1 Introduction

Throughout this paper, \(\mathbb {K}\) denotes an algebraically closed field. We begin by recalling the notion of a tridiagonal pair. We will use the following terms. Let V  denote a vector space over \(\mathbb {K}\) with finite positive dimension. For a linear transformation A : V → V  and a subspace W ⊆ V , we say that W is an eigenspace of A whenever W ≠ 0 and there exists \(\theta \in \mathbb {K}\) such that W = {v ∈ V |Av = θv}. In this case, θ is called the eigenvalue of A associated with W. We say that A is diagonalizable whenever V  is spanned by the eigenspaces of A.

Definition 1 ([9, Definition 1.1])

Let V  denote a vector space over \(\mathbb {K}\) with finite positive dimension. By a tridiagonal pair (or TD pair) on V  we mean an ordered pair of linear transformations A : V → V  and A  : V → V  that satisfy the following four conditions.

  1. (i)

    Each of A, A is diagonalizable.

  2. (ii)

    There exists an ordering \(\{V_i\}_{i=0}^d\) of the eigenspaces of A such that

    $$\displaystyle \begin{aligned} A^{\ast} V_i \subseteq V_{i-1} + V_i+ V_{i+1} \qquad \qquad (0 \leq i \leq d), \end{aligned} $$
    (1)

    where V −1 = 0 and V d+1 = 0.

  3. (iii)

    There exists an ordering \(\{V^{\ast }_i\}_{i=0}^{\delta }\) of the eigenspaces of A such that

    $$\displaystyle \begin{aligned} A V^{\ast}_i \subseteq V^{\ast}_{i-1} + V^{\ast}_i+ V^{\ast}_{i+1} \qquad \qquad (0 \leq i \leq \delta), \end{aligned} $$
    (2)

    where \(V^{\ast }_{-1} = 0\) and \(V^{\ast }_{\delta +1}= 0\).

  4. (iv)

    There does not exist a subspace W of V  such that AW ⊆ W, A W ⊆ W, W ≠ 0, W ≠ V .

We say the pair A, A is over \(\mathbb {K}\).

Note 1

According to a common notational convention A denotes the conjugate-transpose of A. We are not using this convention. In a TD pair A, A the linear transformations A and A are arbitrary subject to (i)–(iv) above.

Referring to the TD pair in Definition 1, by [9, Lemma 4.5] the scalars d and δ are equal. We call this common value the diameter of A, A . To avoid trivialities, throughout this paper we assume that the diameter is at least one.

TD pairs first arose in the study of Q-polynomial distance-regular graphs and provided a way to study the irreducible modules of the Terwilliger algebra associated with such a graph. Since their introduction, TD pairs have been found to appear naturally in a variety of other contexts including representation theory [1, 7, 10,11,12, 14, 15, 25], orthogonal polynomials [23, 24], partially ordered sets [22], statistical mechanical models [3, 6, 19], and other areas of physics [16, 18]. As a result, TD pairs have become an area of interest in their own right. Among the above papers on representation theory, there are several works that connect TD pairs to quantum groups [1, 5, 7, 11, 12]. These papers consider certain special classes of TD pairs. We call particular attention to [5], in which the present author describes a new relationship between TD pairs in the q-Racah class and quantum groups. The present paper builds off of this work.

In the present paper, we give a new relationship between the maps Δ, ψ : V → V  introduced in [4], as well as describe a new decomposition of the underlying vector space that, in some sense, lies between the first and second split decompositions associated with a TD pair. In order to motivate our results, we now recall some basic facts concerning TD pairs. For the rest of this section, let A, A denote a TD pair on V , as in Definition 1. Fix an ordering \(\{V_i\}_{i=0}^d\) (resp. \(\{V_i^{\ast }\}_{i=0}^d\)) of the eigenspaces of A (resp. A ) which satisfies (1) (resp. (2)). For 0 ≤ i ≤ d let θ i (resp. \(\theta _i^{\ast }\)) denote the eigenvalue of A (resp. A ) corresponding to V i (resp. \(V_i^{\ast }\)). By [9, Theorem 11.1] the ratios

$$\displaystyle \begin{aligned} \frac{\theta_{i-2}-\theta_{i+1}}{\theta_{i-1}-\theta_i},\qquad \qquad \frac{\theta_{i-2}^{\ast}-\theta_{i+1}^{\ast}}{\theta_{i-1}^{\ast}-\theta_i^{\ast}} \end{aligned}$$

are equal and independent of i for 2 ≤ i ≤ d − 1. This gives two recurrence relations, whose solutions can be written in closed form. There are several cases [9, Theorem 11.2]. The most general case is called the q-Racah case [12, Section 1]. We will discuss this case shortly.

We now recall the split decompositions of V  [9]. For 0 ≤ i ≤ d define

$$\displaystyle \begin{aligned} \begin{array}{rcl} U_i&\displaystyle = (V^{\ast}_0+V^{\ast}_1+\cdots + V^{\ast}_i)\cap (V_i+V_{i+1}+\cdots + V_d),\\ U_i^{\Downarrow} &\displaystyle = (V^{\ast}_0+V^{\ast}_1+\cdots + V^{\ast}_i)\cap (V_0+V_1+\cdots + V_{d-i}). \end{array} \end{aligned} $$

By [9, Theorem 4.6], both the sums \(V=\sum _{i=0}^d U_i\) and \(V=\sum _{i=0}^d U_i^{\Downarrow }\) are direct. We call \(\{U_i\}_{i=0}^d\) (resp. \(\{U_i^{\Downarrow }\}_{i=0}^d\)) the first split decomposition (resp. second split decomposition) of V . In [9], the authors showed that A, A act on the first and second split decomposition in a particularly attractive way. This will be described in more detail in Sect. 3.

We now describe the q-Racah case. We say that the TD pair A, A has q-Racah type whenever there exist nonzero scalars \(q,a,b\in \mathbb {K}\) such that q 4 ≠ 1 and

$$\displaystyle \begin{aligned} \theta_i=aq^{d-2i}+a^{-1}q^{2i-d},\qquad \qquad \theta_i^{\ast}=bq^{d-2i}+b^{-1}q^{2i-d} \end{aligned}$$

for 0 ≤ i ≤ d. For the rest of this section assume that A, A has q-Racah type.

We recall the maps K and B [13, Section 1.1]. Let K : V → V  denote the linear transformation such that for 0 ≤ i ≤ d, U i is an eigenspace of K with eigenvalue q d−2i. Let B : V → V  denote the linear transformation such that for 0 ≤ i ≤ d, \(U_i^{\Downarrow }\) is an eigenspace of B with eigenvalue q d−2i. The relationship between K and B is discussed in considerable detail in [5].

We now bring in the linear transformation Ψ : V → V  [4, Lemma 11.1]. As in [5], we work with the normalization ψ = (q − q −1)(q d − q d) Ψ. A key feature of ψ is that by [4, Lemma 11.2, Corollary 15.3],

$$\displaystyle \begin{aligned} \psi U_i\subseteq U_{i-1},\qquad \qquad \psi U_i^\Downarrow\subseteq U_{i-1}^\Downarrow \end{aligned}$$

for 1 ≤ i ≤ d and both ψU 0 = 0 and \(\psi U_0^\Downarrow =0\). In [5], it is shown how ψ is related to several maps, including the maps K, B, as well as the map Δ which we now recall. By [4, Lemma 9.5], there exists a unique linear transformation Δ : V → V  such that

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle \Delta U_i\subseteq U_i^{\Downarrow} &\displaystyle \ (0\leq i \leq d),\\ &\displaystyle (\Delta -I)U_i\subseteq U_0+U_1+\cdots +U_{i-1} &\displaystyle \ (0\leq i \leq d). \end{array} \end{aligned} $$

In [4, Theorem 17.1], the present author showed that both

$$\displaystyle \begin{aligned} \Delta = \sum_{i=0}^d \left(\prod_{j=1}^i \frac{aq^{j-1}-a^{-1}q^{1-j}}{q^j-q^{-j}}\right)\psi^i, \ \Delta^{-1} =\sum_{i=0}^d \left(\prod_{j=1}^i \frac{a^{-1}q^{j-1}-aq^{1-j}}{q^j-q^{-j}}\right)\psi^i. \end{aligned}$$

The primary goal of this paper is to provide factorizations of these power series in ψ and to investigate the consequences of these factorizations. We accomplish this goal using a linear transformation \(\mathcal {M}:V\to V\) given by

$$\displaystyle \begin{aligned} \mathcal{M} =\frac{aK-a^{-1}B}{a-a^{-1}}. \end{aligned}$$

By construction, \(\mathcal {M}^\Downarrow =\mathcal {M}\). One can quickly check that \(\mathcal {M}\) is invertible. We show that the map \(\mathcal {M}\) is equal to each of

$$\displaystyle \begin{aligned} (I-a^{-1}q\psi)^{-1}K, \quad K(I-a^{-1}q^{-1}\psi)^{-1}, \quad (I-aq\psi)^{-1}B, \quad B(I-aq^{-1}\psi)^{-1}. \end{aligned}$$

We give a number of different relations involving the maps \(\mathcal {M},K,B,\psi \), the most significant of which are the following:

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle K \exp_q\left(\frac{a^{-1}}{q-q^{-1}}\psi\right)=\exp_q\left(\frac{a^{-1}}{q-q^{-1}}\psi\right) \mathcal{M},\\ &\displaystyle B \exp_q\left(\frac{a}{q-q^{-1}}\psi\right)=\exp_q\left(\frac{a}{q-q^{-1}}\psi\right) \mathcal{M}. \end{array} \end{aligned} $$

Using these equations, we obtain our main result which is that both

$$\displaystyle \begin{aligned} \begin{array}{rcl} \Delta&\displaystyle =\exp_{q}\left(\frac{a}{q-q^{-1}}\psi\right)\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\psi\right),\\ \Delta^{-1}&\displaystyle =\exp_{q}\left(\frac{a^{-1}}{q-q^{-1}}\psi\right)\exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\psi\right). \end{array} \end{aligned} $$

Due to its important role in the factorization of Δ, we explore the map \(\mathcal {M}\) further. We show that \(\mathcal {M}\) is diagonalizable with eigenvalues q d, q d−2, q d−4, …, q d. For 0 ≤ i ≤ d, let W i denote the eigenspace of \(\mathcal {M} \) corresponding to the eigenvalue q d−2i. We show that for 0 ≤ i ≤ d,

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle U_i=\exp_q\left(\frac{a^{-1}}{q-q^{-1}}\psi\right) W_i,\qquad \qquad &\displaystyle U_i^\Downarrow=\exp_q\left(\frac{a}{q-q^{-1}}\psi\right) W_i,\\ &\displaystyle W_i=\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\psi\right) U_i, \qquad \qquad &\displaystyle W_i=\exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\psi\right) U_i^\Downarrow. \end{array} \end{aligned} $$

In light of this result, we interpret the decomposition \(\{W_i\}_{i=0}^d\) as a sort of halfway point between the first and second split decompositions. We explore this decomposition further and give the actions of ψ, K, B,  Δ, A, A on \(\{W_i\}_{i=0}^d\). We then give the actions of \(\mathcal {M}^{\pm 1}\) on \(\{U_i\}_{i=0}^d,\{U_i^\Downarrow \}_{i=0}^d, \{V_i\}_{i=0}^d, \{V_i^{\ast }\}_{i=0}^d\). We conclude the paper with a discussion of the special case when A, A is a Leonard pair.

The present paper is organized as follows. In Sect. 2 we discuss some preliminary facts concerning TD pairs and TD systems. In Sect. 3 we discuss the split decompositions of V  as well as the maps K and B. In Sect. 4 we discuss the map ψ. In Sect. 5 we recall the map Δ and give Δ as a power series in ψ. In Sect. 6 we introduce the map \(\mathcal {M}\) and describe its relationship with A, K, B, ψ. In Sect. 7 we express Δ as a product of two linear transformations; one is a q-exponential in ψ and the other is a q −1-exponential in ψ. In Sect. 8 we describe the eigenvalues and eigenspaces of \(\mathcal {M}\) and discuss how the eigenspace decomposition of \(\mathcal {M}\) is related to the first and second split decompositions. In Sect. 9 we discuss the actions of ψ, K, B,  Δ, A, A on the eigenspace decomposition of \(\mathcal {M}\). In Sect. 10 we describe the action of \(\mathcal {M} \) on the first and second split decompositions of V , as well as on the eigenspace decompositions of A, A . In Sect. 11 we consider the case when A, A is a Leonard pair.

2 Preliminaries

When working with a tridiagonal pair, it is useful to consider a closely related object called a tridiagonal system. In order to define this object, we first recall some facts from elementary linear algebra [9, Section 2].

We use the following conventions. When we discuss an algebra, we mean a unital associative algebra. When we discuss a subalgebra, we assume that it has the same unit as the parent algebra.

Let V  denote a vector space over \(\mathbb {K}\) with finite positive dimension. By a decomposition of V, we mean a sequence of nonzero subspaces whose direct sum is V . Let End(V ) denote the \(\mathbb {K}\)-algebra consisting of all linear transformations from V  to V . Let A denote a diagonalizable element in End(V ). Let \(\{V_i\}_{i=0}^d\) denote an ordering of the eigenspaces of A. For 0 ≤ i ≤ d let θ i be the eigenvalue of A corresponding to V i. Define E i ∈End(V ) by (E i − I)V i = 0 and E iV j = 0 if j ≠ i (0 ≤ j ≤ d). In other words, E i is the projection map from V  onto V i. We refer to E i as the primitive idempotent of A associated with θ i. By elementary linear algebra, (i) AE i = E iA = θ iE i (0 ≤ i ≤ d); (ii) E iE j = δ ijE i (0 ≤ i, j ≤ d); (iii) V i = E iV  (0 ≤ i ≤ d); (iv) \(I=\sum _{i=0}^d E_i\). Moreover

$$\displaystyle \begin{aligned} E_i = \prod_{\genfrac{}{}{0 pt}{}{0 \leq j \leq d}{j\neq i}} \frac{A-\theta_j I}{\theta_i-\theta_j}\qquad \qquad (0 \leq i \leq d).{} \end{aligned}$$

Let M denote the subalgebra of End(V ) generated by A. Note that each of \(\{ A^i\}_{i=0}^d\), \(\{E_i\}_{i=0}^d\) is a basis for the \(\mathbb {K}\)-vector space M.

Let A, A denote a TD pair on V . An ordering of the eigenspaces of A (resp. A ) is said to be standard whenever it satisfies (1) (resp. (2)). Let \(\{V_i\}_{i=0}^d\) denote a standard ordering of the eigenspaces of A. By [9, Lemma 2.4], the ordering \(\{V_{d-i}\}_{i=0}^d\) is standard and no further ordering of the eigenspaces of A is standard. A similar result holds for the eigenspaces of A . An ordering of the primitive idempotents of A (resp. A ) is said to be standard whenever the corresponding ordering of the eigenspaces of A (resp. A ) is standard.

Definition 2 ([17, Definition 2.1])

Let V  denote a vector space over \(\mathbb {K}\) with finite positive dimension. By a tridiagonal system (or TD system) on V, we mean a sequence

$$\displaystyle \begin{aligned} \Phi = (A; \{ E_i\}_{i=0}^d; A^{\ast};\{ E_i^{\ast}\}_{i=0}^d) \end{aligned}$$

that satisfies (i)–(iii) below.

  1. (i)

    A, A is a tridiagonal pair on V .

  2. (ii)

    \(\{ E_i\}_{i=0}^d\) is a standard ordering of the primitive idempotents of A.

  3. (iii)

    \(\{ E_i^{\ast }\}_{i=0}^d\) is a standard ordering of the primitive idempotents of A .

We call d the diameter of Φ, and say Φ is over \(\mathbb {K}\). For notational convenience, set E −1 = 0, E d+1 = 0, \( E^{\ast }_{-1}=0\), \(E^{\ast }_{d+1}=0\).

In Definition 2 we do not assume that the primitive idempotents \(\{ E_i\}_{i=0}^d,\{ E_i^{\ast }\}_{i=0}^d\) all have rank 1. A TD system for which each of these primitive idempotents has rank 1 is called a Leonard system [20]. The Leonard systems are classified up to isomorphism [20, Theorem 1.9].

For the rest of this paper, fix a TD system Φ on V  as in Definition 2. Our TD system Φ can be modified in a number of ways to get a new TD system [9, Section 3]. For example, the sequence

$$\displaystyle \begin{aligned} \Phi^{\Downarrow} = (A;\{ E_{d-i}\}_{i=0}^d;A^{\ast};\{ E_i^{\ast}\}_{i=0}^d) \end{aligned}$$

is a TD system on V . Following [9, Section 3], we call Φ the second inversion of Φ. When discussing Φ, we use the following notational convention. For any object f associated with Φ, let f denote the corresponding object associated with Φ.

Definition 3

For 0 ≤ i ≤ d let θ i (resp. \(\theta _i^{\ast }\)) denote the eigenvalue of A (resp. A ) associated with E i (resp. \(E_i^{\ast }\)). We refer to \(\{\theta _i\}_{i=0}^d\) (resp. \(\{\theta _i^{\ast }\}_{i=0}^d\)) as the eigenvalue sequence (resp. dual eigenvalue sequence) of Φ.

By construction \(\{\theta _i\}_{i=0}^d\) are mutually distinct and \(\{\theta _i^{\ast }\}_{i=0}^d\) are mutually distinct. By [9, Theorem 11.1], the scalars

$$\displaystyle \begin{aligned} \frac{\theta_{i-2}-\theta_{i+1}}{\theta_{i-1}-\theta_i},\qquad \qquad \frac{\theta_{i-2}^{\ast}-\theta_{i+1}^{\ast}}{\theta_{i-1}^{\ast}-\theta_i^{\ast}}{} \end{aligned}$$

are equal and independent of i for 2 ≤ i ≤ d − 1. For this restriction, the solutions have been found in closed form [9, Theorem 11.2]. The most general solution is called q-Racah [12, Section 1]. This solution is described as follows.

Definition 4

Let Φ denote a TD system on V  as in Definition 2. We say that Φ has q-Racah type whenever there exist nonzero scalars \(q,a,b\in \mathbb {K}\) such that such that q 4 ≠ 1 and

$$\displaystyle \begin{aligned} \theta_i=aq^{d-2i}+a^{-1}q^{2i-d},\qquad \qquad \theta_i^{\ast}=bq^{d-2i}+b^{-1}q^{2i-d} {} \end{aligned} $$
(3)

for 0 ≤ i ≤ d.

Note 2

Referring to Definition 4, the scalars q, a, b are not uniquely defined by Φ. If q, a, b is one solution, then their inverses give another solution.

For the rest of the paper, we make the following assumption.

Assumption 1

We assume that our TD system Φ has q-Racah type. We fix q, a, b as in Definition 4.

Lemma 1 ([5, Lemma 2.4])

With reference to Assumption 1 , the following hold.

  1. (i)

    Neither of a 2, b 2 is among q 2d−2, q 2d−4, …, q 2−2d.

  2. (ii)

    q 2i ≠ 1 for 1 ≤ i  d.

Proof

The result follows from the comment below Definition 3. □

3 The First and Second Split Decomposition of V

Recall the TD system Φ from Assumption 1. In this section we consider two decompositions of V  associated with Φ, called the first and second split decomposition.

For 0 ≤ i ≤ d define

$$\displaystyle \begin{aligned} U_i = (E^{\ast}_0V+E^{\ast}_1V+\cdots + E^{\ast}_iV)\cap (E_iV+E_{i+1}V+\cdots + E_dV). \end{aligned}$$

For notational convenience, define U −1 = 0 and U d+1 = 0. Note that for 0 ≤ i ≤ d,

$$\displaystyle \begin{aligned} U_i^{\Downarrow}=(E_0^{\ast}V+E_1^{\ast}V +\cdots + E_i^{\ast} V)\cap (E_0V+E_1V+\cdots + E_{d-i} V). \end{aligned}$$

By [9, Theorem 4.6], the sequence \(\{U_i\}_{i=0}^d\) (resp. \(\{U_i^\Downarrow \}_{i=0}^d\)) is a decomposition of V . Following [9], we refer to \(\{U_i\}_{i=0}^d\) (resp. \(\{U_i^\Downarrow \}_{i=0}^d\)) as the first split decomposition (resp. second split decomposition) of V  with respect to Φ. By [9, Corollary 5.7], for 0 ≤ i ≤ d the dimensions of E iV , \(E_i^{\ast }V\), U i, \(U_i^\Downarrow \) coincide; we denote the common dimension by ρ i. By [9, Theorem 4.6],

$$\displaystyle \begin{aligned} \begin{array}{rcl} E_iV+E_{i+1}V+\cdots+E_dV=U_{i}+U_{i+1}+\cdots+U_d,{} \end{array} \end{aligned} $$
(4)
$$\displaystyle \begin{aligned} \begin{array}{rcl} E_0V+E_1V+\cdots+E_iV=U_{d-i}^\Downarrow+U_{d-i+1}^\Downarrow+\cdots+U_d^\Downarrow,{} \end{array} \end{aligned} $$
(5)
$$\displaystyle \begin{aligned} \begin{array}{rcl} E_0^{\ast}V+E_1^{\ast}V+\cdots E_i^{\ast}V=U_0+U_1+\cdots +U_i=U_0^\Downarrow+U_1^\Downarrow+\cdots +U_i^\Downarrow. {} \end{array} \end{aligned} $$
(6)

By [9, Theorem 4.6], A and A act on the first split decomposition in the following way:

$$\displaystyle \begin{aligned} \begin{array}{rcl} (A-\theta_i I)U_i\subseteq U_{i+1} \qquad &\displaystyle (0\leq i \leq d-1 ), \qquad &\displaystyle (A-\theta_d I)U_d=0,\\ (A^{\ast}-\theta_i^{\ast} I)U_i\subseteq U_{i-1} \qquad &\displaystyle (1\leq i\leq d ), \qquad &\displaystyle (A^{\ast}-\theta_0^{\ast} I)U_0=0. \end{array} \end{aligned} $$

By [9, Theorem 4.6], A and A act on the second split decomposition in the following way:

$$\displaystyle \begin{aligned} \begin{array}{rcl} (A-\theta_{d-i} I)U_i^{\Downarrow}\subseteq U_{i+1}^{\Downarrow}\qquad &\displaystyle (0\leq i\leq d-1 ), \qquad &\displaystyle (A-\theta_0 I)U_d^{\Downarrow}=0,\\ (A^{\ast}-\theta_i^{\ast} I)U_i^{\Downarrow}\subseteq U_{i-1}^{\Downarrow}\qquad &\displaystyle (1\leq i\leq d ), \qquad &\displaystyle (A^{\ast}-\theta_0^{\ast} I)U_0^{\Downarrow}=0. \end{array} \end{aligned} $$

Definition 5 ([5, Definitions 3.1 and 3.2])

Define K, B ∈End(V ) such that for 0 ≤ i ≤ d, U i (resp. \(U_i^\Downarrow \)) is the eigenspace of K (resp. B) with eigenvalue q d−2i. In other words,

$$\displaystyle \begin{aligned} (K-q^{d-2i}I)U_i=0, \qquad \qquad (B-q^{d-2i}I)U_i^\Downarrow=0 \qquad \qquad (0\leq i\leq d).{} \end{aligned} $$
(7)

Observe that B = K .

By construction each of K, B is invertible and diagonalizable on V .

We now describe how K and B act on the eigenspaces of the other one.

Lemma 2 ([5, Lemma 3.3])

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \begin{array}{rcl} (B-q^{d-2i}I)U_i&\displaystyle \subseteq U_0+U_1+\cdots +U_{i-1},{} \end{array} \end{aligned} $$
(8)
$$\displaystyle \begin{aligned} \begin{array}{rcl} (K-q^{d-2i}I)U_i^\Downarrow &\displaystyle \subseteq U_0^\Downarrow+U_1^\Downarrow+\cdots +U_{i-1}^\Downarrow.{} \end{array} \end{aligned} $$
(9)

Next we describe how A, K, B are related.

Lemma 3 ([13, Section 1.1])

Both

$$\displaystyle \begin{aligned} \frac{qKA-q^{-1}AK}{q-q^{-1}}=aK^2+a^{-1}I,\qquad \qquad \frac{qBA-q^{-1}AB}{q-q^{-1}}=a^{-1}B^2+aI. {} \end{aligned} $$
(10)

Lemma 4 ([5, Theorem 9.9])

We have

$$\displaystyle \begin{aligned} aK^2-\frac{a^{-1}q-aq^{-1}}{q-q^{-1}}\ KB-\frac{aq-a^{-1}q^{-1}}{q-q^{-1}}\ BK+a^{-1}B^2=0.{} \end{aligned} $$
(11)

4 The Linear Transformation ψ

We continue to discuss the situation of Assumption 1. In [4, Section 11] we introduced an element Ψ ∈ End(V ). In [5] we used the normalization ψ = (q − q −1)(q d − q d) Ψ. In [5, Theorem 9.8], we showed that ψ is equal to some rational expressions involving K, B. We now recall this result. We start with a comment.

Lemma 5 ([5, Lemma 9.7])

Each of the following is invertible:

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle aI-a^{-1}BK^{-1}, \qquad \qquad &\displaystyle a^{-1}I-aKB^{-1},{} \end{array} \end{aligned} $$
(12)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle aI-a^{-1}K^{-1}B, \qquad \qquad &\displaystyle a^{-1}I-aB^{-1}K. \end{array} \end{aligned} $$
(13)

Lemma 6 ([5, Theorem 9.8])

The following four expressions coincide:

$$\displaystyle \begin{aligned} \begin{array}{rcl} \frac{I-BK^{-1}}{q(aI-a^{-1}BK^{-1})}, \qquad \qquad &\displaystyle &\displaystyle \frac{I-KB^{-1}}{q(a^{-1}I-aKB^{-1})},{} \end{array} \end{aligned} $$
(14)
$$\displaystyle \begin{aligned} \begin{array}{rcl} \frac{q(I-K^{-1}B)}{aI-a^{-1}K^{-1}B}, \qquad \qquad &\displaystyle &\displaystyle \frac{q(I-B^{-1}K)}{a^{-1}I-aB^{-1}K}.{} \end{array} \end{aligned} $$
(15)

In (14), (15) the denominators are invertible by Lemma 5.

Definition 6

Define ψ ∈End(V ) to be the common value of the four expressions in Lemma 6.

We now recall some facts concerning ψ.

Lemma 7 ([5, Lemma 5.4])

Both

$$\displaystyle \begin{aligned} K\psi =q^2\psi K, \qquad \qquad B\psi =q^2\psi B.{} \end{aligned} $$
(16)

Lemma 8 ([4, Lemma 11.2, Corollary 15.3])

We have

$$\displaystyle \begin{aligned} \psi U_i\subseteq U_{i-1},\qquad \qquad \psi U_i^\Downarrow\subseteq U_{i-1}^\Downarrow\qquad \qquad (1\leq i\leq d) \end{aligned} $$
(17)

and also ψU 0 = 0 and \(\psi U_0^\Downarrow =0\). Moreover ψ d+1 = 0.

In Lemma 6 we obtained ψ as a rational expression in BK −1 or K −1B. Next we solve for BK −1 and K −1B as a rational function in ψ. In order to state the answer, we will need the following result.

Lemma 9 ([5, Lemma 9.2])

Each of the following is invertible:

$$\displaystyle \begin{aligned} I-aq\psi,\qquad I-a^{-1}q\psi, \qquad I-aq^{-1}\psi,\qquad I-a^{-1}q^{-1}\psi.{} \end{aligned} $$
(18)

Their inverses are as follows:

$$\displaystyle \begin{aligned} \begin{array}{rcl} (I-aq\psi)^{-1}&\displaystyle =\sum_{i=0}^da^iq^i\psi^i,\qquad (I-a^{-1}q\psi)^{-1}&\displaystyle =\sum_{i=0}^d a^{-i}q^i\psi^i,{} \end{array} \end{aligned} $$
(19)
$$\displaystyle \begin{aligned} \begin{array}{rcl} (I-aq^{-1}\psi)^{-1}&\displaystyle =\sum_{i=0}^d a^iq^{-i}\psi^i,\qquad (I-a^{-1}q^{-1}\psi)^{-1}&\displaystyle =\sum_{i=0}^d a^{-i}q^{-i}\psi^i.{} \end{array} \end{aligned} $$
(20)

The next result is an immediate consequence of Lemma 6, Definition 6, and Lemma 9.

Theorem 1 ([5, Theorem 9.4])

The following hold:

$$\displaystyle \begin{aligned} \begin{array}{rcl} BK^{-1}&\displaystyle =\displaystyle\frac{I-aq\psi}{I-a^{-1}q\psi}, \qquad \qquad &\displaystyle KB^{-1}=\frac{I-a^{-1}q\psi}{I-aq\psi},{} \end{array} \end{aligned} $$
(21)
$$\displaystyle \begin{aligned} \begin{array}{rcl} K^{-1}B&\displaystyle =\displaystyle\frac{I-aq^{-1}\psi}{I-a^{-1}q^{-1}\psi}, \qquad \qquad &\displaystyle B^{-1}K=\frac{I-a^{-1}q^{-1}\psi}{I-aq^{-1}\psi}.{} \end{array} \end{aligned} $$
(22)

In (21), (22) the denominators are invertible by Lemma 9.

Lemma 10 ([5, Equation (22)])

We have

$$\displaystyle \begin{aligned} \frac{\psi A-A\psi}{q-q^{-1}}=\left(I-aq\psi\right)K-\left(I-a^{-1}q^{-1}\psi\right)K^{-1}. \end{aligned} $$
(23)

Proof

This result is a reformulation of [5, Equation (22)] using [5, Equation (14)]. □

5 The Linear Transformation Δ

We continue to discuss the situation of Assumption 1. In [4, Section 9] we introduced an invertible element Δ ∈ End(V ). In [4] we showed that Δ, ψ commute and in fact both Δ, Δ−1 are power series in ψ. These power series will be the central focus of this paper. We will show that each of those power series factors as a product of two power series, each of which is a quantum exponential in ψ.

Lemma 11 ([4, Lemma 9.5])

There exists a unique Δ ∈End(V ) such that

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle \Delta U_i\subseteq U_i^{\Downarrow}\qquad &\displaystyle \ (0\leq i \leq d),{} \end{array} \end{aligned} $$
(24)
$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle (\Delta -I)U_i\subseteq U_0+U_1+\cdots +U_{i-1} \qquad &\displaystyle \ (0\leq i \leq d).{} \end{array} \end{aligned} $$
(25)

Lemma 12 ([4, Lemmas 9.3 and 9.6])

The map Δ is invertible. Moreover Δ −1 = Δ and

$$\displaystyle \begin{aligned} (\Delta^{-1} -I)U_i\subseteq U_0+U_1+\cdots +U_{i-1} \qquad \qquad (0\leq i\leq d). {} \end{aligned} $$
(26)

Lemma 13

The map Δ  I is nilpotent. Moreover ΔK = B Δ.

Proof

The first assertion follows from (25). The last assertion follows from (24) and Definition 5. □

The map Δ is characterized as follows.

Lemma 14 ([4, Lemma 9.8])

The map Δ is the unique element of End(V ) such that

$$\displaystyle \begin{aligned} \begin{array}{rcl} (\Delta -I)E_i^{*} V\subseteq E_0^{*}V+E_1^{*}V+\cdots +E_{i-1}^{*}V \quad &\displaystyle \ (0\leq i \leq d),\\{} \end{array} \end{aligned} $$
(27)
$$\displaystyle \begin{aligned} \begin{array}{rcl} \Delta (E_iV+E_{i+1}V+\cdots+E_d V)= E_0 V+E_{1}V+\cdots+E_{d-i} V\quad &\displaystyle (0\leq i \leq d). \\{} \end{array} \end{aligned} $$
(28)

Theorem 2 ([4, Theorem 17.1])

Both

$$\displaystyle \begin{aligned} \begin{array}{rcl} \Delta &\displaystyle =&\displaystyle \sum_{i=0}^d \left(\prod_{j=1}^i \frac{aq^{j-1}-a^{-1}q^{1-j}}{q^j-q^{-j}}\right)\psi^i, {} \end{array} \end{aligned} $$
(29)
$$\displaystyle \begin{aligned} \begin{array}{rcl} \Delta^{-1} &\displaystyle =&\displaystyle \sum_{i=0}^d \left(\prod_{j=1}^i \frac{a^{-1}q^{j-1}-aq^{1-j}}{q^j-q^{-j}}\right)\psi^i. {} \end{array} \end{aligned} $$
(30)

In (29) and (30), the elements Δ, Δ−1 are expressed as a power series in ψ. In the present paper, we factor these power series and interpret the results. This interpretation will involve a linear transformation \(\mathcal {M}\). We introduce \(\mathcal {M}\) in the next section.

6 The Linear Transformation \(\mathcal {M}\)

We continue to discuss the situation of Assumption 1. In this section we introduce an element \(\mathcal {M} \in \mathrm {End}(V)\). We explain how \(\mathcal {M}\) is related to K, B, ψ, A.

Definition 7

Define \(\mathcal {M} \in \mathrm {End}(V)\) by

$$\displaystyle \begin{aligned} \mathcal{M} =\frac{aK-a^{-1}B}{a-a^{-1}}.{} \end{aligned} $$
(31)

By construction, \(\mathcal {M}^\Downarrow =\mathcal {M} \). Evaluating (31) using Lemma 5, we see that \(\mathcal {M}\) is invertible.

Lemma 15

The map \(\mathcal {M}\) is equal to each of:

$$\displaystyle \begin{aligned} (I-a^{-1}q\psi)^{-1}K, \qquad K(I-a^{-1}q^{-1}\psi)^{-1}, \qquad (I-aq\psi)^{-1}B, \qquad B(I-aq^{-1}\psi)^{-1}. \end{aligned}$$

Proof

We first show that \(\mathcal {M}=(I-a^{-1}q\psi )^{-1}K\). By Definition 7,

$$\displaystyle \begin{aligned} (a-a^{-1})\mathcal{M} K^{-1}=aI-a^{-1}BK^{-1}. \end{aligned}$$

The result follows from this fact along with the equation on the left in (21).

The remaining assertions follow from Theorem 1. □

Lemma 15 can be reformulated as follows.

Lemma 16

We have

$$\displaystyle \begin{aligned} \begin{array}{rcl} K&\displaystyle =\left(I-a^{-1}q\psi\right)\mathcal{M}, \qquad \qquad &\displaystyle K=\mathcal{M} \left(I-a^{-1}q^{-1}\psi\right), {} \end{array} \end{aligned} $$
(32)
$$\displaystyle \begin{aligned} \begin{array}{rcl} B&\displaystyle =\left(I-aq\psi\right)\mathcal{M},\qquad \qquad &\displaystyle B=\mathcal{M}\left(I-aq^{-1}\psi\right). {} \end{array} \end{aligned} $$
(33)

For later use, we give several descriptions of \(\mathcal {M}^{\pm 1}\).

Lemma 17

The map \(\mathcal {M}^{-1}\) is equal to each of:

$$\displaystyle \begin{aligned} K^{-1}(I-a^{-1}q\psi), \quad (I-a^{-1}q^{-1}\psi)K^{-1}, \quad B^{-1}(I-aq\psi), \quad (I-aq^{-1}\psi)B^{-1}. \end{aligned}$$

Proof

Immediate from Lemma 15. □

Lemma 18

The map \(\mathcal {M}\) is equal to each of:

$$\displaystyle \begin{aligned} K\sum_{n=0}^d a^{-n}q^{-n}\psi^n,\qquad \sum_{n=0}^d a^{-n}q^{n}\psi^nK,\qquad B\sum_{n=0}^d a^{n}q^{-n}\psi^n,\qquad \sum_{n=0}^d a^{n}q^{n}\psi^n B \end{aligned} $$
(34)

Proof

Use Lemmas 9 and 15. □

We now give some attractive equations that show how \(\mathcal {M}\) is related to ψ, K, B, A.

Lemma 19

We have

$$\displaystyle \begin{aligned} \mathcal{M} \psi=q^2\psi \mathcal{M} . \end{aligned} $$
(35)

Proof

Use Lemma 7 and Definition 7. □

Lemma 20

We have

$$\displaystyle \begin{aligned} \frac{q\mathcal{M}^{-1}K-q^{-1}K\mathcal{M}^{-1}}{q-q^{-1}}=I,\qquad \qquad \frac{q\mathcal{M}^{-1}B-q^{-1}B\mathcal{M}^{-1}}{q-q^{-1}}=I.{} \end{aligned} $$
(36)

Proof

Use Lemma 17. □

Lemma 21

We have

$$\displaystyle \begin{aligned} \frac{qA\mathcal{M}^{-1}-q^{-1}\mathcal{M}^{-1}A}{q-q^{-1}}=(a+a^{-1})I-(q+q^{-1})\psi.{} \end{aligned} $$
(37)

Proof

Use Lemmas 3, 7, 10, and 17. □

Lemma 22

We have

$$\displaystyle \begin{aligned} \mathcal{M}^{-2}A-(q^2+q^{-2})\mathcal{M}^{-1}A\mathcal{M}^{-1}+A\mathcal{M}^{-2}=-(q-q^{-1})^2(a+a^{-1})\mathcal{M}^{-1}. \end{aligned} $$
(38)

Proof

Use Lemmas 19 and 21. □

7 A Factorization of Δ

We continue to discuss the situation of Assumption 1. We now bring in the q-exponential function [8]. In [4, Theorem 17.1] we expressed Δ as a power series in ψ. In this section we strengthen this result in the following way. We express Δ as a product of two linear transformations; one is a q-exponential in ψ and the other is a q −1-exponential in ψ.

For an integer n, define

$$\displaystyle \begin{aligned}{}[n]_q=\frac{q^n-q^{-n}}{q-q^{-1}} \end{aligned} $$
(39)

and for n ≥ 0, define

$$\displaystyle \begin{aligned}{}[n]_q^!=[n]_q[n-1]_q\cdots [1]_q. \end{aligned} $$
(40)

We interpret \([0]_q^!=1\).

We now recall the q-exponential function [8]. For a nilpotent T ∈End(V ),

$$\displaystyle \begin{aligned} \exp_q(T)=\sum_{n=0}^\infty \frac{q^{\binom{n}{2}}}{[n]_q^!} T^n.{} \end{aligned} $$
(41)

The map expq(T) is invertible. Its inverse is given by

$$\displaystyle \begin{aligned} \exp_{q^{-1}}(-T)=\sum_{n=0}^\infty \frac{(-1)^n q^{-\binom{n}{2}}}{[n]_q^!} T^n.{} \end{aligned} $$
(42)

Using (41) we obtain

$$\displaystyle \begin{aligned} (I-(q^2-1)T)\exp_q(q^2T)=\exp_q(T).{} \end{aligned} $$
(43)

For S ∈End(V ) such that ST = q 2TS, we have

$$\displaystyle \begin{aligned} S\exp_q(T)S^{-1}=\exp_q(STS^{-1})=\exp_q(q^2T). \end{aligned}$$

Consequently

$$\displaystyle \begin{aligned} S\exp_q(T)=\exp_q(q^2T)S.{} \end{aligned} $$
(44)

Combining (43) and (44),

$$\displaystyle \begin{aligned} (I-(q^2-1)T)S\exp_q(T)=\exp_q(T)S.{} \end{aligned} $$
(45)

We return our attention to \(K,B,\psi ,\mathcal {M}\).

Proposition 1

Both

$$\displaystyle \begin{aligned} \begin{array}{rcl} K \exp_q\left(\frac{a^{-1}}{q-q^{-1}}\psi\right)&\displaystyle =&\displaystyle \exp_q\left(\frac{a^{-1}}{q-q^{-1}}\psi\right) \mathcal{M} ,{} \end{array} \end{aligned} $$
(46)
$$\displaystyle \begin{aligned} \begin{array}{rcl} B \exp_q\left(\frac{a}{q-q^{-1}}\psi\right)&\displaystyle =&\displaystyle \exp_q\left(\frac{a}{q-q^{-1}}\psi\right) \mathcal{M} .{} \end{array} \end{aligned} $$
(47)

Proof

Recall from Lemma 19 that \(\mathcal {M} \psi =q^2\psi \mathcal {M}\). We first obtain (46). To do this, in (45) take \(S=\mathcal {M}\) and \(T=\frac {a^{-1}}{q-q^{-1}}\psi \). Evaluate the result using the equation \(\mathcal {M}=(I-a^{-1}q\psi )^{-1}K\) from Lemma 15.

Next we obtain (47). To do this, in (45) take \(S=\mathcal {M}\) and \(T=\frac {a}{q-q^{-1}}\psi \). Evaluate the result using the equation \(\mathcal {M}=(I-aq\psi )^{-1}B\) from Lemma 15. □

The following is our main result.

Theorem 3

Both

$$\displaystyle \begin{aligned} \begin{array}{rcl} \Delta&\displaystyle =&\displaystyle \exp_{q}\left(\frac{a}{q-q^{-1}}\psi\right)\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\psi\right),{} \end{array} \end{aligned} $$
(48)
$$\displaystyle \begin{aligned} \begin{array}{rcl} \Delta^{-1}&\displaystyle =&\displaystyle \exp_{q}\left(\frac{a^{-1}}{q-q^{-1}}\psi\right)\exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\psi\right).{} \end{array} \end{aligned} $$
(49)

Proof

We first show (48). Let \(\tilde \Delta \) denote the expression on the right in (48). Combining (46) and (47), we see that \(\tilde \Delta K=B\tilde \Delta \). Therefore \(\tilde \Delta U_i=U_i^\Downarrow \) for 0 ≤ i ≤ d. Observe that \(\tilde \Delta -I\) is a polynomial in ψ with zero constant term. By Lemma 8, \((\tilde \Delta -I)U_i\subseteq U_0+U_1+\cdots +U_{i-1}\) for 0 ≤ i ≤ d. By Lemma 11, \(\tilde \Delta =\Delta \).

To obtain (49) from (48), use (42). □

Corollary 1

We have

$$\displaystyle \begin{aligned} \begin{array}{rcl} \exp_{q}\left(\frac{a}{q-q^{-1}}\psi\right)\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\psi\right)&\displaystyle =&\displaystyle \sum_{i=0}^d \left(\prod_{j=1}^i \frac{aq^{j-1}-a^{-1}q^{1-j}}{q^j-q^{-j}}\right)\psi^i,\\ \exp_{q}\left(\frac{a^{-1}}{q-q^{-1}}\psi\right)\exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\psi\right)&\displaystyle =&\displaystyle \sum_{i=0}^d \left(\prod_{j=1}^i \frac{a^{-1}q^{j-1}-aq^{1-j}}{q^j-q^{-j}}\right)\psi^i. \end{array} \end{aligned} $$

Proof

Combine Theorems 2 and 3. The equations can also be obtained directly by expanding their left-hand sides using (41) and (42), and evaluating the results using the q-binomial theorem [2, Theorem 10.2.1]. □

8 The Eigenvalues and Eigenspaces of \(\mathcal {M}\)

We continue to discuss the situation of Assumption 1. In Sect. 6 we introduced the linear transformation \(\mathcal {M}\). Proposition 1 indicates the role of \(\mathcal {M}\) in the factorization of Δ in Theorem 3. In this section we show that \(\mathcal {M}\) is diagonalizable. We describe the eigenvalues and eigenspaces of \(\mathcal {M}\). We also explain how the eigenspace decomposition for \(\mathcal {M}\) is related to the first and second split decompositions.

Lemma 23

The map \(\mathcal {M} \) is diagonalizable with eigenvalues q d, q d−2, q d−4, …, q d.

Proof

Let \(E=\exp _{q}\left (\frac {a^{-1}}{q-q^{-1}}\psi \right )\). By (46), \(\mathcal {M}=E^{-1}KE\). By construction K is diagonalizable with eigenvalues q d, q d−2, q d−4, …, q d. The result follows. □

Definition 8

For 0 ≤ i ≤ d, let W i denote the eigenspace of \(\mathcal {M} \) corresponding to the eigenvalue q d−2i. Note that \(\{W_i\}_{i=0}^d\) is a decomposition of V , and that \(W_i^\Downarrow =W_i\) for 0 ≤ i ≤ d. For notational convenience, let W −1 = 0 and W d+1 = 0.

Proposition 2

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \begin{array}{rcl} U_i&\displaystyle =\exp_q\left(\frac{a^{-1}}{q-q^{-1}}\psi\right) W_i, \qquad &\displaystyle U_i^\Downarrow=\exp_q\left(\frac{a}{q-q^{-1}}\psi\right) W_i, {} \end{array} \end{aligned} $$
(50)
$$\displaystyle \begin{aligned} \begin{array}{rcl} W_i&\displaystyle =\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\psi\right) U_i,\qquad &\displaystyle W_i=\exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\psi\right) U_i^\Downarrow.{} \end{array} \end{aligned} $$
(51)

Proof

Define E as in the proof of Lemma 23. We show that U i = EW i. By (46), \(KE=E\mathcal {M}\). Recall that U i (resp. W i) is the eigenspace of K (resp. \(\mathcal {M}\)) corresponding to the eigenvalue q d−2i. By these comments U i = EW i.

Define \(F=\exp _{q}(\frac {a}{q-q^{-1}}\psi )\). We show \(U_i^\Downarrow =FW_i\). By (47), \(BF=F\mathcal {M}\). Recall that \(U_i^\Downarrow \) (resp. W i) is the eigenspace of B (resp. \(\mathcal {M}\)) corresponding to the eigenvalue q d−2i. By these comments \(U_i^\Downarrow =FW_i\).

To obtain (51) from (50), use (42). □

Lemma 24

For 0 ≤ i  d, the dimension of W i is ρ i.

Proof

This follows from Proposition 2 and the fact that \(U_i,U_i^\Downarrow \) have dimension ρ i. □

Recall from (6) that

$$\displaystyle \begin{aligned} \sum_{h=0}^i E_h^{\ast}V=\sum_{h=0}^i U_h=\sum_{h=0}^i U_h^{\Downarrow} {} \end{aligned} $$
(52)

for 0 ≤ i ≤ d.

Lemma 25

For 0 ≤ i  d, the sum \(\sum _{h=0}^iW_h\) is equal to the common value of (52).

Proof

Define \(W=\sum _{h=0}^i W_h\) and let U denote the common value of (52). We show that W = U. By Lemma 8 and the equation on the left in (51), W ⊆ U. By Lemma 24, W and U have the same dimension. Thus W = U. □

9 The Actions of ψ, K, B,  Δ, A, A on \(\{W_i\}_{i=0}^d\)

We continue to discuss the situation of Assumption 1. Recall the eigenspace decomposition \(\{W_i\}_{i=0}^d\) for \(\mathcal {M}\). In this section, we discuss the actions of ψ, K, B,  Δ, A, A on \(\{W_i\}_{i=0}^d\).

Lemma 26

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \psi W_i\subseteq W_{i-1}. \end{aligned} $$
(53)

Proof

Use Lemma 19. □

Lemma 27

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} (K-q^{d-2i}I)W_i\subseteq W_{i-1},\qquad \qquad (B-q^{d-2i}I)W_i\subseteq W_{i-1}. \end{aligned} $$
(54)

Proof

Use Lemmas 16 and 26. □

Lemma 28

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \begin{array}{rcl} (\Delta-I)W_i&\displaystyle \subseteq W_0+W_1+\cdots +W_{i-1},{} \end{array} \end{aligned} $$
(55)
$$\displaystyle \begin{aligned} \begin{array}{rcl} (\Delta^{-1}-I)W_i&\displaystyle \subseteq W_0+W_1+\cdots +W_{i-1}{}. \end{array} \end{aligned} $$
(56)

Proof

To show (55), use (25) and Lemma 25.

To show (56), use (26) and Lemma 25. □

Lemma 29

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} (A-(a+a^{-1})q^{d-2i}I)W_i\subseteq W_{i-1}+W_{i+1}. \end{aligned} $$
(57)

Proof

By Lemma 22, the expression

$$\displaystyle \begin{aligned} (\mathcal{M}^{-1}-q^{2i+2-d}I)(\mathcal{M}^{-1}-q^{2i-2-d}I)(A-(a+a^{-1})q^{d-2i}I) \end{aligned}$$

vanishes on W i. Therefore \((\mathcal {M}^{-1}-q^{2i+2-d}I)(\mathcal {M}^{-1}-q^{2i-2-d}I)\) vanishes on (A − (a + a −1)q d−2iI)W i. The result follows. □

Lemma 30

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} (A^{\ast}-\theta_i^{\ast}I)W_i\subseteq W_0+W_1+\cdots +W_{i-1}. \end{aligned} $$
(58)

Proof

Use \((A^{\ast }-\theta _i^{\ast }I)E_i^{\ast }V=0\) together with (25) and Lemma 25. □

10 The Actions of \(\mathcal {M}^{\pm 1}\) on \(\{U_i\}_{i=0}^d\),\(\{U_i^\Downarrow \}_{i=0}^d\), \(\{E_iV\}_{i=0}^d\), \(\{E_i^{\ast }V\}_{i=0}^d\)

We continue to discuss the situation of Assumption 1. In Sect. 8 we saw how various operators act on the decomposition \(\{W_i\}_{i=0}^d\). In this section we investigate the action of \(\mathcal {M} \) on the first and second split decompositions of V , as well as on the eigenspace decompositions of A, A .

Lemma 31

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \begin{array}{rcl} (\mathcal{M} -q^{d-2i}I)U_i&\displaystyle \subseteq U_0+U_1+\cdots+U_{i-1},{} \end{array} \end{aligned} $$
(59)
$$\displaystyle \begin{aligned} \begin{array}{rcl} (\mathcal{M} -q^{d-2i}I)U_i^\Downarrow&\displaystyle \subseteq U_0^\Downarrow+U_1^\Downarrow+\cdots+U_{i-1}^\Downarrow.{} \end{array} \end{aligned} $$
(60)

Proof

To show (59), use Definition 5, Lemma 2, and Definition 7.

To show (60), use (59) applied to Φ, along with \(\mathcal {M}^\Downarrow =\mathcal {M}\). □

Lemma 32

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} (\mathcal{M}^{-1}-q^{2i-d}I)U_i\subseteq U_{i-1},\qquad \qquad (\mathcal{M}^{-1}-q^{2i-d}I)U_i^\Downarrow \subseteq U_{i-1}^\Downarrow.{} \end{aligned} $$
(61)

Proof

We first show the equation on the left in (61). By Lemma 17,

$$\displaystyle \begin{aligned} \mathcal{M}^{-1}=(I-a^{-1}q^{-1}\psi)K^{-1}. \end{aligned} $$
(62)

From this and Definition 5, it follows that on U i,

$$\displaystyle \begin{aligned} \mathcal{M}^{-1}-q^{2i-d}I=a^{-1}q^{2i-d-1}\psi. \end{aligned} $$
(63)

The result follows from this along with Lemma 8.

The proof of the equation on the right in (61) follows from the equation on the left in (61) applied to Φ, along with the fact that \(\mathcal {M}^\Downarrow =\mathcal {M}\). □

Lemma 33

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \mathcal{M}^{-1}E_iV\subseteq E_{i-1}V+E_iV+E_{i+1}V. \end{aligned} $$
(64)

Proof

We first show that \(\mathcal {M}^{-1}E_iV\subseteq \sum _{h=0}^{i+1}E_hV\). Recall from (5) that \(E_iV\subseteq \sum _{h=d-i}^d U_h^\Downarrow \). By this, Lemma 32, and (5), we obtain \(\mathcal {M}^{-1}E_iV\subseteq \sum _{h=0}^{i+1} E_hV\).

We now show that \(\mathcal {M}^{-1}E_iV\subseteq \sum _{h=i-1}^d E_hV\). Recall from (4) that \(E_iV\subseteq \sum _{h=i}^d U_h\). By this, Lemma 32, and (4), we obtain \(\mathcal {M}^{-1}E_iV\subseteq \sum _{h=i-1}^d E_hV\).

Thus \(\mathcal {M}^{-1}E_iV\) is contained in the intersection of \(\sum _{h=0}^{i+1} E_hV\) and \(\sum _{h=i-1}^d E_hV\), which is E i−1V + E iV + E i+1V . □

Lemma 34

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \begin{array}{rcl} (\mathcal{M} -q^{d-2i}I)E_i^{\ast}V&\displaystyle \subseteq E_0^{\ast}V+E_1^{\ast}V+\cdots+E_{i-1}^{\ast}V,\\ (\mathcal{M}^{-1} -q^{2i-d}I)E_i^{\ast}V&\displaystyle \subseteq E_0^{\ast}V+E_1^{\ast}V+\cdots+E_{i-1}^{\ast}V. \end{array} \end{aligned} $$

Proof

Note that \(E_i^{\ast }V\subseteq E_0^{\ast }V+E_1^{\ast }V+\cdots +E_i^{\ast }V= W_0+W_1+\cdots +W_i\) by Lemma 25. The result follows from this fact along with Definition 8. □

11 When Φ Is a Leonard System

We continue to discuss the situation of Assumption 1. For the rest of the paper we assume ρ i = 1 for 0 ≤ i ≤ d. In this case Φ is called a Leonard system.

We use the following notational convention. Let \(\{V_i\}_{i=0}^d\) denote a basis for V . The sequence of subspaces \(\{\mathbb {K} v_i\}_{i=0}^d\) is a decomposition of V  said, to be induced by the basis \(\{V_i\}_{i=0}^d\).

We display a basis \(\{U_i\}_{i=0}^d\) (resp. \(\{U_i^\Downarrow \}_{i=0}^d\)) (resp. \(\{W_i\}_{i=0}^d\)) that induces the decomposition \(\{U_i\}_{i=0}^d\) (resp. \(\{U_i^\Downarrow \}_{i=0}^d\)) (resp. \(\{W_i\}_{i=0}^d\)). We find the actions of ψ, K, B, Δ±1, A on these bases. We also display the transition matrices between these bases.

For the rest of this section fix 0 ≠ u 0 ∈ U 0. Let M denote the subalgebra of End(V ) generated by A. By [21, Lemma 5.1], the map M → V , XXu 0 is an isomorphism of vector spaces. Consequently, the vectors \(\{A^iu_0\}_{i=0}^d\) form a basis for V .

We now define a basis \(\{U_i\}_{i=0}^d\) of V  that induces \(\{U_i\}_{i=0}^d\). For 0 ≤ i ≤ d, define

$$\displaystyle \begin{aligned} u_i=\left(\prod_{j=0}^{i-1}\left(A-\theta_jI\right)\right) u_0.{} \end{aligned} $$
(65)

Observe that u i ≠ 0. By [9, Theorem 4.6], u i ∈ U i. So u i is a basis for U i. Consequently, \(\{U_i\}_{i=0}^d\) is a basis for V  that induces \(\{U_i\}_{i=0}^d\).

Next we define a basis \(\{U_i^\Downarrow \}_{i=0}^d\) of V  that induces \(\{U_i^\Downarrow \}_{i=0}^d\). For 0 ≤ i ≤ d, define

$$\displaystyle \begin{aligned} u_i^\Downarrow=\left(\prod_{j=0}^{i-1}\left(A-\theta_{d-j}I\right)\right) u_0.{}\end{aligned} $$
(66)

Observe that \(u_i^\Downarrow \neq 0\). By Lemma 11, \(u_i^\Downarrow \in U_i^\Downarrow \). So \(U_i^\Downarrow \) is a basis for \(U_i^\Downarrow \). Consequently, \(\{U_i^\Downarrow \}_{i=0}^d\) is a basis for V  that induces \(\{U_i^\Downarrow \}_{i=0}^d\).

Lemma 35

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} u_i^\Downarrow=\Delta u_i.\end{aligned} $$
(67)

Proof

By Lemma 11, \(\Delta U_i=U_i^\Downarrow \). So there exists \(0\neq \lambda \in \mathbb {K}\) such that \(\Delta u_i=\lambda u_i^\Downarrow \). We show that λ = 1. By [4, Lemma 7.3] and (25), Δu i − A iu is a linear combination of \(\{A^ju\}_{j=0}^{i-1}\). Also, \(u_i^\Downarrow -A^iu\) is a linear combination of \(\{A^ju\}_{j=0}^{i-1}\). The vectors \(\{A^ju\}_{j=0}^{i-1}\) are linearly independent. By these comments λ = 1. □

We next define a basis \(\{W_i\}_{i=0}^d\) of V  that induces \(\{W_i\}_{i=0}^d\). For 0 ≤ i ≤ d, define

$$\displaystyle \begin{aligned} w_i=\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\ \psi\right) u_i.{}\end{aligned} $$
(68)

Since \(\{u_i \}_{i=0}^d\) is a basis of V  and \(\exp _{q^{-1}}(-\frac {a^{-1}}{q-q^{-1}}\ \psi )\) is invertible, w i is a basis for W i. Consequently, \(\{W_i\}_{i=0}^d\) is a basis for V  that induces \(\{W_i\}_{i=0}^d\).

Lemma 36

For 0 ≤ i  d,

$$\displaystyle \begin{aligned} \begin{array}{rcl} u_i&\displaystyle =\exp_q\left(\frac{a^{-1}}{q-q^{-1}}\psi\right) w_i, \qquad \qquad u_i^\Downarrow&\displaystyle =\exp_q\left(\frac{a}{q-q^{-1}}\psi\right) w_i, {} \end{array} \end{aligned} $$
(69)
$$\displaystyle \begin{aligned} \begin{array}{rcl} w_i&\displaystyle =\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\psi\right) u_i, \qquad \qquad w_i&\displaystyle =\exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\psi\right) u_i^\Downarrow.{} \end{array} \end{aligned} $$
(70)

Proof

Use (68) to obtain the equations on the left in (69),(70). To obtain the equations on the right in (69),(70), use Theorem 3, Lemma 35, and (68). □

We now describe the actions of \(\psi ,K,B,\mathcal {M}, \Delta ,A\) on the bases \(\{U_i\}_{i=0}^d\), \(\{U_i^\Downarrow \}_{i=0}^d\), \(\{W_i\}_{i=0}^d\). First we recall a notion from linear algebra. Let \(\mathrm {Mat}_{d+1}(\mathbb {K})\) denote the \(\mathbb {K}\)-algebra of (d + 1) × (d + 1) matrices that have all entries in \(\mathbb {K}\). We index the rows and columns by 0, 1, …, d. Let \(\{V_i\}_{i=0}^d\) denote a basis of V . For T ∈End(V ) and \(X\in \mathrm {Mat}_{d+1}(\mathbb {K})\), we say that X represents T with respect to \(\{V_i\}_{i=0}^d\) whenever \(Tv_j=\sum _{i=0}^d X_{ij}v_i\) for 0 ≤ j ≤ d.

By (65) and (66), the matrices that represent A with respect to \(\{U_i\}_{i=0}^d\) and \(\{U_i^\Downarrow \}_{i=0}^d\) are, respectively,

(71)

By construction, the matrix diag(q d, q d−2, …, q d) represents K with respect to \(\{U_i\}_{i=0}^d\), and B with respect to \(\{U_i^\Downarrow \}_{i=0}^d\), and \(\mathcal {M}\) with respect to \(\{W_i\}_{i=0}^d\).

Definition 9

We define a matrix \(\widehat {\psi }\in \mathrm {Mat}_{d+1}(\mathbb {K})\). For 1 ≤ i ≤ d, the (i − 1, i)-entry is (q i − q i)(q di+1 − q id−1). All other entries are 0.

Proposition 3

The matrix \(\widehat \psi \) represents ψ with respect to each of the bases \(\{U_i\}_{i=0}^d\), \(\{U_i^\Downarrow \}_{i=0}^d\), \(\{W_i\}_{i=0}^d\).

Proof

By [5, Line (23)], \(\widehat \psi \) represents ψ with respect to \(\{U_i\}_{i=0}^d\). The remaining assertions follow from Lemma 36. □

Next we give the matrices that represent \(\mathcal {M}^{\pm 1}\) with respect to the bases \(\{U_i\}_{i=0}^d\), \(\{U_i^\Downarrow \}_{i=0}^d\).

Lemma 37

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents \(\mathcal {M}\) with respect to \(\{U_i\}_{i=0}^d\). This matrix is upper triangular. For 0 ≤ i  j  d, the (i, j)-entry is

$$\displaystyle \begin{aligned} a^{i-j}q^{d-j-i}\left(q-q^{-1}\right)^{2(j-i)} \frac{\ \ [j]_q^![d-i]_q^!\ \ }{[i]_q^![d-j]_q^!}. \end{aligned} $$
(72)

Proof

The matrix diag(q d, q d−2, …, q d) represents K with respect to \(\{U_i\}_{i=0}^d\). Use this fact along with Lemma 18 and Proposition 3. □

Lemma 38

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents \(\mathcal {M}^{-1}\) with respect to \(\{U_i\}_{i=0}^d\). For 0 ≤ i  d, the (i, i)-entry is q 2id. For 1 ≤ i  d, the (i − 1, i)-entry is

$$\displaystyle \begin{aligned} -a^{-1}q^{2i-d-1}\left(q^i-q^{-i}\right)\left(q^{d-i+1}-q^{i-d-1}\right). \end{aligned}$$

All other entries are zero.

Proof

The matrix diag(q d, q 2−d, …, q d) represents K −1 with respect to \(\{U_i\}_{i=0}^d\). Use this fact along with Lemma 17 and Proposition 3. □

Lemma 39

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents \(\mathcal {M}\) with respect to \(\{U_i^\Downarrow \}_{i=0}^d\). This matrix is upper triangular. For 0 ≤ i  j  d, the (i, j)-entry is

$$\displaystyle \begin{aligned} a^{j-i}q^{d-j-i}\left(q-q^{-1}\right)^{2(j-i)} \frac{\ \ [j]_q^![d-i]_q^!\ \ }{[i]_q^![d-j]_q^!}. \end{aligned} $$
(73)

Proof

The matrix diag(q d, q d−2, …, q d) represents B with respect to \(\{U_i^\Downarrow \}_{i=0}^d\). Use this fact along with Lemma 18 and Proposition 3. □

Lemma 40

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents \(\mathcal {M}^{-1}\) with respect to \(\{U_i^\Downarrow \}_{i=0}^d\). For 0 ≤ i  d, the (i, i)-entry is q 2id. For 1 ≤ i  d, the (i − 1, i)-entry is

$$\displaystyle \begin{aligned} -aq^{2i-d-1}\left(q^i-q^{-i}\right)\left(q^{d-i+1}-q^{i-d-1}\right). \end{aligned}$$

All other entries are zero.

Proof

The matrix diag(q d, q 2−d, …, q d) represents B −1 with respect to \(\{U_i^\Downarrow \}_{i=0}^d\). Use this fact along with Lemma 17 and Proposition 3. □

Next we give the matrices that represent K with respect to the bases \(\{U_i^\Downarrow \}_{i=0}^d\), \(\{W_i\}_{i=0}^d\).

Lemma 41

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents K with respect to \(\{U_i^\Downarrow \}_{i=0}^d\). For 0 ≤ i  d, the (i, i)-entry is q d−2i. For 0 ≤ i < j  d, the (i, j)-entry is

$$\displaystyle \begin{aligned} \left(1-a^{-2}\right)a^{j-i}q^{d-j-i} \left(q-q^{-1}\right)^{2(j-i)} \frac{\ \ [j]_q^![d-i]_q^!\ \ }{[i]_q^![d-j]_q^!}. \end{aligned} $$
(74)

All other entries are zero.

Proof

Evaluating the equation on the right in (14) using the equation on the left in (12) we get

$$\displaystyle \begin{aligned} K=\left(a^{-2}I+(1-a^{-2})\sum_{n=0}^d a^nq^{n}\psi^n\right)B. {} \end{aligned} $$
(75)

The result follows from this along with Proposition 3 and the fact that the matrix diag(q d, q d−2, …, q d) represents B with respect to \(\{U_i^\Downarrow \}_{i=0}^d\). □

Lemma 42

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents K with respect to \(\{W_i\}_{i=0}^d\). For 0 ≤ i  d, the (i, i)-entry is q d−2i. For 1 ≤ i  d, the (i − 1, i)-entry is

$$\displaystyle \begin{aligned} -a^{-1}q^{d-2i+1} (q^{i}-q^{-i})(q^{d-i+1}-q^{i-d-1}). \end{aligned}$$

All other entries are zero.

Proof

The matrix diag(q d, q d−2, …, q d) represents \(\mathcal {M}\) with respect to \(\{W_i\}_{i=0}^d\). Use this fact along with Proposition 3 and the equation on the left in (32). □

Next we give the matrices that represent B with respect to the bases \(\{U_i\}_{i=0}^d\), \(\{W_i\}_{i=0}^d\).

Lemma 43

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents B with respect to \(\{U_i\}_{i=0}^d\). For 0 ≤ i  d, the (i, i)-entry is q d−2i. For 0 ≤ i < j  d, the (i, j)-entry is

$$\displaystyle \begin{aligned} \left(1-a^{2}\right)a^{i-j}q^{d-j-i}\left(q-q^{-1}\right)^{2(j-i)} \frac{\ \ [j]_q^![d-i]_q^!\ \ }{[i]_q^![d-j]_q^!}. \end{aligned} $$
(76)

All other entries are zero.

Proof

Evaluating the equation on the left in (14) using the equation on the right in (12) we get

$$\displaystyle \begin{aligned} B=\left(a^{2}I+(1-a^{2})\sum_{n=0}^d a^{-n}q^{n}\psi^n\right)K. {} \end{aligned} $$
(77)

The result follows from this along with Proposition 3 and the fact that the matrix diag(q d, q d−2, …, q d) represents K with respect to \(\{U_i\}_{i=0}^d\). □

Lemma 44

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents B with respect to \(\{W_i\}_{i=0}^d\). For 0 ≤ i  d, the (i, i)-entry is q d−2i. For 1 ≤ i  d, the (i − 1, i)-entry is

$$\displaystyle \begin{aligned} -aq^{d-2i+1} (q^{i}-q^{-i})(q^{d-i+1}-q^{i-d-1}). \end{aligned}$$

All other entries are zero.

Proof

The matrix diag(q d, q d−2, …, q d) represents \(\mathcal {M}\) with respect to \(\{W_i\}_{i=0}^d\). Use this fact along with Proposition 3 and the equation on the left in (33). □

Next we consider the matrices

$$\displaystyle \begin{aligned} \exp_q\left(\frac{a}{q-q^{-1}}\widehat{\psi}\right),\qquad \qquad \exp_q\left(\frac{a^{-1}}{q-q^{-1}}\widehat{\psi}\right).{} \end{aligned} $$
(78)

Their inverses are

$$\displaystyle \begin{aligned} \exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\widehat{\psi}\right),\qquad \qquad \exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\widehat{\psi}\right){} \end{aligned} $$
(79)

respectively. The matrices in (78), (79) are upper triangular. We now consider the entries of (78), (79).

Lemma 45

For \(0\neq x\in \mathbb {K}\), the matrix \(\exp _q(x\widehat {\psi })\) is upper triangular. For 0 ≤ i  j  d, the (i, j)-entry is

$$\displaystyle \begin{aligned} x^{j-i}q^{\binom{j-i}{2}}\left(q-q^{-1}\right)^{2(j-i)}\cdot \frac{ [j]_q^![d-i]_q^! }{[i]_q^![j-i]_q^![d-j]_q^!}. \end{aligned} $$
(80)

The matrix \(exp_{q^{-1}}(x\widehat {\psi })\) is upper triangular. For 0 ≤ i  j  d, the (i, j)-entry is

$$\displaystyle \begin{aligned} x^{j-i}q^{-\binom{j-i}{2}}\left(q-q^{-1}\right)^{2(j-i)}\cdot \frac{ [j]_q^![d-i]_q^! }{[i]_q^![j-i]_q^![d-j]_q^!}. \end{aligned} $$
(81)

Lemma 46

The transition matrices between the basis \(\{W_i\}_{i=0}^d\) and the bases \(\{U_i\}_{i=0}^d\), \(\{U_i^\Downarrow \}_{i=0}^d\) are given in the table below.

From

To

Transition matrix

\(\{U_i\}_{i=0}^d\)

\(\{W_i\}_{i=0}^d\)

\(\exp _{q^{-1}}\left (-\frac {a^{-1}}{q-q^{-1}}\widehat \psi \right )\)

\(\{W_i\}_{i=0}^d\)

\( \{u_i\}_{i=0}^d\)

\(\exp _q\left (\frac {a^{-1}}{q-q^{-1}}\widehat \psi \right )\)

\(\{U_i^\Downarrow \}_{i=0}^d\)

\(\{W_i\}_{i=0}^d\)

\(\exp _{q^{-1}}\left (-\frac {a}{q-q^{-1}}\widehat \psi \right )\)

\(\{W_i\}_{i=0}^d\)

\(\{U_i^\Downarrow \}_{i=0}^d\)

\(\exp _q\left (\frac {a}{q-q^{-1}}\widehat \psi \right )\)

Proof

Use Lemma 36 and Proposition 3. □

We next consider the product

$$\displaystyle \begin{aligned} \exp_q\left(\frac{a}{q-q^{-1}}\widehat\psi\right)\exp_{q^{-1}}\left(-\frac{a^{-1}}{q-q^{-1}}\widehat\psi\right).{} \end{aligned} $$
(82)

The inverse of (82) is

$$\displaystyle \begin{aligned} \exp_q\left(\frac{a^{-1}}{q-q^{-1}}\widehat\psi\right)\exp_{q^{-1}}\left(-\frac{a}{q-q^{-1}}\widehat\psi\right). {} \end{aligned} $$
(83)

The matrices in (82), (83) are upper triangular.

Lemma 47

The transition matrices between the bases \(\{U_i\}_{i=0}^d\), \(\{U_i^\Downarrow \}_{i=0}^d\) are given in the table below.

From

To

Transition matrix

\(\{U_i\}_{i=0}^d\)

\(\{u^\Downarrow _i\}_{i=0}^d\)

\(\exp _q\left (\frac {a}{q-q^{-1}}\widehat \psi \right )\exp _{q^{-1}}\left (-\frac {a^{-1}}{q-q^{-1}}\widehat \psi \right )\)

\(\{u^\Downarrow _i\}_{i=0}^d\)

\(\{U_i\}_{i=0}^d\)

\(\exp _q\left (\frac {a^{-1}}{q-q^{-1}}\widehat \psi \right )\exp _{q^{-1}}\left (-\frac {a}{q-q^{-1}}\widehat \psi \right )\)

Proof

Use Lemma 46. □

Lemma 48

With respect to each of the bases \(\{u_i\}_{i=0}^d, \{u_i^\Downarrow \}_{i=0}^d, \{w_i\}_{i=0}^d\) , the matrices that represent Δ and Δ −1 are \(\exp _q\left (\frac {a}{q-q^{-1}}\widehat \psi \right )\exp _{q^{-1}}\left (-\frac {a^{-1}}{q-q^{-1}}\widehat \psi \right )\) and

\(\exp _q\left (\frac {a^{-1}}{q-q^{-1}}\widehat \psi \right )\exp _{q^{-1}}\left (-\frac {a}{q-q^{-1}}\widehat \psi \right )\) respectively.

Proof

Use Theorem 3 and Proposition 3. □

We give the entries of the matrices representing Δ, Δ−1 in the following lemma.

Lemma 49

The matrix in (82) is upper triangular. For 0 ≤ i  j  d, the (i, j)-entry of (82) is

$$\displaystyle \begin{aligned} \frac{\left(q-q^{-1}\right)^{j-i} [j]_q^![d-i]_q^! }{[i]_q^![j-i]_q^![d-j]_q^! } \ \prod_{n=1}^{j-i} \left(aq^{n-1}-a^{-1}q^{1-n}\right). \end{aligned} $$
(84)

The matrix in (83) is upper triangular. For 0 ≤ i  j  d, the (i, j)-entry of (83) is

$$\displaystyle \begin{aligned} \frac{\left(q-q^{-1}\right)^{j-i} [j]_q^![d-i]_q^! }{[i]_q^![j-i]_q^![d-j]_q^! } \ \prod_{n=1}^{j-i} \left(a^{-1}q^{n-1}-aq^{1-n}\right). \end{aligned} $$
(85)

Proof

Use Corollary 1 and Proposition 3. □

We finish the paper by giving the matrix that represents A with respect to \(\{W_i\}_{i=0}^d\).

Lemma 50

We give the matrix in \(\mathrm {Mat}_{d+1}(\mathbb {K})\) that represents A with respect to \(\{W_i\}_{i=0}^d\). For 1 ≤ i  d, the (i, i − 1)-entry is 1. For 0 ≤ i  d, the (i, i)-entry is (a + a −1)q d−2i. For 1 ≤ i  d, the (i − 1, i)-entry is

$$\displaystyle \begin{aligned} -q^{d-2i+1}(q^i-q^{-i})(q^{d-i+1}-q^{i-d-1}). \end{aligned}$$

All other entries are zero.

Proof

Let \(\mathcal {A}\) denote the matrix that represents A with respect to \(\{W_i\}_{i=0}^d\). By Lemma 29, \(\mathcal {A}\) is tridiagonal with (i, i)-entry given by (a + a −1)q d−2i for 0 ≤ i ≤ d.

We now show that the subdiagonal entries of \(\mathcal {A}\) are all 1. Let \(\mathcal {A}^{\prime }\) denote the matrix that represents A with respect to \(\{U_i\}_{i=0}^d\). Recall that this matrix is displayed on the left in (71). Observe that \(\mathcal {A}\) is equal to \(\exp _{q^{-1}}(-\frac {a^{-1}}{q-q^{-1}}\widehat \psi ) \mathcal {A}' \exp _{q}(\frac {a^{-1}}{q-q^{-1}}\widehat \psi )\). It follows from this fact that the subdiagonal entries of \(\mathcal {A}\) are all 1.

We next obtain the superdiagonal entries of \(\mathcal {A}\). Let 0 ≤ i ≤ d. Apply both sides of (37) to w i. Evaluate the result using Proposition 3 and the fact that the w i is an eigenvector for \(\mathcal {M}\) with eigenvalue q 2id. Analyze the result in light of the above comments concerning the entries of \(\mathcal {A}\) to obtain the desired result. □