
Borodin, A. and Wheeler, M., ‘Nonsymmetric Macdonald polynomials via integrable vertex models’, Trans. Amer. Math. Soc. 375 (2022), 83538397.CrossRefGoogle Scholar
Bourbaki, N., Lie Groups and Lie Algebras. Chapters 4-6 (Springer, 2002).CrossRefGoogle Scholar
Chalykh, O., ‘Macdonald polynomials and algebraic integrability’, Adv. Math. 166, (2002), 193259.CrossRefGoogle Scholar
Chalykh, O. and Etingof, P., ‘Orthogonality relations and Cherednik identities for multivariable Baker-Akhiezer functions’, Adv. Math. 238 (2013), 246289.Google Scholar
Cherednik, I., ‘Nonsymmetric Macdonald polynomials’, Int. Math. Res. Not. IMRN 1995(10), 483515.CrossRefGoogle Scholar
Cherednik, I., ‘Double affine Hecke algebras and Macdonald’s conjectures’, Ann. Math. 141 (1997), 191216.CrossRefGoogle Scholar
Cherednik, I., Double Affine Hecke Algebras (London Math. Soc. Lecture Note Series) vol. 319 (Cambridge, Cambridge Univ. Press, 2005).CrossRefGoogle Scholar
Cherednik, I., ‘Intertwining operators of double affine Hecke algebras’, Selecta Math. (N.S.) 3 (1997), 459495.Google Scholar
Cherednik, I., ‘Difference Macdonald-Mehta conjecture’, Int. Math. Res. Not. IMRN 1997(10), 449467.CrossRefGoogle Scholar
Cherednik, I., Ma, X., ‘Spherical and Whittaker functions via DAHA I’, Selecta Math. (N.S.) 19 (2013), 737817.CrossRefGoogle Scholar
Chinta, G. and Offen, O., ‘A metaplectic Casselman-Shalika formula for $\textrm{GL}_r$ ’, Amer. J. Math. 135 (2013), 403441.CrossRefGoogle Scholar
Chinta, G. and Gunnells, P. E., ‘Weyl group multiple Dirichlet series constructed from quadratic characters’, Invent. Math. 167(2) (2007), 327353.CrossRefGoogle Scholar
Chinta, G. and Gunnells, P. E., ‘Constructing Weyl group multiple Dirichlet series’, J. Amer. Math. Soc. 23 (2010), 189215.CrossRefGoogle Scholar
Chinta, G., Gunnells, P. E. and Puskas, A., ‘Metaplectic Demazure operators and Whittaker functions’, Indiana Univ. Math. J. 66(3) (2017), 10451064.Google Scholar
Deodhar, V., ‘Some characterizations of Bruhat ordering on a Coxeter group and determination of the relative Möbius function’, Invent. Math. 39 (1977), 187198.CrossRefGoogle Scholar
Etingof, P. and Kirilliv, A. A. Jr., ‘Macdonald’s polynomials and representations of quantum groups’, Math. Res. Lett. 1 (1994), 279296.Google Scholar
Haglund, J., Haiman, M. and Loehr, N., ‘A combinatorial formula for non-symmetric Macdonald polynomials’, Amer. J. Math. 130(2) (2008), 359383.CrossRefGoogle Scholar
Haiman, M., Cherednik Algebras, Macdonald Polynomials and Combinatorics (International Congress of Mathematicians) Vol. III (Eur. Math. Soc., Zürich, 2006), 843872.Google Scholar
Humphreys, J. E., Introduction to Lie Algebras and Representation Theory (Graduate Texts in Math.) vol. 9 (Springer, 1972).Google Scholar
Humphreys, J. E., Reflection Groups and Coxeter Groups (Cambridge Studies in Advanced Mathematics) vol. 29 (Cambridge, Cambridge Univ. Press, 1990).Google Scholar
Ion, B., ‘Nonsymmetric Macdonald polynomials and matrix coefficients for unramified principal series’, Adv. Math. 201 (2006), 3662.CrossRefGoogle Scholar
Lee, K.-H. and Zhang, Y., ‘Weyl group multiple Dirichlet series for symmetrizable Kac-Moody root systems’, Trans. Amer. Math. Soc. 367 (2015), 597625.CrossRefGoogle Scholar
Letzter, G., ‘Quantum zonal spherical functions and Macdonald polynomials’, Adv. Math. 189 (2004), 88147.CrossRefGoogle Scholar
Lusztig, G., ‘Affine Hecke algebras and their graded version’, J. Amer. Math. Soc. 2 (1989), 599635.CrossRefGoogle Scholar
Macdonald, I. G., Spherical Functions on a Group of $p$ -adic Type (Publications of the Ramanujan Institute) no. 2 (University of Madras, Centre for Advanced Study in Mathematics, Ramanujan Institute, Madras, 1971).Google Scholar
Macdonald, I. G., ‘Orthogonal polynomials associated with root systems’, Sém. Lotharingien de Comb. 45 (2000), Art. B45a.Google Scholar
Macdonald, I. G., Affine Hecke Algebras and Orthogonal Polynomials Séminaire Bourbaki, Vol. 1994/95 (Astérisque) no. 237 (1996), Exp. No. 797, 4, 189207.Google Scholar
Macdonald, I. G., Affine Hecke Algebras and Orthogonal Polynomials (Cambridge Tracts in Math.) vol. 157 (Cambridge, Cambridge University Press, 2003).CrossRefGoogle Scholar
McNamara, P. J., ‘The metaplectic Casselman-Shalika formula’, Trans. Amer. Math. Soc. 368 (2016), 29132937.Google Scholar
Mukhin, E. and Varchenko, A., ‘Quasi-polynomials and the Bethe Ansatz’, Geometry & Topology Monographs 13 (2008), 385420.Google Scholar
Mukhin, E., Tarasov, V. and Varchenko, A., ‘Spaces of quasi-exponentials and representations of the Yangian $\mathrm{Y}\left({\mathfrak{gl}}_N\right)$ ’, Transform. Groups 19 (2014), 861885.Google Scholar
Noumi, M., ‘Macdonald’s symmetric polynomials as zonal spherical functions on some quantum homogeneous spaces’, Adv. Math. 123 (1996), 1677.CrossRefGoogle Scholar
Patnaik, M. M. and Puskas, A., ‘On Iwahori-Whittaker functions for metaplectic groups’, Adv. Math. 313 (2017), 875914.CrossRefGoogle Scholar
Patnaik, M. M. and Puskas, A., ‘Metaplectic covers of Kac-Moody groups and Whittaker functions’, Duke Math. J. 168 (2019), 553653.CrossRefGoogle Scholar
Ram, A. and Yip, M., ‘A combinatorial formula for Macdonald polynomials’, Adv. Math. 226 (2011), 309331.CrossRefGoogle Scholar
Sahi, S., ‘Nonsymmetric Koornwinder polynomials and duality’, Ann. of Math. (2) 150 (1999), 267282.CrossRefGoogle Scholar
Sahi, S., Stokman, J. V. and Venkateswaran, V., ‘Metaplectic representations of Hecke algebras, Weyl group actions, and associated polynomials’, Selecta Math. (N.S.) 27 (2021), Paper No. 47.Google Scholar
Saied, J., ‘A combinatorial formula for Sahi, Stokman, and Venkateswaran’s generalization of Macdonald polynomials’, Adv. Math. 404 (2022), Paper No. 108440, 51pp.Google Scholar
Steinberg, R., ‘Finite reflection groups’, Trans. Amer. Math. Soc. 91 (1959), 493504.Google Scholar
Stokman, J. V., ‘Quantum affine Knizhnik-Zamolodchikov equations and quantum spherical functions, I’, Int. Math. Res. Not. IMRN (2011) 5, 10231090.Google Scholar
Stokman, J. V., ‘Connection coefficients for basic Harish-Chandra series’, Adv. Math. 250 (2014), 351386.Google Scholar
Stokman, J. V., ‘The $c$ -function expansion of a basic hypergeometric function associated to root systems’, Ann. of Math. 179 (2014), 253299.CrossRefGoogle Scholar
Stokman, J. V., Macdonald-Koornwinder Polynomials , in Koornwinder, T. H. and Stokman, J. V. (eds.), Multivariable Special Functions (Encyclopedia of Special Functions, The Askey-Bateman Project) vol. 2 (Cambridge, Cambridge Univ. Press, 2020), 258313.CrossRefGoogle Scholar
Vasserot, E., ‘Induced and simple modules of double affine Hecke algebras’, Duke Math. J. 126 (2005), 251323.CrossRefGoogle Scholar