1 Introduction and Preliminaries

The Bernoulli polynomials \( \left( B_{n}(x)\right) _{n}\) are defined by the generating function

$$\begin{aligned} \frac{te^{xt}}{e^{t}-1}=\sum _{n=0}^{\infty }B_{n}(x)\frac{t^{n}}{n!},\quad \mid t\mid <2\pi . \end{aligned}$$

In a series of papers, Frappier [9,10,11] studied the generalized Bernoulli polynomials \( B_{n,\alpha }(x)\), defined by the generating function

$$\begin{aligned} \frac{e^{(x-\frac{1}{2})t}}{g_{\alpha }(\frac{it}{2})}=\sum _{n=0}^{\infty }B_{n,\alpha }(x)\frac{t^{n}}{n!},\quad \mid t\mid <2 j_{1,\alpha }, \end{aligned}$$
(1.1)

where

$$\begin{aligned} g_{\alpha }(t)=2^{\alpha }\Gamma (\alpha +1)\frac{ J_{\alpha }(t)}{t^{\alpha }}, \end{aligned}$$

\( J_{\alpha }(t)\) is the Bessel function of the first kind of order \(\alpha ,\) and \(j_{1,\alpha } \) is the smallest positive zero of \(J_{\alpha }(t)\). Ismail and Mansour, see [19], introduced a pair of q-analogs of the Bernoulli polynomials by the generating functions

$$\begin{aligned} \begin{aligned} \frac{te_{q}(xt)}{e_{q}(\frac{t}{2})E_{q}(\frac{t}{2})-1}&=\sum _{n=0}^{\infty }b_{n}(x;q)\frac{t^{n}}{[n]_{q}!},\\ \frac{t E_{q}(xt)}{e_{q}(\frac{t}{2})E_{q}(\frac{t}{2})-1}&=\sum _{n=0}^{\infty }B_{n}(x;q)\frac{t^{n}}{[n]_{q}!}.\end{aligned} \end{aligned}$$
(1.2)

They also defined a pair of q-analogs of the Euler polynomials by the generating functions

$$\begin{aligned} \begin{aligned} \frac{2e_{q}(xt)}{E_{q}(\frac{t}{2})e_{q}(\frac{t}{2})+1}&=\sum _{n=0}^{\infty }e_{n}(x;q)\frac{t^{n}}{[n]_{q}!},\\ \frac{2 E_{q}(xt)}{E_{q}(\frac{t}{2})e_{q}(\frac{t}{2})+1}&=\sum _{n=0}^{\infty }E_{n}(x;q)\frac{t^{n}}{[n]_{q}!},\end{aligned} \end{aligned}$$
(1.3)

where

$$\begin{aligned}\left[ n\right] _q!=\frac{(q;q)_{n}}{(1-q)^{n}}\quad (n\in {\mathbb {N}}), \quad (a;q)_{n}=\left\{ \begin{array}{ll} 1, &{}{} {n=0;} \\ \displaystyle \prod _{k=0}^{n-1}(1-aq^{k}), &{}{} {n\in {\mathbb {N}},} \end{array} \right. \end{aligned}$$

and \(a\in {\mathbb {C}}\), see [12]. The functions \(E_q(x)\) and \(e_q(x)\) are the q-analogs of the exponential functions defined by

$$\begin{aligned} \begin{aligned} E_q(x):= (-x(1-q);q)_{\infty }&=\sum _{n=0}^{\infty }\frac{q^{\frac{n(n-1)}{2}}(1-q)^{n}x^{n}}{(q;q)_{n}}, \,\, x\in {\mathbb {C}}, \\ e_q(x):=\frac{1}{(x(1-q);q)_{\infty }}&=\sum _{n=0}^{\infty }\frac{(1-q)^{n}x^{n}}{(q;q)_{n}}, \,\, \mid x\mid < \frac{1}{1-q}, \end{aligned} \end{aligned}$$
(1.4)

see [12].

In [22], Mansour and Al-Towalib introduced q-analogs of Bernoulli and Euler polynomials by the generating functions

$$\begin{aligned} \begin{aligned}&\frac{t \exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{\exp _{q}(\frac{t}{2})-\exp _{q}(\frac{-t}{2})}=\sum _{n=0}^{\infty }{\tilde{B}}_{n}(x;q)\frac{t^{n}}{[n]_{q}!},\\&\frac{2 \exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{\exp _{q}(\frac{t}{2})+\exp _{q}(\frac{-t}{2})}=\sum _{n=0}^{\infty }{\tilde{E}}_{n}(x;q)\frac{t^{n}}{[n]_{q}!}, \end{aligned} \end{aligned}$$
(1.5)

where

$$\begin{aligned} \exp _{q}(x)=\sum _{n=0}^{\infty }q^{\frac{n(n-1)}{4}}\frac{x^{n}}{[n]_{q}!},\quad x\in {\mathbb {C}}, \end{aligned}$$

is a q-analog of the exponential function. This q-exponential function has the property \(\lim _{q\rightarrow 1}\exp _{q}(x) = e^{x}\) for \(x \in {\mathbb {C}}\). It is an entire function of x of order zero, see [12, Eq. (1.3.27), p. 12].

In this paper, we use \({\mathbb {N}} \) to denote the set of positive integers and \({\mathbb {N}}_{0} \) to denote the set of non-negative integers. Throughout this paper, unless otherwise is stated, q is a positive number that is less than one. We follow Gasper and Rahman [12] to define the q-shifted factorial, the q-binomial coefficients, and the q-gamma function. The q-integer number \([n]_{q}\) is defined by

$$\begin{aligned}{}[n]_{q}=\frac{1-q^n}{1-q},\quad n\in {\mathbb {N}}_0. \end{aligned}$$

Jackson in [20] defined the q-difference operator by

$$\begin{aligned} D_q f(z)= \dfrac{f(qz)-f(z)}{z(q-1)}, \quad z\ne 0. \end{aligned}$$

The symmetric q-difference operator is defined by, see [8, 12],

$$\begin{aligned} \delta _{q,z} f(z)= \frac{{}f(q^{\frac{1}{2}}z)-f(q^{\frac{-1}{2}}z)}{(q^{\frac{1}{2}}-q^{\frac{-1}{2}})z}, \quad z\ne 0. \end{aligned}$$

The q-trigonometric functions \(\sin _{q}z,\, \cos _{q}z,\,Sin_{q}z\) and \(Cos_{q} z\) are defined by

$$\begin{aligned} \begin{aligned}&\sin _{q}z=\dfrac{e_{q}(iz)-e_{q}(-iz)}{2i},\quad \cos _{q}z=\dfrac{e_{q}(iz)+e_{q}(-iz)}{2},\,\,\mid z\mid <1,\\&Sin_{q}z=\dfrac{E_{q}(iz)-E_{q}(-iz)}{2i},\quad Cos_{q}z=\dfrac{E_{q}(iz)+E_{q}(-iz)}{2},\,\,z\in {\mathbb {C}}, \end{aligned} \end{aligned}$$

see [5, 12]. The q-sine and cosine functions \(S_{q}(z),\,\, C_{q}(z)\) are defined by the q-Euler formula

$$\begin{aligned} \exp _{q}(iz):=C_{q}(z)+iS_{q}(z), \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} C_{q}(z)=\sum _{n=0}^{\infty }(-1)^{n}\frac{q^{n(n-\frac{1}{2})}}{[2n]_{q}!}z^{2n},\quad S_{q}(z)=\sum _{n=0}^{\infty }(-1)^{n}\frac{q^{n(n+\frac{1}{2})}}{[2n+1]_{q}!}z^{2n+1}, \end{aligned} \end{aligned}$$

cf. [8, p. 2]. The hyperbolic functions \(Sh_{q}(z)\) and \(Ch_{q}(z)\) are defined for \(z\in {\mathbb {C}}\) by

$$\begin{aligned} \begin{aligned}&Sh_{q}(z):=-iS_{q}(iz)=\frac{\exp _{q}(z)-\exp _{q}(-z)}{2},\\&Ch_{q}(z):= C_{q}(iz)=\frac{\exp _{q}(z)+\exp _{q}(-z)}{2}. \end{aligned} \end{aligned}$$
(1.6)

There are three known q-analogs of the Bessel function that are due to Jackson [20]. These are denoted by \( J^{(k)}_{\alpha }(t;q)\, (k=1,2,3)\) and defined by

$$\begin{aligned} J_{\alpha }^{(1)}(t;q)= & {} \frac{(q^{\alpha +1};q)_{\infty }}{(q;q)_{\infty }}\sum _{n=0}^{\infty }(-1)^n\frac{(\frac{t}{2})^{2n+\alpha }}{(q;q)_{n}(q^{\alpha +1};q)_{n}} \quad (\mid t\mid <2), \\ J_{\alpha }^{(2)}(t;q)= & {} \frac{(q^{\alpha +1};q)_{\infty }}{(q;q)_{\infty }}\sum _{n=0}^{\infty }(-1)^n\frac{q^{n(\alpha +n)}(\frac{t}{2})^{2n+\alpha }}{(q;q)_{n}(q^{\alpha +1};q)_{n}}\quad (t\in {\mathbb {C}} ), \\ J_{\alpha }^{(3)}(t;q)= & {} \frac{(q^{\alpha +1};q)_{\infty }}{(q;q)_{\infty }}\sum _{n=0}^{\infty }(-1)^n\frac{q^{\frac{n(n+1)}{2}}t^{2n+\alpha }}{(q;q)_{n}(q^{\alpha +1};q)_{n}} \quad (t\in {\mathbb {C}}). \end{aligned}$$

For convenience, we set

$$\begin{aligned} \begin{aligned}&{\mathcal {J}}_{\alpha }^{(k)} (t;q):=\frac{(q;q)_{\infty }}{(q^{\alpha +1};q)_{\infty }}(\frac{t}{2})^{-\alpha } J_{\alpha }^{(k)} (t;q)\quad ( k=1,2),\\&{\mathcal {J}}_{\alpha }^{(3)} (t;q):=\frac{(q;q)_{\infty }}{(q^{\alpha +1};q)_{\infty }}t^{-\alpha } J_{\alpha }^{(3)}(t;q).\end{aligned} \end{aligned}$$
(1.7)

The functions \({\mathcal {J}}_{\alpha }^{(k)} (t;q)\, (k=1,2,3)\) are called the modified Jackson q-Bessel functions. From now on, we use \((j^{(k)}_{m,\alpha })_{m=1}^{\infty }\) to denote the positive zeros of \(J^{(k)}_{\alpha }(\cdot ;q^{2})\) arranged in increasing order of magnitude. Consequently, \(j^{(k)}_{1,\alpha }\) is the smallest positive zero of \(J^{(k)}_{\alpha }(\cdot ;q^{2})\,( k=1,2,3)\).

This paper is organized as follows. In Sect. 2, we introduce three q-analogs of the generalized Bernoulli polynomials defined in (1.1). The generating functions of these q-analogs include the three q-analogs of Jackson q-Bessel functions mentioned above. We also include the main properties of these q-analogs. Section 3 introduces a q-Fourier expansion for the generalized Bernoulli numbers related to the first and second Jackson q-Bessel functions. Also, their large n degree asymptotic is derived. Finally, in Sect. 4 as an application, we introduce the connection coefficients between q-analogs and certain q-orthogonal polynomials.

2 Generalized q-Bernoulli Polynomials Generated by Jackson q-Bessel Functions

This section introduces three q-analogs of the generalized Bernoulli polynomials introduced by Frappier in [9,10,11].

Definition 2.1

The generalized q-Bernoulli polynomials \(B^{(k)}_{n,\alpha }(x;q)\,( k=1,2,3)\) are defined by the generating functions

$$\begin{aligned}&\frac{e_{q}(xt)e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!},\quad \mid t\mid < \frac{ j^{(1)}_{1,\alpha }}{1-q}, \end{aligned}$$
(2.1)
$$\begin{aligned}&\frac{E_{q}(xt)E_{q}(\frac{-t}{2})}{g^{(2)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!},\quad \mid t\mid <\frac{ j^{(2)}_{1,\alpha }}{1-q}, \end{aligned}$$
(2.2)
$$\begin{aligned}&\frac{\exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{g^{(3)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }B^{(3)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}, \quad \mid t\mid <\frac{2q^{\frac{1}{4}}j^{(3)}_{1,\alpha }}{1-q}, \end{aligned}$$
(2.3)

where \(g^{(k)}_{\alpha }(t;q) \,( k=1,2,3 )\) are the functions defined for \((k=1,2)\) by

$$\begin{aligned} \begin{aligned} g^{(k)}_{\alpha }(t;q):=(1+q)^{\alpha }\Gamma _{q^{2}}(\alpha +1)\,\left( \frac{t}{2}\right) ^{-\alpha }\,J_{\alpha }^{(k)}(t(1-q);q^{2})={\mathcal {J}}_{\alpha }^{(k)} (t(1-q);q^{2}), \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} g^{(3)}_{\alpha }(t;q)\,{{:\!}=}\,&(1+q)^{\alpha }\Gamma _{q^{2}}(\alpha +1)\,\left( \frac{ q^{\frac{-1}{4}}t}{2}\right) ^{-\alpha }\,J_{\alpha }^{(3)}\left( \frac{t}{2}(1-q)q^{\frac{-1}{4}};q^{2}\right) \\ =\,&{\mathcal {J}}_{\alpha }^{(3)}\left( \frac{t}{2}(1-q)q^{\frac{-1}{4}};q^{2}\right) . \end{aligned} \end{aligned}$$

Since the generating functions in (1.2), (1.3), and (1.5) can be written as

$$\begin{aligned}&\frac{te_{q}(xt)e_{q}(\frac{-t}{2})}{2\sinh _{q}\frac{t}{2}}=\sum _{n=0}^{\infty }b_{n}(x;q)\frac{t^{n}}{[n]_{q}!},\nonumber \\&\frac{t E_{q}(xt)E_{q}(\frac{-t}{2})}{2Sinh_{q}\frac{t}{2}}=\sum _{n=0}^{\infty }B_{n}(x;q)\frac{t^{n}}{[n]_{q}!}, \end{aligned}$$
(2.4)
$$\begin{aligned}&\frac{e_{q}(xt)e_{q}(\frac{-t}{2})}{\cosh _{q}\frac{t}{2}}=\sum _{n=0}^{\infty }e_{n}(x;q)\frac{t^{n}}{[n]_{q}!},\nonumber \\&\frac{ E_{q}(xt)E_{q}(\frac{-t}{2})}{Cosh_{q}\frac{t}{2}}=\sum _{n=0}^{\infty }E_{n}(x;q)\frac{t^{n}}{[n]_{q}!}, \end{aligned}$$
(2.5)

and

$$\begin{aligned} \begin{aligned}&\frac{t \exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{2Sh_{q}(\frac{t}{2})}=\sum _{n=0}^{\infty }{\tilde{B}}_{n}(x;q)\frac{t^{n}}{[n]_{q}!},\\&\frac{ \exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{Ch_{q}(\frac{t}{2})}=\sum _{n=0}^{\infty }{\tilde{E}}_{n}(x;q)\frac{t^{n}}{[n]_{q}!}, \end{aligned} \end{aligned}$$
(2.6)

then, if we substitute with \(\alpha =\pm \frac{1}{2}\) in (2.1), (2.2), and (2.3), we obtain the q-Bernoulli and Euler polynomials defined in (2.4), (2.5) and (2.6), respectively.

Lemma 2.2

For \(n\in {\mathbb {N}}_{0}\) and \(Re \,\alpha >-1\),

$$\begin{aligned} \frac{e_{q}(\frac{-t}{2})}{ g^{(1)}_{\alpha }(it;q)}=\frac{E_{q}(\frac{-t}{2})}{ g^{(2)}_{\alpha }(it;q)},\,\,\, \mid t \mid <\frac{1}{1-q}\min \{ j^{(1)}_{1,\alpha }, j^{(2)}_{1,\alpha },2\}. \end{aligned}$$

Proof

Hahn in [14] proved the identity

$$\begin{aligned} J^{(2)}_{\alpha }(t;q)=\left( \frac{-t^{2}}{4};q\right) _{\infty }\,J^{(1)}_{\alpha }(t;q),\quad \mid t\mid <2. \end{aligned}$$
(2.7)

Since

$$\begin{aligned} \begin{aligned}&g^{(k)}_{\alpha }(it;q)=(1+q)^{\alpha }\Gamma _{q^{2}}(\alpha +1)\,\left( \frac{it}{2}\right) ^{-\alpha }\,J_{\alpha }^{(k)}(it(1-q);q^{2})\,\, (k=1,2),\end{aligned} \end{aligned}$$
(2.8)

then, substituting from (2.8) into (2.7), we conclude that

$$\begin{aligned} g^{(2)}_{\alpha }(it;q)=\left( \frac{t^{2}}{4}(1-q)^{2};q^{2}\right) _{\infty }\,{g^{(1)}_{\alpha }(it;q)}= E_{q}\left( \frac{t}{2}\right) E_{q}\left( \frac{-t}{2}\right) \, {g^{(1)}_{\alpha }(it;q)}.\nonumber \\ \end{aligned}$$
(2.9)

Hence

$$\begin{aligned} \begin{aligned} \frac{E_{q}(\frac{-t}{2})}{ g^{(2)}_{\alpha }(it;q)}= \frac{E_{q}(\frac{-t}{2})}{ E_{q}(\frac{t}{2})E_{q}(\frac{-t}{2})\, g^{(1)}_{\alpha }(it;q)}=\frac{e_{q}(\frac{-t}{2})}{ g^{(1)}_{\alpha }(it;q)},\end{aligned} \end{aligned}$$

which completes the proof. \(\square \)

Definition 2.3

The generalized q-Bernoulli numbers \({\beta }_{n,\alpha }(q),\) \( \beta ^{(3)}_{n,\alpha }(q)\) are defined respectively in terms of the generating functions

$$\begin{aligned} \frac{e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}&=\frac{E_{q}(\frac{-t}{2})}{g^{(2)}_{\alpha }(it;q)} =\sum _{n=0}^{\infty }\beta _{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}, \end{aligned}$$
(2.10)
$$\begin{aligned} \frac{\exp _{q}(\frac{-t}{2})}{g^{(3)}_{\alpha }(it;q)}&=\sum _{n=0}^{\infty }\beta ^{(3)}_{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.11)

Proposition 2.4

For \(n\in {\mathbb {N}}\), we have

$$\begin{aligned} B^{(k)}_{2n+1,\alpha }\left( \frac{1}{2};q\right) =0\quad (k=1,2,3). \end{aligned}$$

Proof

If we substitute with \(x=\frac{1}{2}\) in Eqs. (2.1)–(2.3), we find that their left hand side are even functions. Therefore, the coefficients of the odd powers of \(t^{n}\) on the right hand sides of Eqs. (2.1)–(2.3) vanish. This proves the proposition. \(\square \)

Proposition 2.5

For \( k\in \{1,2,3\}\) and \(n\in {\mathbb {N}}\), the polynomials \(B_{n,\alpha }^{(k)}(x;q)\) have the representation \(B_{0,\alpha }^{(k)}(x;q)=1\),

$$\begin{aligned} B^{(1)}_{n,\alpha }(x;q)= & {} \sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \beta _{n-k,\alpha } (q) x^{k}, \end{aligned}$$
(2.12)
$$\begin{aligned} \quad \quad \, B^{(2)}_{n,\alpha }(x;q)= & {} \sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} q^{\frac{k(k-1)}{2}}\beta _{n-k,\alpha }(q) x^{k}, \end{aligned}$$
(2.13)
$$\begin{aligned} \quad \quad \, B^{(3)}_{n,\alpha }(x;q)= & {} \sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} q^{\frac{k(k-1)}{4}}\beta ^{(3)}_{n-k,\alpha }(q) x^{k}. \end{aligned}$$
(2.14)

Proof

We prove the case \((k=1)\). The proofs for \((k=2,3)\) are similar and are omitted. Substituting with the series representation of \(e_{q}(x)\) from (1.4) into (2.1) gives

$$\begin{aligned} \begin{aligned} \sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}&=\frac{e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}e_{q}(xt)\\&=\left( \sum _{n=0}^{\infty } \beta _{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}\right) \left( \sum _{n=0}^{\infty }\frac{(xt)^{n}}{[n]_{q}!}\right) . \end{aligned} \end{aligned}$$

Hence

$$\begin{aligned} \sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}= \sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\sum _{k=0}^{n} \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \beta _{n-k,\alpha }\,x^{k}, \end{aligned}$$
(2.15)

where we applied the Cauchy product formula. Equating the nth power of t in (2.15), we obtain (2.12). \(\square \)

Proposition 2.6

For \(n\in {\mathbb {N}}\) and \( k\in \{1,2,3\}\), the polynomials \(B_{n,\alpha }^{(k)}(x;q)\) satisfy the q-difference equations

$$\begin{aligned} D_{q,x}\,B^{(1)}_{n,\alpha }(x;q)&=[n]_{q}\,B^{(1)}_{n-1,\alpha }(x;q), \end{aligned}$$
(2.16)
$$\begin{aligned} D_{q^{-1},x}\,B^{(2)}_{n,\alpha }(x;q)&=[n]_{q}\,B^{(2)}_{n-1,\alpha }(x;q), \end{aligned}$$
(2.17)
$$\begin{aligned} \delta _{q,x}\,B^{(3)}_{n,\alpha }(x;q)&=[n]_{q}B^{(3)}_{n-1,\alpha }(x;q). \end{aligned}$$
(2.18)

Proof

We only prove the case (\(k=1\)) and the proofs of \((k=2,3)\) are similar. Calculating the q-derivative of both sides of (2.1) with respect to the variable x and taking into consideration that

$$\begin{aligned} D_{q,x}\,e_{q}(xt)=t\,e_{q}(xt), \end{aligned}$$

we obtain

$$\begin{aligned} \frac{t e_{q}(xt)e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}=\sum _{n=1}^{\infty }D_{q,x} B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$

Therefore,

$$\begin{aligned} \sum _{n=0}^{\infty } B^{(1)}_{n,\alpha }(x;q)\frac{t^{n+1}}{[n]_{q}!}=\sum _{n=1}^{\infty }D_{q,x} B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.19)

Equating the corresponding nth power of t in (2.19), we obtain (2.16). \(\square \)

Corollary 2.7

Let \(n\in {\mathbb {N}}\) and k be a positive integer such that \(k\le n\). Then for \(x\in {\mathbb {C}},\)

$$\begin{aligned} \begin{aligned} D^{k}_{q,x}\,\frac{B^{(1)}_{n,\alpha }(x;q)}{[n]_{q}!}&=\frac{B^{(1)}_{n-k,\alpha }(x;q) }{[n-k]_{q}!},\\ D^{k}_{q^{-1},x}\,\frac{B^{(2)}_{n,\alpha }(x;q)}{[n]_{q}!}&=\frac{B^{(2)}_{n-k,\alpha }(x;q) }{[n-k]_{q}!},\\ \delta ^{k}_{q,x}\,\frac{B^{(3)}_{n,\alpha }(x;q)}{[n]_{q}!}&=\frac{B^{(3)}_{n-k,\alpha }(x;q)}{[n-k]_{q}!}.\end{aligned} \end{aligned}$$

Proof

The proofs follow from Proposition 2.6 and the mathematical induction. \(\square \)

Proposition 2.8

For \(\mid t\mid <\frac{1}{1-q}\min \{ j^{(1)}_{1,\alpha }, j^{(2)}_{1,\alpha },2\}\),

$$\begin{aligned}&\sum _{n=0}^{\infty } B^{(1)}_{n,\alpha }\left( \frac{1}{2};q\right) \frac{t^{n}}{[n]_{q}!}=\frac{1}{g^{(2)}_{\alpha }(it;q)}. \end{aligned}$$
(2.20)
$$\begin{aligned}&\sum _{n=0}^{\infty } B^{(2)}_{n,\alpha }\left( \frac{1}{2};q\right) \frac{t^{n}}{[n]_{q}!}=\frac{1}{g^{(1)}_{\alpha }(it;q)}. \end{aligned}$$
(2.21)

Proof

Set \( x=\frac{1}{2} \) in (2.1), we obtain

$$\begin{aligned} \frac{e_{q}(\frac{t}{2})e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }\left( \frac{1}{2};q\right) \frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.22)

Substituting from (2.9) into (2.22), we obtain (2.20). Similarly, we can prove (2.21). \(\square \)

The following Lemma from [22] gives the reciprocal of \( \exp _{q}(z)\) in a certain domain.

Lemma 2.9

Let \(z\in \Omega , \,\displaystyle \Omega := \{ z\in {\mathbb {C}}: \, \mid 1-\exp _{q}(-z) \mid < 1\}\). Then

$$\begin{aligned} \dfrac{1}{\exp _{q}(z)}:=\sum _{n=0}^{\infty } c_n\,z^{n}, \end{aligned}$$

where

$$\begin{aligned} c_n=\, \sum _{k=1}^{n} (-1)^{k}\sum _{\begin{array}{c} {s_1+s_2+\ldots +s_k=n} \\ {s_i>0\,(i=1,\ldots ,k)} \end{array}} \dfrac{ q^{ \sum _{i=1}^k s_i(s_i-1)/4 }}{[s_1]_q! [s_2]_q!\ldots [s_{k}]_q!}. \end{aligned}$$
(2.23)

Proposition 2.10

For \( Re\,\alpha >-1 \) and \(t\in \displaystyle \Omega = \{ t\in {\mathbb {C}}: \, \mid 1-\exp _{q}(-t)\mid < 1\}\),

$$\begin{aligned} \frac{1}{g^{(3)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }t^{n} \sum _{k=0}^{n} \frac{(-1)^{k}\,c_{k}}{2^{k}[n-k]_{q}!}\beta ^{(3)}_{n-k,\alpha }(q), \end{aligned}$$
(2.24)

where \(c_{n}\) is defined in (2.23).

Proof

Substitute with \( x=0\) in Eq. (2.3). This gives

$$\begin{aligned} \frac{\exp _{q}(\frac{-t}{2})}{g^{(3)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }\beta ^{(3)}_{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$

From Lemma 2.9,

$$\begin{aligned} \begin{aligned} \frac{1}{g^{(3)}_{\alpha }(it;q)}&=\frac{1}{\exp _{q}(\frac{-t}{2})}\sum _{n=0}^{\infty }\beta ^{(3)}_{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}\\&= \left( \sum _{n=0}^{\infty }c_{n}\frac{(-1)^{n}t^{n}}{2^{n}}\right) \left( \sum _{n=0}^{\infty }\beta ^{(3)}_{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}\right) . \end{aligned} \end{aligned}$$

Applying the Cauchy product formula, we obtain (2.24) and completes the proof. \(\square \)

Theorem 2.11

For \( n\in {\mathbb {N}}_{0} \) and \(x\in {\mathbb {C}}\),

$$\begin{aligned} \sum _{k =0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} B^{(1)}_{k,\alpha } (-x;q) B^{(2)}_{n-k,\alpha } (x;q)= \sum _{k =0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \beta _{k,\alpha }(q) \beta _{n-k,\alpha } (q). \end{aligned}$$

Proof

If we replace x by \( -x \) in (2.1), then

$$\begin{aligned} \frac{e_{q}(-xt)e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(-x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.25)

Since \( e_{q}(-xt)E_{q}(xt)=1,\) then multiplying (2.2) by (2.25) gives

$$\begin{aligned} \frac{E_{q}(\frac{-t}{2})e_{q}(\frac{-t}{2})}{g^{(2)}_{\alpha }(it;q)g^{(1)}_{\alpha }(it;q)}=\left( \sum _{n=0}^{\infty } B^{(1)}_{n,\alpha }(-x;q)\frac{t^{n}}{[n]_{q}!}\right) \left( \sum _{n=0}^{\infty }B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\right) . \end{aligned}$$

From (2.10), we obtain

$$\begin{aligned} \left( \sum _{n=0}^{\infty } \beta _{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}\right) ^{2}=\sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\sum _{k=0}^{n} \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} B^{(1)}_{k,\alpha } (-x;q) B^{(2)}_{n-k,\alpha } (x;q). \end{aligned}$$

Hence

$$\begin{aligned}&\sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\sum _{k=0}^{n} \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \beta _{k,\alpha }(q) \beta _{n-k,\alpha }(q) \nonumber \\&\quad = \sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\sum _{k=0}^{n} \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} B^{(1)}_{k,\alpha } (-x;q) B^{(2)}_{n-k,\alpha } (x;q). \end{aligned}$$
(2.26)

So, equating the nth power of t in (2.26), we obtain the required result. \(\square \)

Proposition 2.12

For \( n\in {\mathbb {N}}_{0},\) \( x\in {\mathbb {C}}\) and \(q\ne 0, \)

$$\begin{aligned} B^{(2)}_{n,\alpha } (x;q)=q^{\frac{n(n-1)}{2}} B^{(1)}_{n,\alpha } \left( x;\frac{1}{q}\right) . \end{aligned}$$
(2.27)

In particular,

$$\begin{aligned} \beta _{n,\alpha } (q)=q^{\frac{n(n-1)}{2}} \beta _{n,\alpha } \left( \frac{1}{q}\right) . \end{aligned}$$
(2.28)

Proof

Replacing q by \(\frac{1}{q}\) on the generating function in (2.1) and using \( E_{q}(x)=e_{\frac{1}{q}}(x)\), we obtain

$$\begin{aligned} \frac{E_{q}(xt)E_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;\frac{1}{q})}=\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }\left( x;\frac{1}{q}\right) \frac{t^{n}}{[n]_{\frac{1}{q}}!}. \end{aligned}$$
(2.29)

Since

$$\begin{aligned} \begin{aligned} g_{\alpha }^{(1)}\left( it;\frac{1}{q}\right)&=\sum _{n=0}^{\infty }\frac{(1-q^{-1})^{2n}(\frac{t}{2})^{2n}}{(q^{-2};q^{-2})_{n}(q^{-2\alpha -2};q^{-2})_{n}}\\&=\sum _{n=0}^{\infty }\frac{(1-q)^{2n}q^{2n(n+\alpha )}(\frac{t}{2})^{2n}}{(q^{2},q^{2\alpha +2};q^{2})_{n}}=g_{\alpha }^{(2)}(it;q), \end{aligned} \end{aligned}$$

where we used the identity \((a;q^{-1})_{n}=(a^{-1};q)_{n}(-a)^{n}q^{-\frac{n(n-1)}{2}}\). Since \( [n]_{1/q}!=q^{\frac{n(1-n)}{2}}[n]_{q}!\), then (2.29) takes the form

$$\begin{aligned} \frac{E_{q}(xt)E_{q}(\frac{-t}{2})}{g^{(2)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }\left( x;\frac{1}{q}\right) q^{\frac{n(n-1)}{2}}\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$

Therefore,

$$\begin{aligned} \sum _{n=0}^{\infty }B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}=\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }\left( x;\frac{1}{q}\right) \frac{q^{\frac{n(n-1)}{2}}t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.30)

Equating the coefficients of \( t^{n}\) in (2.30) gives (2.27) and substituting with \( x=0 \) into (2.27) yields directly (2.28). \(\square \)

Al-Salam in [3] introduced the polynomials

$$\begin{aligned} H_{n}(x):=\sum _{k=0}^{n} \, \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\,x^{k} ,\quad G_{n}(x):=\sum _{k=0}^{n} \, \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\, q^{k^{2}-nk} x^{k}. \end{aligned}$$
(2.31)

He also proved that

$$\begin{aligned}&E_{q}(x) E_{q}(-x)=\sum _{n=0}^{\infty } q^{\frac{n(n-1)}{2}} G_{n}(-1)\frac{x^{n}}{[n]_{q}!} ,\quad x \in \,{\mathbb {C}}, \end{aligned}$$
(2.32)
$$\begin{aligned}&e_{q}(x) e_{q}(-x)=\sum _{n=0}^{\infty } H_{n}(-1)\frac{x^{n}}{[n]_{q}!} ,\quad \mid x \mid \,<\,\frac{1}{1-q}. \end{aligned}$$
(2.33)

The following theorem introduces connection relations between the polynomials \(B^{(1)}_{n,\alpha }(x;q)\) and \(B^{(2)}_{n,\alpha }(x;q)\).

Theorem 2.13

For \( n \in {\mathbb {N}}_{0} \),

$$\begin{aligned} B^{(1)}_{n,\alpha }(x;q)&=\sum _{k=0}^{n} \, \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\, x^{k}H_{k}(-1) B^{(2)}_{n-k,\alpha }(x;q), \end{aligned}$$
(2.34)
$$\begin{aligned} B^{(2)}_{n,\alpha }(x;q)&=\sum _{k=0}^{n} \, \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\, q^{\frac{k(k-1)}{2}}x^{k}G_{k}(-1) B^{(1)}_{n-k,\alpha }(x;q). \end{aligned}$$
(2.35)

Proof

Since \(E_{q}(xt) e_{q}(-xt)= 1,\,\mid xt \mid <\frac{1}{1-q},\) then from (2.9), the generating function of \( B^{(1)}_{n,\alpha }(x;q) \) can be represented as

$$\begin{aligned} \begin{aligned} \frac{e_{q}(xt)e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}=\frac{E_{q}(xt) E_{q}(\frac{-t}{2})}{g^{(2)}_{\alpha }(it;q)}e_{q}(xt)e_{q}(-xt). \end{aligned} \end{aligned}$$

From (2.1), (2.2) and (2.33), we obtain

$$\begin{aligned} \begin{aligned} \sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}&= \left( \sum _{n=0}^{\infty }B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\right) \left( \sum _{n=0}^{\infty } H_{n}(-1)\frac{(xt)^{n}}{[n]_{q}!}\right) \\&=\sum _{n=0}^{\infty } \frac{t^{n}}{[n]_{q}!} \sum _{k=0}^{n} \, \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\, x^{k}H_{k}(-1) B^{(2)}_{n-k,\alpha }(x;q).\end{aligned}\nonumber \\ \end{aligned}$$
(2.36)

Therefore, equating the coefficients of the nth power of t in the series of the outside parts of (2.36) gives (2.34). The proof for \(B^{(2)}_{n,\alpha }(x;q)\) follows similarly from the generating function of \( B^{(2)}_{n,\alpha }(x;q) \) and the identity (2.32), and is omitted. \(\square \)

Theorem 2.14

Let x and \(\alpha \) be two complex numbers, with \(Re\,\alpha >-1 \), and n a positive integer. Then

$$\begin{aligned} \begin{aligned} \sum _{k = 0}^{[\frac{n}{2}]} \, \frac{ (1-q)^{2k} \,B^{(1)}_{n-2k,\alpha }(-\frac{x}{2};q)}{2^{2k}\,[n-2k]_{q}!\, (q^{2},q^{2n+\alpha };q^{2})_{k}}&=\frac{(-1/2)^{n}}{[n]_{q}!} H_{n}(x),\\ \sum _{k =0 }^{[\frac{n}{2}]} \, \frac{ (1-q)^{2k} q^{2k(k+\alpha )}\,B^{(2)}_{n-2k,\alpha }(-\frac{x}{2};q)}{2^{2k}\,[n-2k]_{q}!\, (q^{2},q^{2n+\alpha };q^{2})_{k}}&=\frac{(-1/2)^{n}}{[n]_{q}!}q^{\frac{n(n-1)}{2}} G_{n}(x),\\ \sum _{k = 0}^{[\frac{n}{2}]} \, \frac{ (1-q)^{2k} q^{k^{2}+k/2}\,B^{(3)}_{n-2k,\alpha }(-\frac{x}{2};q)}{2^{2k}\,[n-2k]_{q}!\, (q^{2},q^{2\alpha +2};q^{2})_{k}}&=\frac{(-1/2)^{n}}{[n]_{q}!}q^{\frac{n(n-1)}{4}} (-xq^{\frac{1-n}{2}};q)_{n} . \end{aligned} \end{aligned}$$
(2.37)

Proof

We can write the generating function of the polynomials \(B^{(1)}_{n,\alpha }(x;q)\) as

$$\begin{aligned} \begin{aligned} e_{q}(xt)e_{q}\left( \frac{-t}{2}\right)&= g^{(1)}_{\alpha }(it;q)\sum _{n=0}^{\infty } B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\\&=\left( \sum _{n=0}^{\infty } \frac{(1-q)^{2n}t^{2n}}{2^{2n}(q^{2},q^{2\alpha +2};q^{2})_{n}}\right) \,\left( \sum _{n=0}^{\infty } B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\right) .\end{aligned}\nonumber \\ \end{aligned}$$
(2.38)

On one hand, applying the Cauchy product formula in (2.38), we obtain

$$\begin{aligned} e_{q}(xt)e_{q}\left( \frac{-t}{2}\right) =\sum _{n=0}^{\infty }t^{n}\sum _{k=0}^{[\frac{n}{2}]}\frac{(1-q)^{2k}B^{(1)}_{n-2k,\alpha }(x;q)}{2^{2k}[n-2k]_{q}!(q^{2},q^{2\alpha +2};q^{2})_{k}} . \end{aligned}$$

On the other hand, using the series representation of \(e_{q}(x)\) in (1.4) followed by the Cauchy product formula, and using (2.31) yields

$$\begin{aligned} e_{q}(xt)e_{q}\left( \frac{-t}{2}\right) =\sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\,\left( \frac{-1}{2}\right) ^{n} H_{n}(-2x). \end{aligned}$$
(2.39)

Hence

$$\begin{aligned} \sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\,\left( \frac{-1}{2}\right) ^{n} H_{n}(-2x)=\sum _{n=0}^{\infty }t^{n}\sum _{k=0}^{[\frac{n}{2}]}\frac{(1-q)^{2k}B^{(1)}_{n-2k,\alpha }(x;q)}{2^{2k}[n-2k]_{q}!(q^{2},q^{2\alpha +2};q^{2})_{k}},\nonumber \\ \end{aligned}$$
(2.40)

equating the coefficients of \(t^{n}\) in (2.40), we get

$$\begin{aligned} \frac{(\frac{-1}{2})^{n}\,H_{n}(-2x)}{[n]_{q}!}=\sum _{k=0}^{[\frac{n}{2}]}\frac{(1-q)^{2k}B^{(1)}_{n-2k,\alpha }(x;q)}{2^{2k}[n-2k]_{q}!(q^{2},q^{2\alpha +2};q^{2})_{k}}. \end{aligned}$$
(2.41)

Replacing \(x \,by\,\, \frac{-x}{2}\) in (2.41) gives

$$\begin{aligned} \frac{(\frac{-1}{2})^{n}\,H_{n}(x)}{[n]_{q}!}=\sum _{k=0}^{[\frac{n}{2}]}\frac{(1-q)^{2k}B^{(1)}_{n-2k,\alpha }(\frac{-x}{2};q)}{2^{2k}[n-2k]_{q}!(q^{2},q^{2\alpha +2};q^{2})_{k}}, \end{aligned}$$

which readily completes the proof for \(B^{(1)}_{n,\alpha }(x;q)\). The proofs for \(B^{(2)}_{n,\alpha }(x;q)\) and \(B^{(3)}_{n,\alpha }(x;q)\) are similar and are omitted. \(\square \)

If we set \(x=0\) in (2.37), we obtain the following recurrence relations for \(\beta _{n,\alpha }(q)\) and \(\beta ^{(3)}_{n,\alpha }(q),\)

$$\begin{aligned} \begin{aligned} \sum _{k = 0}^{[\frac{n}{2}]} \, \frac{ (1-q)^{2k} \,\beta _{n-2k,\alpha }(q)}{2^{2k}\,[n-2k]_{q}!\, (q^{2},q^{2\alpha +2};q^{2})_{k}}\,&=\frac{(\frac{-1}{2})^{n}}{[n]_{q}!}\quad (n\in {\mathbb {N}}),\\ \quad \sum _{k = 0}^{[\frac{n}{2}]} \, \frac{(1-q)^{2k} q^{k^{2}+k/2}\,\beta ^{(3)}_{n-2k,\alpha }(q)}{2^{2k}\,[n-2k]_{q}!\, (q^{2},q^{2\alpha +2};q^{2})_{k}}\,&=\frac{q^{\frac{n(n-1)}{4}}(\frac{-1}{2})^{n}}{[n]_{q}!}\quad (n\in {\mathbb {N}}). \end{aligned} \end{aligned}$$
(2.42)

As a consequence of the recursive relations in (2.42), and the fact that

$$\begin{aligned} \beta _{0,\alpha }(q)=\beta ^{(3)}_{0,\alpha }(q)=1, \end{aligned}$$

we can prove that

$$\begin{aligned} \beta _{1,\alpha }(q)&=-\frac{1}{2},\quad \beta _{2,\alpha }(q)=\frac{q(1-q^{2\alpha +1})}{4(1-q^{2\alpha +2})},\quad \beta _{3,\alpha }(q)=\frac{-q^{3}(1-q^{2\alpha -1})}{8(1-q^{2\alpha +2})},\\ \beta _{4,\alpha }(q)&=\frac{1}{16}-\frac{(q+q^{3})(1-q^{3})(1-q^{2\alpha +1})}{16(1-q^{2\alpha +2})^{2}}-\frac{(1-q)(1-q^{3})}{16(q^{2\alpha +2};q^{2})_{2}},\\ \beta _{5,\alpha }(q)&=\frac{(1+q^{2})(1-q^{5})(q^{3}-q^{2\alpha +2})}{32(1-q^{2\alpha +2})^{2}}+\frac{(1-q^{3})(1-q^{5})}{32(q^{2\alpha +2};q^{2})_{2}} -\frac{1}{32}, \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} \beta ^{(3)}_{1,\alpha }(q)&=\frac{-1}{2},\quad \quad \beta ^{(3)}_{2,\alpha }(q)= \frac{q^{1/2}(1-q^{2\alpha +2})-q^{3/2}(1-q)}{4(1-q^{2\alpha +2})},\\ \beta ^{(3)}_{3,\alpha }(q)&=\frac{-q^{3/2}(q^{3}-q^{2\alpha +2})}{8(1-q^{2\alpha +2})}, \\ \beta ^{(3)}_{4,\alpha }(q)&=\frac{q^{3}(q^{2\alpha +2};q^{2})_{2}(1-q^{2\alpha +2}) }{16(1-q^{2\alpha +2})^{2}(1-q^{2\alpha +4})}-\frac{[3]_{q}q^{5}(1-q)^{2}(1-q^{2\alpha +2})}{16(1-q^{2\alpha +2})^{2}(1-q^{2\alpha +4})}\\&\quad +\,\frac{[4]_{q}[3]_{q}q^{3/2} (1-q^{2\alpha +4})\left( q^{1/2}(1-q^{2\alpha +2})-q^{3/2}(1-q)\right) }{16(1-q^{2\alpha +2})^{2}(1-q^{2\alpha +4})}, \\ \quad \beta ^{(3)}_{5,\alpha }(q)&=\frac{[5]_{q}q^{3}(1-q)(1+q^{2})(q^{3}-q^{2\alpha +2})(1-q^{2\alpha +4})}{32(1-q^{2\alpha +2})^{2}(1-q^{2\alpha +4})} \\&\quad +\,\frac{[5]_{q}q^{5}(1-q)(1-q^{3})(1-q^{2\alpha +2})}{32(1-q^{2\alpha +2})^{2}(1-q^{2\alpha +4})}-\frac{q^{5}(1-q^{2\alpha +2})^{2}(1-q^{2\alpha +4})}{32(1-q^{2\alpha +2})^{2}(1-q^{2\alpha +4})}.\end{aligned} \end{aligned}$$

Theorem 2.15

For \( n\in \,{\mathbb {N}}_{0} \) and complex numbers a and x,

$$\begin{aligned} B^{(1)}_{n,\alpha }(x;q)&=\sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\,(a;q)_{k} \,x^{k} B^{(1)}_{n-k,\alpha }(ax;q), \end{aligned}$$
(2.43)
$$\begin{aligned} B^{(2)}_{n,\alpha }(x;q)&=\sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\,(-a)^{k}(1/a;q)_{k} \,x^{k} B^{(2)}_{n-k,\alpha }(ax;q). \end{aligned}$$
(2.44)

Proof

The proof of (2.43) follows from the generating function (2.1) since

$$\begin{aligned} \begin{aligned} \frac{e_{q}(xt)e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}=\frac{e_{q}(tax)e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}\,\frac{e_{q}(tx)}{e_{q}(atx)},\quad \mid tx \mid <\frac{1}{1-q}. \end{aligned} \end{aligned}$$

From the q-binomial theorem (see [12, Eq.(1.3.2), p. 8]), we can prove that

$$\begin{aligned} \frac{e_{q}(tx)}{e_{q}(atx)}=\sum _{n=0}^{\infty }\frac{(a;q)_{n}}{(q;q)_{n}}((1-q)tx)^{n},\quad \mid tx \mid <\frac{1}{1-q}. \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned} \sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}&=\left( \sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(ax;q)\frac{t^{n}}{[n]_{q}!}\right) \,\, \left( \sum _{n=0}^{\infty }\frac{(a;q)_{n}}{[n]_{q}!}(tx)^{n}\right) \\&=\sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\, (a;q)_{k}x^{k} B^{(1)}_{n-k,\alpha }(ax;q), \end{aligned}\nonumber \\ \end{aligned}$$
(2.45)

where we used the Cauchy product formula. Equating the coefficients of \( t^{n} \) in (2.45), we obtain (2.43). The proof for \(B^{(2)}_{n,\alpha }(x;q)\) is similar and is omitted. \(\square \)

Lemma 2.16

For \(n\in {\mathbb {N}}_{0}\), \(Re \,\alpha >-1\), and \(\mid \frac{(1-q)t}{2}\mid <1\),

$$\begin{aligned} g^{(1)}_{\alpha }(it;q)E_{q}(\frac{t}{2})= g^{(2)}_{\alpha }(it;q)e_{q}(\frac{t}{2}) ={}_{2}\phi _{1}\,(q^{\alpha +\frac{1}{2}},-q^{\alpha +\frac{1}{2}};q^{2\alpha +1};q,\frac{(1-q)t}{2}).\nonumber \\ \end{aligned}$$
(2.46)

Proof

From Lemma 2.2, we conclude that

$$\begin{aligned} \begin{aligned} g^{(2)}_{\alpha }(it;q) e_{q}\left( \frac{t}{2}\right) =g^{(1)}_{\alpha }(it;q)E_{q}\left( \frac{t}{2}\right) .\end{aligned} \end{aligned}$$

From the series representations of \(E_{q}(x)\) and \(g^{(1)}_{\alpha }(it;q)\) in (1.4) and (2.38), respectively we obtain

$$\begin{aligned} \begin{aligned} g^{(1)}_{\alpha }(it;q)E_{q}(\frac{t}{2})&=\left( \sum _{n=0}^{\infty }\frac{(1-q)^{2n}t^{2n}}{2^{2n}(q^{2},q^{2\alpha +2};q^{2})_{n}}\right) \left( \sum _{n=0}^{\infty }\frac{q^{\frac{n(n-1)}{2}}(1-q)^{n}t^{n}}{2^{n}(q;q)_{n}}\right) \\&=\sum _{n=0}^{\infty }\frac{(1-q)^{n}q^{\frac{n(n-1)}{2}}t^{n}}{2^{n}} \sum _{k=0}^{[\frac{n}{2}]}\frac{q^{2k^{2}-2nk +k} }{(q;q)_{n-2k}(q^{2},q^{2\alpha +2};q^{2})_{k}}\\&=\sum _{n=0}^{\infty }\frac{q^{\frac{n(n-1)}{2}}(1-q)^{n}t^{n}}{2^{n}(q;q)_{n}}\sum _{k=0}^{[\frac{n}{2}]} \frac{q^{2k}(q^{-n};q)_{2k} }{(q^{2},q^{2\alpha +2};q^{2})_{k}},\end{aligned} \end{aligned}$$

where we used the identity, see [12, Eq. (1.2.32), p. 6],

$$\begin{aligned} \begin{aligned} (a;q)_{n-k}=\frac{(a;q)_{n}}{(a^{-1}q^{1-n};q)_{k}}(-qa^{-1})^{k}q^{\frac{k(k-1)}{2} -nk}\quad (k=0,1,\ldots ,n). \end{aligned} \end{aligned}$$
(2.47)

Therefore, using the identity \((a;q)_{2n}=(a;q^{2})_{n}(aq;q^{2})_{n}\) yields

$$\begin{aligned} \begin{aligned} g^{(1)}_{\alpha }(it;q)E_{q}\left( \frac{t}{2}\right)&=\sum _{n=0}^{\infty }\frac{q^{\frac{n(n-1)}{2}}(1-q)^{n}(\frac{t}{2})^{n}}{(q;q)_{n}}\sum _{k=0}^{[\frac{n}{2}]} \frac{q^{2k}(q^{-n};q^{2})_{k}(q^{-n+1};q^{2})_{k}}{(q^{2};q^{2\alpha +2};q^{2})_{k}}\\&= \sum _{n=0}^{\infty }\dfrac{q^{\frac{n(n-1)}{2}}(1-q)^{n}(\frac{t}{2})^{n}}{(q;q)_{n}}\,\,_{2}\phi _{1}(q^{-n},q^{-n+1};q^{2\alpha +2};q^{2},q^{2}). \end{aligned} \end{aligned}$$

Since

$$\begin{aligned} _{2}\phi _{1}\,(q^{-n},q^{1-n};qb^{2};q^{2},q^{2})=\frac{(b^{2};q^{2})_{n}}{(b^{2};q)_{n}}q^{\frac{-n(n-1)}{2}}\,\,(n\in {\mathbb {N}}), \end{aligned}$$

see [12, p. 26], then

$$\begin{aligned} \begin{aligned} g^{(1)}_{\alpha }(it;q)E_{q}\left( \frac{t}{2}\right)&=\sum _{n=0}^{\infty }\dfrac{(q^{\alpha +\frac{1}{2}};q)_{n}(-q^{\alpha +\frac{1}{2}};q)_{n}(\frac{(1-q)t}{2})^{n}}{ (q;q)_{n}(q^{2\alpha +1};q)_{n}}\\&={}_{2}\phi _{1}\,\left( q^{\alpha +\frac{1}{2}},-q^{\alpha +\frac{1}{2}};q^{2\alpha +1};q,\frac{(1-q)t}{2}\right) . \end{aligned}\nonumber \\ \end{aligned}$$
(2.48)

Hence from Lemma 2.2 and (2.48), we obtain (2.46) and completes the proof. \(\square \)

Theorem 2.17

Let \( \alpha \) be a complex number such that \( Re\,\alpha >-1\). Then

$$\begin{aligned} \sum _{m=0}^{n}\left[ \begin{array}{c} n \\ m \\ \end{array} \right] _{q}\dfrac{(q^{2\alpha +1};q^{2})_{m}B^{(1)}_{n-m,\alpha }(x;q)}{2^{m} (q^{2\alpha +1};q)_{m}}&= x^{n},\\ \sum _{m=0}^{n}\left[ \begin{array}{c} n \\ m \\ \end{array} \right] _{q}\dfrac{(q^{2\alpha +1};q^{2})_{m}B^{(2)}_{n-m,\alpha }(x;q)}{2^{m} (q^{2\alpha +1};q)_{m}}&=q^{\frac{n(n-1)}{2}}x^{n},\\ \sum _{m=0}^{n}(-\frac{1}{2})^{m}\left( \sum _{k = 0}^{[\frac{m}{2}]}\, \frac{ \,q^{k^{2}+k/2}\,(1-q)^{2k}\,c_{m-2k}}{\,(q^{2},q^{2\alpha +2};q^{2})_{k}}\right) \frac{B^{(3)}_{n-m,\alpha }(x;q)}{[n-m]_{q}!}&= \frac{q^{\frac{n(n-1)}{4}}x^{n}}{[n]_{q}!}, \end{aligned}$$

where \((c_{k})_{k}\) are the coefficients defined in (2.23).

Proof

We can write Eq. (2.1) in the form

$$\begin{aligned} \begin{aligned} e_{q}(xt)&=E_{q}\left( \frac{t}{2}\right) g^{(1)}_{\alpha }(it;q)\sum _{n=0}^{\infty } B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\\&= \left( \sum _{n=0}^{\infty } d_{n} t^{n}\right) \,\left( \sum _{n=0}^{\infty } B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\right) .\end{aligned}\nonumber \\ \end{aligned}$$
(2.49)

From Lemma 2.16, we obtain

$$\begin{aligned} \begin{aligned} g^{(1)}_{\alpha }(it;q)E_{q}\left( \frac{t}{2}\right) =\sum _{n=0}^{\infty } d_{n}t^{n}, \end{aligned} \end{aligned}$$
(2.50)

where

$$\begin{aligned} d_{n}=\dfrac{(1-q)^{n}(q^{2\alpha +1};q^{2})_{n}}{2^{n}(q;q)_{n} (q^{2\alpha +1};q)_{n}} . \end{aligned}$$
(2.51)

Now, applying the Cauchy product formula in (2.49) gives

$$\begin{aligned} \begin{aligned} e_{q}(xt)=\sum _{n=0}^{\infty }t^{n}\,\sum _{m=0}^{n}\frac{d_{m} B^{(1)}_{n-m,\alpha }(x;q) }{ [n-m]_{q}!}= \sum _{n=0}^{\infty }\frac{(xt)^{n}}{[n]_{q}!}. \end{aligned} \end{aligned}$$
(2.52)

Equating the coefficients of the nth power of t in (2.52) gives

$$\begin{aligned} \sum _{m=0}^{n}\frac{d_{m} B^{(1)}_{n-m,\alpha }(x;q) }{ [n-m]_{q}!}=\frac{x^{n}}{[n]_{q}!}. \end{aligned}$$
(2.53)

Substituting from (2.51) into (2.53), we get the result for \(B^{(1)}_{n,\alpha }(x;q)\). Similarly, we can prove the result for \( B^{(k)}_{n,\alpha }(x;q)\,(k=2,3) \). \(\square \)

Theorem 2.18

Let n be a positive integer and x be a complex number. If \(Re\,\alpha >-1 \), then

$$\begin{aligned}&B^{(1)}_{n,\alpha }(x;q)-(-1)^{n} B^{(1)}_{n,\alpha }(-x;q)\nonumber \\&\quad =\sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\left( \left( \frac{-1}{2}\right) ^{k}H_{k}(-2x)-\left( \frac{1}{2}\right) ^{k}H_{k}(2x)\right) B^{(2)}_{n-k,\alpha }\left( \frac{1}{2};q\right) ,\nonumber \\ \end{aligned}$$
(2.54)
$$\begin{aligned}&B^{(2)}_{n,\alpha }(x;q)-(-1)^{n} B^{(2)}_{n,\alpha }(-x;q)\nonumber \\&\quad =\sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}q^{\frac{k(k-1)}{2}}\left( (\frac{-1}{2})^{k}G_{k}(-2x)-(\frac{1}{2})^{k}G_{k}(2x)\right) B^{(1)}_{n-k,\alpha }(\frac{1}{2};q).\nonumber \\ \end{aligned}$$
(2.55)

Proof

We give only the proof of (2.54) since the proof of (2.55) is similar. From (2.1),

$$\begin{aligned} \frac{e_{q}(xt)e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}-\frac{e_{q}(xt)e_{q}(\frac{t}{2})}{g^{(1)}_{\alpha }(-it;q)} =\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}-\sum _{n=0}^{\infty }B^{(1)}_{n,\alpha }(-x;q)\frac{(-t)^{n}}{[n]_{q}!}.\nonumber \\ \end{aligned}$$
(2.56)

Since

$$\begin{aligned} g^{(1)}_{\alpha }(-it;q)=g^{(1)}_{\alpha }(it;q), \end{aligned}$$

then Eq. (2.56) can be written as

$$\begin{aligned} \frac{e_{q}(xt)e_{q}(\frac{-t}{2}) -e_{q}(xt)e_{q}(\frac{t}{2})}{g^{(1)}_{\alpha }(it;q)} =\sum _{n=0}^{\infty }\left[ B^{(1)}_{n,\alpha }(x;q)-(-1)^{n}B^{(1)}_{n,\alpha }(-x;q)\right] \frac{t^{n}}{[n]_{q}!}.\nonumber \\ \end{aligned}$$
(2.57)

Replacing xt by \(-x, -t\), respectively in (2.39) gives

$$\begin{aligned} e_{q}(xt)e_{q}(\frac{t}{2})= \sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}(\frac{1}{2})^{n}H_{n}(2x). \end{aligned}$$
(2.58)

From (2.39) and (2.58), the left hand side of (2.57) can be written as

$$\begin{aligned} \begin{aligned} \frac{e_{q}(xt)e_{q}(\frac{-t}{2}) -e_{q}(xt)e_{q}(\frac{t}{2})}{g^{(1)}_{\alpha }(it;q)}&=\frac{1}{g^{(1)}_{\alpha }(it;q)}\sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\\&\quad \times \left( \left( \frac{-1}{2}\right) ^{n}H_{n}(-2x)-\left( \frac{1}{2}\right) ^{n}H_{n}(2x)\right) .\end{aligned} \end{aligned}$$

Therefore, by (2.21) and the Cauchy product formula, we get

$$\begin{aligned} \begin{aligned}&\frac{e_{q}(xt)e_{q}(\frac{-t}{2}) -e_{q}(xt)e_{q}(\frac{t}{2})}{g^{(1)}_{\alpha }(it;q)}\\&=\sum _{n=0}^{\infty }\frac{t^{n}}{[n]_{q}!}\sum _{k=0}^{n} \left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q}\left( \left( \frac{-1}{2}\right) ^{k}H_{k}(-2x)-\left( \frac{1}{2}\right) ^{k}H_{k}(2x)\right) B^{(2)}_{n-k,\alpha }\left( \frac{1}{2};q\right) . \end{aligned}\nonumber \\ \end{aligned}$$
(2.59)

Since the left hand side of (2.57) and (2.59) are equal, then equating the coefficients of \(t^{n}\) on the right hand sides of (2.57) and (2.59) yields (2.54) and completes the proof. \(\square \)

Proposition 2.19

If \(\alpha _{0}>-1\) satisfies the condition

$$\begin{aligned} q^{2(\alpha _{0}+1)}(1-q)^{2}<(1-q^{2})(1-q^{2\alpha _{0}+2}), \end{aligned}$$
(2.60)

then \((t/2)^{-\alpha }J_{\alpha }^{(2)}(t(1-q);q^{2})\) has no zeros in \(\mid t\mid \le 1\) for all \(\alpha \ge \alpha _{0}.\)

Proof

Set

$$\begin{aligned} F(t):= \frac{(q;q)_{\infty }}{(q^{\alpha +1};q)_{\infty }}(t/2)^{-\alpha }J_{\alpha }^{(2)}(t(1-q);q^{2}) =\sum _{k=0}^{\infty }\frac{(-1)^{k}q^{2k(k+\alpha )}(1-q)^{2k}}{2^{2k}(q^{2},q^{2\alpha +2};q^{2})_{k}}t^{2k}, \end{aligned}$$

and

$$\begin{aligned} a_{k}:=\frac{q^{2k(k+\alpha )}(1-q)^{2k}}{2^{2k}(q^{2},q^{2\alpha +2};q^{2})_{k}}. \end{aligned}$$

Then, under hypothesis (2.60) and since \(0< q < 1\),

$$\begin{aligned} q^{2(\alpha +1)}(1-q^{2})\le & {} q^{2(\alpha _{0}+1)}(1-q^{2})<(1-q^{2})(1-q^{2\alpha _{0}+2})\\ {}\le & {} (1-q^{2})(1-q^{2\alpha +2}), \end{aligned}$$

holds whenever \(\alpha \ge \alpha _{0}\). Hence

$$\begin{aligned} \frac{a_{k+1}}{a_{k}}=\frac{q^{4k+2(\alpha +1)}(1-q)^{2}}{4(1-q^{2k+2})(1-q^{2k+2\alpha +2})} \le \frac{q^{2(\alpha +1)}(1-q)^{2}}{4(1-q^{2})(1-q^{2\alpha +2})}<1, \end{aligned}$$

for \(t\in {\mathbb {R}},\, \mid t\mid \le 1\)

$$\begin{aligned} \begin{aligned} F(t)=\sum _{k=0}^{\infty }t^{2k}(a_{2k}-a_{2k+1}t^{2})\ge (a_{0}-a_{1}t^{2})\ge (a_{0}-a_{1})>0. \end{aligned} \end{aligned}$$

This proves that F(t) has no zeros on \([-1,1]\), since F(t) has only real zeros, then F(t) has no zeros in the unit disk. i.e \(\mid F(t)\mid >0,\,\,for\,\,\mid t\mid \le 1.\) \(\square \)

Corollary 2.20

There exists \(\alpha _{0}>-1\) such that \(J_{\alpha }^{(2)}(t(1-q);q^{2})\) has no zeros in the unit disk for all \(\alpha \ge \alpha _{0}\).

Proof

Since for a fixed \(q\in (0,1)\),

$$\begin{aligned} \displaystyle \lim _{\alpha \rightarrow \infty }q^{2\alpha +2}=0,\,\,\displaystyle \lim _{\alpha \rightarrow \infty }(1-q^{2})(1-q^{2\alpha +2})=(1-q^{2}), \end{aligned}$$

then there exists \(\alpha _{0}>-1 \) such that the condition (2.60) holds for all \(\alpha \ge \alpha _{0}\). Consequently from Proposition 2.19, \(J_{\alpha }^{(2)}(t(1-q);q^{2})\) has no zeros in the unit disk for all \(\alpha \ge \alpha _{0}\). \(\square \)

Theorem 2.21

For \(n\in {\mathbb {N}}\),

$$\begin{aligned}&\lim _{\alpha \rightarrow \infty }\,B^{(2)}_{n,\alpha }(x;q)=\left( -\frac{1}{2}\right) ^{n}q^{\frac{n(n-1)}{2}}G_{n}(-2x), \end{aligned}$$
(2.61)
$$\begin{aligned}&\lim _{\alpha \rightarrow \infty }\,B^{(1)}_{n,\alpha }(x;q)=x^{n}\left( \frac{1}{2x};q\right) _{n}. \end{aligned}$$
(2.62)

Proof

Taking the limit on both sides of Eq. (2.2) as \( \alpha \rightarrow \infty \) we get

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } \frac{E_{q}(xt)E_{q}(\frac{-t}{2})}{ g^{(2)}_{\alpha }(it;q)}= \lim _{\alpha \rightarrow \infty }\,\sum _{n=0}^{\infty } B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.63)

From Corollary 2.20, there exists \(\alpha _{0}>-1\) such that \(g^{(2)}_{\alpha }(it;q)\) has no zeros in \(\mid t\mid \le 1\) for all \(\alpha \ge \alpha _{0}\). This means that \(\dfrac{E_{q}(xt)E_{q}(\frac{-t}{2})}{ g^{(2)}_{\alpha }(it;q)}\) is analytic in \(\mid t\mid \le 1\) for all \(\alpha \ge \alpha _{0}\). Therefore, we can interchange the limit with the summation in (2.63) when \(\mid t\mid \le 1\) to obtain

$$\begin{aligned} \frac{E_{q}(xt)E_{q}(\frac{-t}{2})}{\displaystyle \lim _{\alpha \rightarrow \infty } g^{(2)}_{\alpha }(it;q)}= \sum _{n=0}^{\infty } \lim _{\alpha \rightarrow \infty }\, B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$

Since

$$\begin{aligned} \lim _{\alpha \rightarrow \infty } g^{(2)}_{\alpha }(it;q)=1,\quad E_{q}(xt)E_{q}(yt)=\sum _{n=0}^{\infty }\frac{q^{n(n-1)/2}(ty)^{n}}{[n]_{q}!}G_{n}(\frac{x}{y}), \end{aligned}$$

then from (2.32)

$$\begin{aligned} \begin{aligned} \sum _{n=0}^{\infty }\lim _{\alpha \rightarrow \infty } B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!} =\sum _{n=0}^{\infty }\frac{q^{n(n-1)/2}t^{n}}{[n]_{q}!}\left( \frac{-1}{2}\right) ^{n}G_{n}(-2x).\end{aligned} \end{aligned}$$
(2.64)

Equating the coefficients of \( t^{n} \) in (2.64) gives (2.61). The proof of (2.62) follows directly from the relation (2.9) since

$$\begin{aligned} 1=\displaystyle \lim _{\alpha \rightarrow \infty }g^{(2)}_{\alpha }(it;q)= E_{q}\left( \frac{t}{2}\right) E_{q}\left( \frac{-t}{2}\right) \displaystyle \lim _{\alpha \rightarrow \infty }g^{(1)}_{\alpha }(it;q). \end{aligned}$$

Hence

$$\begin{aligned} \displaystyle \lim _{\alpha \rightarrow \infty }g^{(1)}_{\alpha }(it;q)=e_{q}\left( \frac{t}{2}\right) e_{q}\left( \frac{-t}{2}\right) ,\,\,\mid t(1-q)\mid <2. \end{aligned}$$

Therefore, computing the limit in both sides of (2.1) gives

$$\begin{aligned} \frac{e_{q}(xt)}{e_{q}(\frac{t}{2})}= \sum _{n=0}^{\infty } \lim _{\alpha \rightarrow \infty }\, B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$

From the q-binomial theorem (see [12, Eq.(1.3.2), p. 8]), we have

$$\begin{aligned} \frac{e_{q}(xt)}{e_{q}(\frac{t}{2})}= \frac{(\frac{t}{2}(1-q);q)_{\infty }}{(xt(1-q);q)_{\infty }}=\sum _{n=0}^{\infty } \frac{(\frac{1}{2x};q)_{n}}{(q;q)_{n}}(xt(1-q))^{n},\,\,\mid xt(1-q)\mid <1. \end{aligned}$$

Hence

$$\begin{aligned} \sum _{n=0}^{\infty } \lim _{\alpha \rightarrow \infty }\, B^{(1)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}= \sum _{n=0}^{\infty }\frac{(xt)^{n}}{[n]_{q}!}(\frac{1}{2x};q)_{n}, \end{aligned}$$
(2.65)

equating the coefficients of \(t^{n}\) in (2.65) yields the required result. \(\square \)

Corollary 2.22

For \( n\in {\mathbb {N}}\),

$$\begin{aligned} \lim _{\alpha \rightarrow \infty }\,\beta _{n,\alpha }(q)=(-1)^{n}2^{-n}q^{\frac{n(n-1)}{2}}. \end{aligned}$$
(2.66)

Proof

Since

$$\begin{aligned} \lim _{x\rightarrow 0}x^{n}\left( \frac{1}{2x};q\right) _{n}= & {} \lim _{x\rightarrow 0}x^{n}\prod _{k=0}^{n-1}\left( 1-\frac{q^{k}}{2x}\right) =\lim _{x\rightarrow 0}\prod _{k=0}^{n-1}\left( x-\frac{q^{k}}{2}\right) \\ {}= & {} (-1)^{n}2^{-n}q^{\frac{n(n-1)}{2}}, \end{aligned}$$

then substituting with \(x=0\) into (2.62) yields (2.66). \(\square \)

Lemma 2.23

Let \(\alpha _{0}>-1\). If \(q^{3/2}(1-q)^{2}<(1-q^{2})(1-q^{2\alpha _{0}+2})\), then \((q^{\frac{1}{4}}t/2)^{-\alpha }J_{\alpha }^{(3)}(\frac{t}{2}(1-q)q^{\frac{-1}{4}};q^{2})\) has no zeros in \(\mid t\mid \le 1\) for all \(\alpha \ge \alpha _{0}.\)

Proof

The proof is similar to the proof of Proposition 2.19 and is omitted. \(\square \)

Theorem 2.24

For \( n\in {\mathbb {N}}\),

$$\begin{aligned} \lim _{\alpha \rightarrow \infty }\!\!B^{(3)}_{n,\alpha }\!(x;q)\!= & {} \!{} q^{\frac{n(n-1)}{4}}\!\left( \frac{-1}{2}\right) ^{n}\!\sum _{k=0}^{[\frac{n}{2}]}\frac{(-1)^{k}q^{k(n-k+3)} (q^{-n};q)_{2k}}{(q^{2};q^{2})_{k}}\! (2xq^{\frac{1-n}{2}};q)_{n-2k},\nonumber \\ \end{aligned}$$
(2.67)
$$\begin{aligned} \lim _{\alpha \rightarrow \infty }\,\beta ^{(3)}_{n,\alpha }(q)= & {} q^{\frac{n(n-1)}{4}}\left( \frac{-1}{2}\right) ^{n}\sum _{k=0}^{[\frac{n}{2}]}\frac{(-1)^{k} q^{k(n-k+3)}(q^{-n};q)_{2k}}{(q^{2};q^{2})_{k}}. \nonumber \\ \end{aligned}$$
(2.68)

Proof

Taking the limit as \(\alpha \rightarrow \infty \) on both sides of (2.3), we obtain

$$\begin{aligned} \displaystyle \lim _{\alpha \rightarrow \infty } \frac{\exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{ g^{(3)}_{\alpha }(it;q)}= \lim _{\alpha \rightarrow \infty }\,\sum _{n=0}^{\infty } B^{(3)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.69)

We can choose \(\alpha _{0}>-1\) such that

$$\begin{aligned} q^{3/2}\le \frac{(1-q^{2})}{1-q}\frac{(1-q^{2\alpha _{0}+2})}{1-q}\le \frac{(1-q^{2})}{1-q}\frac{(1-q^{2\alpha +2})}{1-q}, \end{aligned}$$

for all \(\alpha \ge \alpha _{0}\). Hence from Lemma 2.23, the function \(g^{(3)}_{\alpha }(it;q)\) does not vanish on the unit disk, and the left hand side of (2.69) is analytic for \(\mid t\mid \le 1\). Therefore, we can interchange the limit as \(\alpha \rightarrow \infty \) with the summation in (2.69) to obtain

$$\begin{aligned} \frac{\exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{\displaystyle \lim _{\alpha \rightarrow \infty } g^{(3)}_{\alpha }(it;q)}= \sum _{n=0}^{\infty } \lim _{\alpha \rightarrow \infty }\, B^{(3)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$

Since

$$\begin{aligned} \begin{aligned} \lim _{\alpha \rightarrow \infty } g^{(3)}_{\alpha }(it;q)&= \sum _{n=0}^{\infty }\lim _{\alpha \rightarrow \infty }\frac{q^{n^{2}+\frac{n}{2}}(1-q)^{2n}}{(q^{2},q^{2\alpha +2};q^{2})_{n}}(\frac{t}{2})^{2n} \\&=\sum _{n=0}^{\infty }\frac{q^{n^{2}+\frac{n}{2}}(1-q)^{2n}(\frac{t}{2})^{2n}}{(q^{2};q^{2})_{n}} =\left( -\frac{q^{\frac{3}{2}}(1-q)^{2}t^{2}}{4};q^{2}\right) _{\infty }.\end{aligned} \end{aligned}$$

Hence

$$\begin{aligned} \frac{ \exp _{q}(xt)\exp _{q}(\frac{-t}{2})}{\left( -\frac{q^{\frac{3}{2}}(1-q)^{2}t^{2}}{4};q^{2}\right) _{\infty }}= \sum _{n=0}^{\infty } \lim _{\alpha \rightarrow \infty }\, B^{(3)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.70)

But

$$\begin{aligned} \begin{aligned} \exp _{q}(xt)\exp _{q}\left( \frac{-t}{2}\right)&=\sum _{n=0}^{\infty }\frac{(\frac{-1}{2})^{n}q^{\frac{n(n-1)}{4}}t^{n}}{[n]_{q}!} (2xq^{\frac{1-n}{2}};q)_{n}. \end{aligned} \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned}&\left( \frac{1}{(- q^{\frac{3}{2}}(1-q)^{2}t^{2}/4;q^{2})_{\infty }}\right) \left( \exp _{q}(xt)\exp _{q}(\frac{-t}{2})\right) \\&=\left( \sum _{n=0}^{\infty }\frac{(-1)^{n}q^{\frac{3}{2}n} (\frac{(1-q)t}{2})^{2n}}{(q^{2};q^{2})_{n}}\right) \left( \sum _{n=0}^{\infty }\frac{q^{\frac{n(n-1)}{4}}(\frac{-t}{2})^{n}}{[n]_{q}!} (2xq^{\frac{1-n}{2}};q)_{n}\right) \\&=\sum _{n=0}^{\infty }q^{\frac{n(n-1)}{4}}\left( \frac{-t(1-q)}{2}\right) ^{n}\sum _{k=0}^{[\frac{n}{2}]}\frac{(-1)^{k}q^{\frac{3}{2}k}q^{k^{2}-nk+k/2}}{(q^{2};q^{2})_{k}(q;q)_{n-2k}} (2xq^{\frac{1-n}{2}};q)_{n-2k}\\&=\sum _{n=0}^{\infty }\frac{q^{\frac{n(n-1)}{4}}(\frac{-t}{2})^{n}}{[n]_{q}!}\sum _{k=0}^{[\frac{n}{2}]}\frac{(-1)^{k} q^{k(n-k+3)}(q^{-n};q)_{2k}}{(q^{2};q^{2})_{k}} (2xq^{\frac{1-n}{2}};q)_{n-2k}. \end{aligned}\nonumber \\ \end{aligned}$$
(2.71)

Substituting from (2.71) into (2.70) and equating the coefficients of \(t^{n} \) yields (2.67). The proof of (2.68) follows directly by setting \(x=0\) in (2.67). \(\square \)

Theorem 2.25

Let \(\alpha \) be a complex number such that \( Re\,\alpha >-1\). Then for \(n\in {\mathbb {N}},\,n\,\ge \,2\),

$$\begin{aligned} \begin{aligned} \beta _{n,\alpha }(q)&=-\frac{[n]_{q}!(1-q)^{2}}{4}\sum _{k=0}^{[\frac{n}{2}]-1}\frac{\left( (1-q)/2\right) ^{2k}\beta _{n-2k-2,\alpha }(q)}{[n-2k-2]_{q}!(q^{2},q^{2\alpha +2};q^{2})_{k+1}} +\frac{(-1)^{n}}{2^{n}},\\ \beta _{n,\alpha }^{(3)}(q)&=-\frac{[n]_{q}!q^{3/2}(1-q)^{2}}{4}\sum _{k=0}^{[\frac{n}{2}]-1}\frac{q^{k^{2}+5k/2}\left( (1-q)/2\right) ^{2k}\beta ^{(3)}_{n-2k-2,\alpha }(q)}{[n-2k-2]_{q}!(q^{2},q^{2\alpha +2};q^{2})_{k+1}}\\ {}&\quad +\frac{(-1)^{n}q^{\frac{n(n-1)}{4}}}{2^{n}}.\end{aligned} \end{aligned}$$
(2.72)

Proof

We give in detail the proof of \(\beta _{n,\alpha }(q)\) in (2.72). The proof for \(\beta _{n,\alpha }^{(3)}(q)\) is similar. Since

$$\begin{aligned} \frac{e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}=\sum _{n=0}^{\infty }\beta _{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}, \end{aligned}$$
(2.73)

then

$$\begin{aligned} \frac{e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}-e_{q} \left( \frac{-t}{2}\right) =\sum _{n=0}^{\infty }\beta _{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}- e_{q} \left( \frac{-t}{2}\right) . \end{aligned}$$

Consequently, from the series representation of \(e_{q}(t)\) in (1.4), we get

$$\begin{aligned} \frac{e_{q}(\frac{-t}{2})}{g^{(1)}_{\alpha }(it;q)}\left( 1-g^{(1)}_{\alpha }(it;q)\right) =\sum _{n=0}^{\infty }\left( \beta _{n,\alpha }(q) -\frac{(-1)^{n}}{2^{n}}\right) \frac{t^{n}}{[n]_{q}!}. \end{aligned}$$
(2.74)

Since

$$\begin{aligned} \begin{aligned} \left( g^{(1)}_{\alpha }(it;q)-1\right) =t^{2} \sum _{m=0}^{\infty }\frac{(1-q)^{2m+2}t^{2m}}{2^{2m+2}(q^{2},q^{2\alpha +2};q^{2})_{m+1}},\end{aligned} \end{aligned}$$
(2.75)

then substituting from (2.75) into (2.74) and using (2.73), we obtain

$$\begin{aligned} \begin{aligned}&\left( \sum _{n=0}^{\infty }\beta _{n,\alpha }(q)\frac{t^{n}}{[n]_{q}!}\right) \left( -t^{2} \sum _{n=0}^{\infty }\frac{(1-q)^{2n+2}t^{2n}}{2^{2n+2}(q^{2},q^{2\alpha +2};q^{2})_{n+1}}\right) \\&\quad =\sum _{n=0}^{\infty }\left( \beta _{n,\alpha }(q)- \frac{(-1)^{n}}{2^{n}}\right) \frac{t^{n}}{[n]_{q}!}.\end{aligned} \end{aligned}$$

Therefore, by the Cauchy product formula

$$\begin{aligned} \begin{aligned}&-\frac{(1-q)^{2}}{4}\sum _{n=2}^{\infty }t^{n}\sum _{k=0}^{[\frac{n}{2}]-1}\frac{(1-q)^{2k}\beta _{n-2k-2,\alpha }(q)}{2^{2k}[n-2k-2]_{q}!(q^{2},q^{2\alpha +2};q^{2})_{k+1}} \\&\quad =\sum _{n=0}^{\infty }\left( \beta _{n,\alpha }(q)-\frac{(-1)^{n}}{2^{n}}\right) \frac{t^{n}}{[n]_{q}!}.\end{aligned}\nonumber \\ \end{aligned}$$
(2.76)

Equating the coefficient of \( t^{n}\) in (2.76), we get the required result and the theorem follows. \(\square \)

The following theorem gives a recursive relations between the polynomials \(B^{(k)}_{n,\alpha }(x;q)\) and \(B^{(k)}_{n,\alpha +1}(x;q)\,(k=2,3)\).

Theorem 2.26

If \( Re\,\alpha > -1 \) , \( x\in {\mathbb {C}},\) and \( k \in \,{\mathbb {N}}\), then

$$\begin{aligned}&\frac{B^{(2)}_{n,\alpha }(x;q)}{[n]_{q}!}=2(1-q^{2\alpha +2})\sum _{k=0}^{[\frac{n}{2}]}(-1)^{k}\frac{(1-q)^{2k}\,h^{(2)}_{k+1}(q^{2})\,}{[n-2k]_{q}!}B^{(2)}_{n-2k,\alpha +1}(x;q),\nonumber \\ \end{aligned}$$
(2.77)
$$\begin{aligned}&\frac{B^{(3)}_{n,\alpha }(x;q)}{[n]_{q}!}=(1-q^{2\alpha +2})\sum _{k=0}^{[\frac{n}{2}]}(-1)^{k}\frac{(1-q)^{2k}q^{\frac{-k}{2}}\,h^{(3)}_{k+1}(q^{2})\,}{2^{2k}[n-2k]_{q}!}B^{(3)}_{n-2k,\alpha +1}(x;q),\nonumber \\ \end{aligned}$$
(2.78)

where

$$\begin{aligned} h^{(r)}_{k}(q^{2})=\sum _{m=1}^{\infty }\frac{-2J^{(r)}_{\alpha +1}(j^{(r)}_{m,\alpha };q^{2})}{ \frac{d}{dz}J^{(r)}_{\alpha }(z;q^{2})\mid _{z=j^{(r)}_{m,\alpha }}}\left( \frac{1}{j^{(r)}_{m,\alpha }}\right) ^{2k}, \end{aligned}$$

and \((j^{(r)}_{m,\alpha })_{m=1}^{\infty }\,\,(r=2,3)\) are the positive zero of \( J^{(r)}_{\alpha }( \cdot ;q^{2}).\)

Proof

We start with the proof of (2.77). From [6, 13], we have the identity

$$\begin{aligned} \frac{J^{(2)}_{\alpha +1}(t;q)}{J^{(2)}_{\alpha }(t;q)}=\sum _{n=1}^{\infty }h^{(2)}_{n}(q)t^{2n-1}, \end{aligned}$$
(2.79)

where

$$\begin{aligned} h^{(2)}_{n}(q)=\sum _{m=1}^{\infty }\frac{-2J^{(2)}_{\alpha +1}(j^{(2)}_{m,\alpha };q^{2})}{ \frac{d}{dz}J^{(2)}_{\alpha }(z;q^{2})\mid _{z=j^{(2)}_{m,\alpha }} }\left( \frac{1}{j^{(2)}_{m,\alpha }}\right) ^{2n}. \end{aligned}$$

Replacing t by \( it(1-q)\) and q by \( q^{2}\) in (2.79), we obtain

$$\begin{aligned} \frac{1}{J^{(2)}_{\alpha }(it(1-q);q^{2})}= \frac{1}{J^{(2)}_{\alpha +1}(it(1-q);q^{2})}\sum _{n=1}^{\infty }h^{(2)}_{n}(it(1-q))^{2n-1}. \end{aligned}$$
(2.80)

Multiplying (2.80) by \( E_{q}(xt) E_{q}(\frac{-t}{2})\) to obtain

$$\begin{aligned} \frac{E_{q}(xt) E_{q}(\frac{-t}{2})}{J^{(2)}_{\alpha }(it(1-q);q^{2})}= \frac{E_{q}(xt) E_{q}(\frac{-t}{2})}{J^{(2)}_{\alpha +1}(it(1-q);q^{2})}\sum _{n=1}^{\infty }h^{(2)}_{n}(q^{2})(it(1-q))^{2n-1}. \end{aligned}$$
(2.81)

Substituting from (2.8) into (2.81), we get

$$\begin{aligned} \frac{E_{q}(xt) E_{q}(\frac{-t}{2})}{g^{(2)}_{\alpha }(it;q)}=\frac{(1+q)[\alpha +1]_{q^{2}}}{(\frac{it}{2})}\frac{E_{q}(xt) E_{q}(\frac{-t}{2})}{ g^{(2)}_{\alpha +1}(it;q)}\sum _{n=1}^{\infty }h^{(2)}_{n}(q^{2})(it(1-q))^{2n-1}.\nonumber \\ \end{aligned}$$
(2.82)

Consequently,

$$\begin{aligned} \begin{aligned} \sum _{n=0}^{\infty }&B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\\ {}&=\frac{2(1+q)}{it}[\alpha +1]_{q^{2}}\left( \sum _{n=0}^{\infty }B^{(2)}_{n,\alpha +1}(x;q) \frac{t^{n}}{[n]_{q}!} \right) \left( \sum _{n=1}^{\infty }h^{(2)}_{n}(q^{2})(it(1-q))^{2n-1}\right) \\ {}&=2(1-q^{2\alpha +2}) \left( \sum _{n=0}^{\infty }B^{(2)}_{n,\alpha +1}(x;q)\frac{t^{n}}{[n]_{q}!}\right) \left( \sum _{n=0}^{\infty }h^{(2)}_{n+1}(q^{2})t^{2n}(i(1-q))^{2n} \right) \\ {}&=2(1-q^{2\alpha +2})\sum _{n=0}^{\infty }t^{n}\sum _{k=0}^{[\frac{n}{2}]}(-1)^{k} \frac{(1-q)^{2k} \,h^{(2)}_{k+1}(q^{2})}{[n-2k]_{q}!}B^{(2)}_{n-2k,\alpha +1}(x;q). \end{aligned} \end{aligned}$$

Hence

$$\begin{aligned} \begin{aligned}&\sum _{n=0}^{\infty }B^{(2)}_{n,\alpha }(x;q)\frac{t^{n}}{[n]_{q}!}\\&\quad ={2(1-q^{2\alpha +2})}\sum _{n=0}^{\infty }t^{n}\sum _{k=0}^{[\frac{n}{2}]}\frac{(-1)^{k}(1-q)^{2k} \,h^{(2)}_{k+1}(q^{2})\,}{[n-2k]_{q}!}B^{(2)}_{n-2k,\alpha +1}(x;q).\end{aligned}\nonumber \\ \end{aligned}$$
(2.83)

Equating the coefficients of \( t^{n}\) in (2.83), we get (2.77). The proof of (2.78) follows from the identity (see [1, Eq. (4.3), p. 1201]),

$$\begin{aligned} \begin{aligned} \frac{J^{(3)}_{\alpha +1}(t;q)}{J^{(3)}_{\alpha }(t;q)}=\sum _{n=1}^{\infty }h^{(3)}_{n}(q)t^{2n-1}, \end{aligned} \end{aligned}$$

where

$$\begin{aligned} h^{(3)}_{n}(q)=\sum _{m=1}^{\infty }\frac{-2J^{(3)}_{\alpha +1}(j^{(3)}_{m,\alpha };q^{2})}{ \frac{d}{dz}J^{(3)}_{\alpha }(z;q^{2})\mid _{z=j^{(3)}_{m,\alpha }} }\left( \frac{1}{j^{(3)}_{m,\alpha }}\right) ^{2n}, \end{aligned}$$

and by using the same technique. \(\square \)

3 Asymptotic Relations for the Generalized q-Bernoulli Numbers

In this section, we derive asymptotic relations for the generalized q-Bernoulli numbers defined in (2.10).

Theorem 3.1

Let n be a non negative integer and \(\alpha \) be a complex number such that \(Re\,\alpha >-1\). Then for \(n\in {\mathbb {N}},\)

$$\begin{aligned} \begin{aligned} \beta _{2n,\alpha }(q)&=2(-1)^{n+1}(q;q)_{2n} \sum _{k=1}^{\infty }\frac{Cos_{q}(\frac{j^{(2)}_{k,\alpha }}{2(1-q)})}{(j^{(2)}_{k,\alpha })^{2n+1}\, \frac{d}{dz}{\mathcal {J}}_{\alpha }^{(2)}(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}},\\ \beta _{2n+1,\alpha }(q)&=2(-1)^{n}(q;q)_{2n+1} \sum _{k=1}^{\infty }\frac{Sin_{q}(\frac{j^{(2)}_{k,\alpha }}{2(1-q)})}{(j^{(2)}_{k,\alpha })^{2n+2}\,\frac{d}{dz}{\mathcal {J}}_{\alpha }^{(2)}(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}}, \end{aligned} \end{aligned}$$
(3.1)

where \({\mathcal {J}}_{\alpha }^{(2)} (z;q)\) is defined in (1.7).

Proof

Since

$$\begin{aligned} G(z):=\frac{E_{q}(\frac{-z}{2})}{g^{(2)}_{\alpha }(iz;q)}=\sum _{n=0}^{\infty }\beta _{n,\alpha }(q)\frac{z^{n}}{[n]_{q}!},\quad \mid z\mid <\frac{j^{(2)}_{1,\alpha }}{1-q}, \end{aligned}$$

then

$$\begin{aligned} \frac{\beta _{n,\alpha }(q)}{[n]_{q}!}=\frac{G^{(n)}(0)}{n!},\quad n\in {\mathbb {N}}_{0}. \end{aligned}$$

Now, we integrate \(f(z):=\dfrac{G(z)}{z^{n+1}},\,G(z)=\dfrac{E_{q}(\frac{-z}{2})}{g^{(2)}_{\alpha }(iz;q)}\) on the contour \(\Gamma _{m},\) where \(\Gamma _{m}\) is a circle of radius \(R_{m},\) \(\mid z_{m}\mid<\,R_{m}<\mid z_{m+1}\mid \). From the Cauchy Residue Theorem, see [2],

$$\begin{aligned} \int _{\Gamma _{m}} f(z) \,dz=2\pi i\sum \,Res(f,z_{k} ), \end{aligned}$$

where \(\{z_{k}\}\) are the poles of f that lie inside \(\Gamma _{m}\). The function f(z) has a pole at \(z=0\) of order \(n+1\) and simple poles at \(\pm z_{k}\) where \(z_{k}= \dfrac{ij^{(2)}_{k,\alpha }}{1-q},\,k\in {\mathbb {N}}\). Consequently,

$$\begin{aligned} I_{m}=\frac{1}{2\pi i}\int _{\Gamma _{m}} f(z) \,dz=Res(f(z),0)+\sum _{k=1}^{m}Res(f(z),\pm z_{k} ). \end{aligned}$$
(3.2)

Since

$$\begin{aligned} Res(f,0)= & {} \frac{f^{n}(0)}{n!}=\frac{\beta _{n,\alpha }(q)}{[n]_{q}!}, \\ \quad Res(f,z_{k} )= & {} \frac{E_{q}(\frac{-z_{k}}{2})}{\frac{d}{dz}g^{(2)}_{\alpha }(iz;q)\mid _{z=z_{k}}}\frac{1}{(z_{k})^{n+1}}\\ {}= & {} \frac{E_{q}(\frac{-ij^{(2)}_{k,\alpha }}{2(1-q)})}{\frac{d}{dz}{\mathcal {J}}^{(2)}_{\alpha }(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}}\frac{(i)^{-n}(1-q)^{n}}{(j^{(2)}_{k,\alpha })^{n+1}}, \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} Res(f,-z_{k} )&=\frac{E_{q}(\frac{z_{k}}{2})}{\frac{d}{dz}g^{(2)}_{\alpha }(iz;q)\mid _{z=-z_{k}}}\frac{1}{(-z_{k})^{n+1}}\\&=\frac{E_{q}(\frac{ij^{(2)}_{k,\alpha }}{2(1-q)})}{\frac{d}{dz}{\mathcal {J}}^{(2)}_{\alpha }(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}}\frac{(-i)^{-n}(1-q)^{n}}{(j^{(2)}_{k,\alpha })^{n+1}}. \end{aligned} \end{aligned}$$

Then Eq. (3.2) can be written as

$$\begin{aligned} \begin{aligned} I_{m}&= \frac{\beta _{n,\alpha }(q)}{[n]_{q}!}\\ {}&\quad +\sum _{k=1}^{m}2Re\,\left( (-i)^{-n}E_{q}\left( \frac{ij^{(2)}_{k,\alpha }}{2(1-q)}\right) \right) \frac{(1-q)^{n}}{(j^{(2)}_{k,\alpha })^{n+1}\frac{d}{dz}{\mathcal {J}}^{(2)}_{\alpha }(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}}, \end{aligned}\nonumber \\ \end{aligned}$$
(3.3)

substituting into (3.3) with \(-i=e^{-\frac{i\pi }{2}}\) gives

$$\begin{aligned} \begin{aligned} I_{m}&= \frac{\beta _{n,\alpha }(q)}{[n]_{q}!}+2(1-q)^{n}\cos \frac{n\pi }{2} \sum _{k=1}^{m}\frac{Cos_{q}\left( \frac{j^{(2)}_{k,\alpha }}{2(1-q)}\right) }{\frac{d}{dz}{\mathcal {J}}^{(2)}_{\alpha }(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}}\frac{1}{(j^{(2)}_{k,\alpha })^{n+1}} \\&\quad -2(1-q)^{n}\sin \frac{n\pi }{2}\sum _{k=1}^{m}\frac{Sin_{q}\left( \frac{j^{(2)}_{k,\alpha }}{2(1-q)}\right) }{\frac{d}{dz}{\mathcal {J}}^{(2)}_{\alpha }(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}} \frac{1}{(j^{(2)}_{k,\alpha })^{n+1}}. \end{aligned} \end{aligned}$$

Now, we show that the integral \(I_{m}\rightarrow 0\) as \(m\rightarrow \infty \). Bergweiller and Hayman [7] introduced the asymptotic relation for \(E_{q}(z)\),

$$\begin{aligned} \mid M(r;E_{q})\mid :=\sup \{\mid E_{q}(z)\mid :\mid z\mid =r\}\sim \,e^{\frac{-(\log r)^{2}}{2\log q}},\quad \text { when} \quad r=\mid z\mid \rightarrow \infty . \end{aligned}$$

In [4], Annaby and Mansour proved that for \(r=\mid z\mid \,\rightarrow \,\infty \)

$$\begin{aligned} z^{-\nu }J_{\nu }^{(2)}(z;q) \sim \,\exp \left( -\frac{(\log \,r)^{2}}{2\log \,q}-\frac{\log \,2}{\log \,q}\log \,r\right) . \end{aligned}$$

Hayman in [15] introduced the higher order asymptotics of \(J_{\nu }^{(2)}(z;q)\). Then, Annaby and Mansour, see [4], pointed out that the first order asymptotics of the zeros of \(J_{\nu }^{(2)}(z;q^{2})\) is given by

$$\begin{aligned} j_{m,\nu }^{(2)}=2q^{-2m}q^{-\nu +1}(1+O(q^{2m})), \quad (m\rightarrow \infty ). \end{aligned}$$

Hence if \((z_{m})_{m}\) are the positive zeros of \(g^{(2)}_{\alpha }(iz;q)\), then

$$\begin{aligned} \lim _{m\rightarrow \infty }\frac{z_{m}}{z_{m+1}}=\lim _{m\rightarrow \infty }\frac{j^{(2)}_{m,\nu }}{j^{(2)}_{m+1,\nu }}=q^{2}, \quad \lim _{m\rightarrow \infty }z_{m}=\infty . \end{aligned}$$
(3.4)

Let \( 0<\epsilon <(q^{-1}-1)\). There exists \(M_{0}\in {\mathbb {N}}\) such that if \(m\in {\mathbb {N}},\,m\ge M_{0}\), then

$$\begin{aligned} q^{2}(1-\epsilon )<\frac{z_{m}}{z_{m+1}}<q^{2}(1+\epsilon ). \end{aligned}$$

Hence \(z_{m}<qz_{m+1}\) for all \(m\ge M_{0}\). We can choose \(R_{m},\,\delta :=q^{-1}\displaystyle \sup _{m\ge M_{0}}\frac{z_{m}}{z_{m+1}}\) such that \((z_{m}<\delta R_{m}<q z_{m+1} <R_{m})\). Indeed,

$$\begin{aligned} \delta =q^{-1}\displaystyle \sup _{m\ge M_{0}}\frac{z_{m}}{z_{m+1}}\ge q^{-1}\frac{z_{m}}{z_{m+1}},\,\, m\ge M_{0}. \end{aligned}$$

But \(q z_{m+1} <R_{m}\) leads to \(\delta >\frac{z_{m}}{R_{m}}\) and so \( z_{m} <\delta R_{m}\). Now,

$$\begin{aligned} \delta =q^{-1}\displaystyle \sup _{ m\ge M_{0}}\frac{z_{m}}{z_{m+1}}\ge q^{-1}\displaystyle \lim _{m\rightarrow \infty }\frac{z_{m}}{z_{m+1}}=q^{-1}q^{2}=q. \end{aligned}$$

Also \(\delta =q^{-1}\displaystyle \sup _{ m\ge M_{0}}\frac{z_{m}}{z_{m+1}}\le q(1+\epsilon )<1\). Hence \(1>\delta \ge q\) and so by

$$\begin{aligned} z_{m}\le R_{m}\le \frac{q}{\delta } z_{m+1}\le z_{m+1}, \end{aligned}$$
(3.5)

the annulus \(\delta R_{m}<|z| <R_{m}\) has no zeros of the function \( g^{(2)}_{\alpha }(iz;q)\). Hence, from the minimum modulus principle we have

$$\begin{aligned} \begin{aligned} \mid g^{(2)}_{\alpha }(iz;q)\mid&\ge c_{1}\,e^{-\frac{(\log \delta R_{m})^{2}}{2\log q}-\frac{\log 2}{\log q}\log \delta R_{m}},\quad c_{1}>0.\\ \mid E_{q}(\frac{-z}{2}) \mid&\le \,c_{2}\,e^{\frac{-(\log \frac{R_{m}}{2})^{2}}{2\log q}},\quad c_{2}>0.\end{aligned} \end{aligned}$$
(3.6)

Therefore, from (3.6), we conclude that

$$\begin{aligned} \begin{aligned} \mid \frac{E_{q}(\frac{-z}{2})}{g^{(2)}_{\alpha }(iz;q)} \mid&\le \,\frac{c_{2}}{c_{1}}\,\,\frac{e^{\frac{-(\log \frac{ R_{m}}{2})^{2}}{2\log q}}}{e^{-\frac{( \log \delta R_{m})^{2}}{2\log q}-\frac{\log 2}{\log q}\log \delta R_{m}}}\\&\le \,\frac{ c_{2}}{c_{1}}\,e^{\frac{1}{2\log q}\left( ((\log \delta R_{m})^{2}-(\log \frac{R_{m}}{2})^{2})\right) +\frac{\log 2}{\log q}\log \delta R_{m}} \\&\le \,\frac{ c_{2}}{c_{1}}\,e^{K}\,e^{\frac{2\log 2 \log R_{m}}{\log q}+\frac{\log \delta \log R_{m}}{\log q}}, \end{aligned} \end{aligned}$$

where

$$\begin{aligned} K=\frac{1}{2\log q}\left( (\log \delta )^{2} -(\log 2)^{2}+2\log 2\log \delta \right) . \end{aligned}$$

Now, using the ML-inequality (see [2]) to obtain

$$\begin{aligned} \begin{aligned} \mid I_{m}\mid&=\mid \int _{\Gamma _{m}}f(z)dz\mid \le (2\pi R_{m})\mid M(r;f(z))\mid \\&\le \,\frac{2\pi R_{m}c_{2}}{c_{1}}\,e^{K}\,e^{\frac{2\log 2 \log R_{m}}{\log q}+\frac{\log \delta \log R_{m}}{\log q}}\frac{1}{R_{m}^{n+1}}\\&\le \,\frac{ 2\pi c_{2}}{c_{1}}\,e^{K}\,R_{m}^{\frac{2 \log 2}{\log q}}R_{m}^{\frac{\log \delta }{\log q}-n}. \end{aligned} \end{aligned}$$
(3.7)

From (3.4) and (3.5), we have \(\displaystyle \lim _{m\rightarrow \infty }R_{m}=\infty \). Also, since \(0<q<1\) and \(1>\delta \ge q\) then

$$\begin{aligned} R_{m}^{\frac{2 \log 2}{\log q}}\rightarrow 0\quad and \quad R_{m}^{\frac{\log \delta }{\log q}-n}\rightarrow 0 \quad as\,\, m\,\,\rightarrow \,\infty . \end{aligned}$$

Hence \(\displaystyle \lim _{m\rightarrow \infty }I_{m}=0.\) Consequently,

$$\begin{aligned} \begin{aligned} \frac{\beta _{n,\alpha }(q)}{[n]_{q}!}&=-2(1-q)^{n}\cos \frac{n\pi }{2} \sum _{k=1}^{\infty }\frac{Cos_{q}\left( \frac{j^{(2)}_{k,\alpha }}{2(1-q)}\right) }{\frac{d}{dz}{\mathcal {J}}^{(2)}_{\alpha }(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}}\frac{1}{(j^{(2)}_{k,\alpha })^{n+1}} \\&\quad +2(1-q)^{n}\sin \frac{n\pi }{2}\sum _{k=1}^{\infty }\frac{Sin_{q}\left( \frac{j^{(2)}_{k,\alpha }}{2(1-q)}\right) }{\frac{d}{dz}{\mathcal {J}}^{(2)}_{\alpha }(z;q)\mid _{z=j^{(2)}_{k,\alpha }}} \frac{1}{(j^{(2)}_{k,\alpha })^{n+1}}. \end{aligned} \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned} \beta _{2n,\alpha }(q)&=2(-1)^{n+1}(q;q)_{2n} \sum _{k=1}^{\infty }\frac{Cos_{q}\left( \frac{j^{(2)}_{k,\alpha }}{2(1-q)}\right) }{(j^{(2)}_{k,\alpha })^{2n+1}\, \frac{d}{dz}{\mathcal {J}}_{\alpha }^{(2)}(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}},\\ \beta _{2n+1,\alpha }(q)&=2(-1)^{n}(q;q)_{2n+1} \sum _{k=1}^{\infty }\frac{Sin_{q}\left( \frac{j^{(2)}_{k,\alpha }}{2(1-q)}\right) }{(j^{(2)}_{k,\alpha })^{2n+2}\,\frac{d}{dz}{\mathcal {J}}_{\alpha }^{(2)}(z;q^{2})\mid _{z=j^{(2)}_{k,\alpha }}}, \end{aligned} \end{aligned}$$

which completes the proof of the theorem. \(\square \)

Remark 3.2

If we substitute with \(\alpha =\frac{1}{2}\) in the second equation in (3.1), then \((z_{k})_{k}\) will be the positive zeros of \(Sin_{q}(z)\) and consequently, the series in the left hand side vanishes which coincide with the known result that the odd Bernoulli numbers vanish \((\beta _{2n+1}(q)=0,\,n\ge \, 1)\) (see [19]). Similarly, if we set \(\alpha =-\frac{1}{2}\) in the first equation in (3.1), the series in the left hand side vanishes and this coincide with the fact that the even Euler\(^{,}\)s numbers are zero \((E_{2n}(q)=0,\,n\ge \,1)\) (see [19]).

Corollary 3.3

The asymptotic relations of the generalized q-Bernoulli numbers \((\beta _{n,\alpha }(q))_{n},\)

$$\begin{aligned} \begin{aligned} \beta _{2n,\alpha }(q)&=2(-1)^{n+1}(q;q)_{2n} \frac{Cos_{q}\left( \frac{j^{(2)}_{1,\alpha }}{2(1-q)}\right) }{(j^{(2)}_{1,\alpha })^{2n+1}\, \frac{d}{dz}{\mathcal {J}}_{\alpha }^{(2)}(z;q^{2})\mid _{z=j^{(2)}_{1,\alpha }}}\left( 1+o(1)\right) ,\\ \beta _{2n+1,\alpha }(q)&=2(-1)^{n}(q;q)_{2n+1} \frac{Sin_{q}\left( \frac{j^{(2)}_{1,\alpha }}{2(1-q)}\right) }{(j^{(2)}_{1,\alpha })^{2n+2}\,\frac{d}{dz}{\mathcal {J}}_{\alpha }^{(2)}(z;q^{2})\mid _{z=j^{(2)}_{1,\alpha }}}\left( 1+o(1)\right) , \end{aligned} \end{aligned}$$

where \({\mathcal {J}}_{\alpha }^{(2)} (z;q)\) is defined in (1.7).

Proof

The proof follows directly from Theorem 3.1. \(\square \)

4 Applications of the Generalized q-Bernoulli Polynomials

In this section, we introduce connection relations between the generalized q-Bernoulli polynomials \( B^{(k)}_{n,\alpha }(x;q)\,(k=1,2,3)\) and the q-Laguerre and the little q-Legendre polynomials.

The q-Laguerre polynomials \( L_{n}^{\alpha }(x;q)\) of degree n are defined by

$$\begin{aligned} \begin{aligned} L_{n}^{\alpha }(x;q):&=\frac{1}{(q;q)_{n}}\,_{2}\phi _{1} \left( \begin{array}{c} q^{-n},-x \\ 0 \\ \end{array} q;q^{n+\alpha +1} \right) \\&=\frac{(q^{\alpha +1};q)_{n}}{(q;q)_{n}}\sum _{k=0}^{n} \frac{(q^{-n};q)_{k}}{(q^{\alpha +1};q)_{k}}(-1)^{k}(q^{n+\alpha +1})^{k}x^{k}. \end{aligned} \end{aligned}$$
(4.1)

The Rodrigues formula is given by

$$\begin{aligned} \begin{aligned} L_{n}^{\alpha }(x;q)=\frac{(1-q)^{n}}{(q;q)_{n}}(-x;q)_{\infty }\,x^{-\alpha }\,D_{q}^{n}\left( \frac{x^{\alpha +n}}{(-x;q)_{\infty }}\right) , \end{aligned} \end{aligned}$$
(4.2)

and the orthogonality relation is

$$\begin{aligned} \begin{aligned}&\int _{0}^{\infty }\frac{x^{\alpha }}{(-x;q)_{\infty }}L_{m}^{\alpha }(x;q)L_{n}^{\alpha }(x;q)dx \\&\quad =\frac{(q^{-\alpha };q)_{\infty }}{(q;q)_{\infty }}\frac{(q^{\alpha +1};q)_{n}}{(q;q)_{n}q^{n}}\Gamma _{q}(-\alpha )\Gamma _{q}(\alpha +1)\delta _{mn},\end{aligned}\nonumber \\ \end{aligned}$$
(4.3)

\(\alpha >-1\), where \(\delta _{mn}\) is the Kronecker delta function, see [17, 21]. The q-Laguerre polynomials \( L_{n}^{\alpha }(x;q)\) satisfy three term recurrence relation

$$\begin{aligned} \begin{aligned} -xa_{n}L_{n}^{\alpha }(x;q)=L_{n+1}^{\alpha }(x;q)-b_{n}L_{n}^{\alpha }(x;q)+d_{n}L_{n-1}^{\alpha }(x;q), \end{aligned} \end{aligned}$$

where

$$\begin{aligned} a_{n}=\frac{q^{2n+\alpha +1}}{1-q^{n+1}},\quad b_{n}=1+\frac{q(1-q^{n+\alpha }}{1-q^{n+1}},\quad d_{n}=\frac{q(1-q^{n+\alpha })}{1-q^{n+1}}. \end{aligned}$$

In the following, let \(\alpha >-1\) and \({\mathbb {P}}_{n}=\{p(x):\text{ deg }\, p(x)\le n\}\) with the inner product

$$\begin{aligned} \langle p(x),g(x)\rangle = \int _{0}^{\infty }\frac{x^{\alpha }}{(-x;q)_{\infty }}p(x)g(x)dx, \end{aligned}$$

where \(p(x),\,g(x)\in {\mathbb {P}}_{n}\).

Theorem 4.1

Let p(x) \(\in {\mathbb {P}}_{n}.\) Then p(x) can be expanded as

$$\begin{aligned} p(x)=\sum _{m=0}^{n}C_{m}L_{m}^{\alpha }(x;q), \end{aligned}$$

where

$$\begin{aligned} C_{m}=\frac{q^{m}(1-q)^{m-1}(q^{\alpha +m+1};q)_{\infty }}{(q;q)_{\infty }}\int _{0}^{\infty } \,D_{q}^{m}\left( \frac{x^{\alpha +m}}{(-x;q)_{\infty } }\right) p(x) dx. \end{aligned}$$

Proof

Since

$$\begin{aligned} p(x)=\sum _{m=0}^{n}C_{m}L_{m}^{\alpha }(x;q), \end{aligned}$$

in order to calculate the constant \(C_{m}\), we use (4.3) to obtain

$$\begin{aligned} \left\langle p(x),L_{k}^{\alpha }(x;q)\right\rangle= & {} \left\langle \sum _{m=0}^{n}C_{m}L_{m}^{\alpha }(x;q),L_{k}^{\alpha }(x;q)\right\rangle \\ {}= & {} \sum _{m=0}^{n}C_{m}\langle L_{m}^{\alpha }(x;q),L_{k}^{\alpha }(x;q)\rangle . \end{aligned}$$

Then

$$\begin{aligned} \begin{aligned} \langle p(x),L_{m}^{\alpha }(x;q)\rangle&=C_{m}\langle L_{m}^{\alpha }(x;q) ,L_{m}^{\alpha }(x;q)\rangle \\&=C_{m}\frac{(q^{\alpha +1};q)_{m}}{q^{m}(q;q)_{m}}(1-q)^{1+\alpha }\Gamma _{q}(\alpha +1). \end{aligned} \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned} C_{m}= \frac{q^{m}(q;q)_{m}}{(q^{\alpha +1};q)_{m}(1-q)^{1+\alpha }\Gamma _{q}(\alpha +1)}\int _{0}^{\infty }\frac{x^{\alpha }}{(-x;q)_{\infty }}L_{m}^{\alpha }(x;q) p(x) dx. \end{aligned} \end{aligned}$$
(4.4)

Using (4.2) with n replaced by m, we obtain

$$\begin{aligned} \begin{aligned} C_{m}=\frac{q^{m}(1-q)^{m-1}(q^{\alpha +m+1};q)_{\infty }}{(q;q)_{\infty }}\int _{0}^{\infty } \,D_{q}^{m}\left( \frac{x^{\alpha +m}}{(-x;q)_{\infty } }\right) p(x) dx, \end{aligned} \end{aligned}$$

and the theorem follows. \(\square \)

The following Lemma, see [18], is essential in the proof of Theorem 4.3.

Lemma 4.2

Let the functions f and g be defined and continuous on \([0,\infty ]\). Assume that the improper Riemann integrals of the functions f(x)g(x) and f(x/q)g(x) exist on \([0,\infty ]\). Then

$$\begin{aligned} \begin{aligned} \int _{0}^{\infty }f(x)D_{q}g(x)dx&=\frac{f(0)g(0)}{1-q}\ln q-\frac{1}{q}\int _{0}^{\infty }g(x)D_{q^{-1}}f(x)dx\\&=\frac{f(0)g(0)}{1-q}\ln q-\int _{0}^{\infty }g(qx)D_{q}f(x)dx. \end{aligned} \end{aligned}$$

Theorem 4.3

If \(n\in {\mathbb {N}}\) and \(x\in {\mathbb {C}},\) then

$$\begin{aligned} B^{(1)}_{n,\alpha }(x;q)&=\sum _{m=0}^{n}A_{m}\left( \sum _{k=m}^{n}q^{\frac{k(2n-k+1)}{2}}\frac{ (q^{-n},q^{-\alpha -k};q)_{k} (q^{-k};q)_{m}}{(q;q)_{k}} \beta _{n-k,\alpha } (q)\right) L_{m} ^{\alpha }(x;q),\\ B^{(2)}_{n,\alpha }(x;q)&=\sum _{m=0}^{n}A_{m}\left( \sum _{k=m}^{n}\frac{q^{nk}(q^{-n},q^{-\alpha -k};q)_{k} (q^{-k};q)_{m}}{(q;q)_{k}} \beta _{n-k,\alpha } (q)\right) L_{m} ^{\alpha }(x;q),\\ B^{(3)}_{n,\alpha }(x;q)&=\sum _{m=0}^{n}A_{m}\left( \sum _{k=m}^{n}q^{\frac{k(4n-k+1)}{4}}\frac{ (q^{-n},q^{-\alpha -k};q)_{k} (q^{-k};q)_{m}}{(q;q)_{k}} \beta ^{(3)}_{n-k,\alpha } (q)\right) L_{m} ^{\alpha }(x;q), \end{aligned}$$

where

$$\begin{aligned} A_{m}=\frac{-q^{m}(q^{\alpha +m+1},q^{-\alpha };q)_{\infty }}{(1-q)(q,q;q)_{\infty }}\frac{\pi }{sin(\alpha \pi )}. \end{aligned}$$

Proof

We prove the identity for \(B^{(1)}_{n,\alpha }(x;q)\) and the proofs for \(B^{(k)}_{n,\alpha }(x;q)\,(k=2,3)\) are similar. Substitute with \(p(x)=B^{(1)}_{n,\alpha }(x;q)\) in (4.4). This gives

$$\begin{aligned} \begin{aligned} C_{m}= \frac{q^{m}(q;q)_{m}}{(q^{\alpha +1};q)_{m}(1-q)^{1+\alpha }\Gamma _{q}(\alpha +1)}\int _{0}^{\infty }\frac{x^{\alpha }}{(-x;q)_{\infty }}L_{m}^{\alpha }(x;q) B^{(1)}_{n,\alpha }(x;q) dx. \end{aligned}\nonumber \\ \end{aligned}$$
(4.5)

Since \(\{L_{m}^{\alpha }(x;q)\}_{n\in {\mathbb {N}}}\) is an orthogonal polynomials sequence then \( C_{m}=0\) for \( m>n\), and

$$\begin{aligned} B^{(1)}_{n,\alpha }(x;q) =\sum _{m=0}^{n}C_{m}L_{m}^{\alpha }(x;q). \end{aligned}$$

Now, we calculate \(C_{m}\). Using (2.12) in (4.5) gives

$$\begin{aligned} \begin{aligned} C_{m}&=\frac{q^{m}(q;q)_{m}}{(q^{\alpha +1};q)_{m}(1-q)^{1+\alpha }\Gamma _{q}(\alpha +1)}\sum _{k=0}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \beta _{n-k,\alpha } (q) \int _{0}^{\infty } \,\frac{x^{\alpha +k}}{(-x;q)_{\infty }}L_{m}^{\alpha }(x;q)dx.\end{aligned} \end{aligned}$$

Since

$$\begin{aligned} \int _{0}^{\infty }\frac{x^{\alpha }}{(-x;q)_{\infty }} L_{m}^{\alpha }(x;q)x^{k} dx=0, \quad \text {for}\,\, k<m, \end{aligned}$$

then

$$\begin{aligned} \begin{aligned} C_{m}&=\frac{q^{m}(q;q)_{m}}{(q^{\alpha +1};q)_{m}(1-q)^{1+\alpha }\Gamma _{q}(\alpha +1)}\sum _{k=m}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \beta _{n-k,\alpha } (q) \int _{0}^{\infty } \,\frac{x^{\alpha +k}}{(-x;q)_{\infty }}L_{m}^{\alpha }(x;q) dx. \end{aligned} \end{aligned}$$

From (4.2), we get

$$\begin{aligned} \begin{aligned} C_{m}&=\frac{q^{m}(1-q)^{m-1}(q^{\alpha +m+1};q)_{\infty }}{(q;q)_{\infty }}\sum _{k=m}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \beta _{n-k,\alpha } (q) \int _{0}^{\infty } \,D_{q}^{m}\left( \frac{x^{\alpha +m}}{(-x;q)_{\infty }}\right) x^{k} dx, \end{aligned} \end{aligned}$$

then applying the q-integration by part introduced in Lemma 4.2m times, we obtain

$$\begin{aligned} \begin{aligned} C_{m}=\,&\frac{(-1)^{m}q^{m}(1-q)^{m-1}(q^{\alpha +m+1};q)_{\infty }}{(q;q)_{\infty }}\\&\times \sum _{k=m}^{n}\left[ \begin{array}{c} n \\ k \\ \end{array} \right] _{q} \left( \prod _{i=0}^{m-1}q^{i-k}\right) \frac{[k]_{q}!}{[k-m]_{q}!} \beta _{n-k,\alpha } (q) \int _{0}^{\infty } \frac{x^{\alpha +k}}{(-x;q)_{\infty }}dx. \end{aligned} \end{aligned}$$

From [16, Eq. (5.4), p. 465],

$$\begin{aligned} \frac{1}{ \Gamma _{q}(z)}=\frac{sin\pi z}{\pi }\int _{0}^{\infty } \frac{ t^{-z}}{(-t(1-q);q)_{\infty }}dt, \,\,\,Re\,z>0. \end{aligned}$$

Then

$$\begin{aligned} \int _{0}^{\infty } \frac{x^{\alpha +k}}{(-x;q)_{\infty }}dx= \frac{\pi }{sin(-\alpha -k)\pi }\frac{1}{ \Gamma _{q}(-\alpha -k)}(1-q)^{\alpha +k+1}. \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned} C_{m}&=\frac{(-1)^{m}q^{m}(q^{\alpha +m+1};q)_{\infty }}{(1-q)(q,q;q)_{\infty }}\\&\quad \times \sum _{k=m}^{n} \left( \prod _{i=0}^{m-1}q^{i-k}\right) \frac{(q;q)_{n}(q^{-\alpha -k};q)_{\infty }}{(q;q)_{n-k}(q;q)_{k-m}}\frac{\pi }{sin(-\alpha -k)\pi }\beta _{n-k,\alpha }(q). \end{aligned}\nonumber \\ \end{aligned}$$
(4.6)

Since

$$\begin{aligned} \frac{\pi }{sin(-\alpha -k)\pi } =(-1)^{k-1}\frac{\pi }{sin(\alpha \pi )},\quad \prod _{i=0}^{m-1}q^{i-k}=q^{\frac{m(m-1)}{2}}q^{-km}, \end{aligned}$$
(4.7)

then substituting from (4.7) into (4.6), we get

$$\begin{aligned} \begin{aligned} C_{m}=\,&\frac{(-1)^{m}q^{m}q^{m(m-1)/2}(q^{\alpha +m+1};q)_{\infty }(q^{-\alpha };q)_{\infty }}{(1-q)(q,q;q)_{\infty }}\frac{\pi }{sin(\alpha \pi )}\\&\times \sum _{k=m}^{n} (-1)^{k-1}q^{-km} \frac{(q;q)_{n}(q^{-\alpha -k};q)_{k}}{(q;q)_{n-k}(q;q)_{k-m}}\beta _{n-k,\alpha }(q). \end{aligned} \end{aligned}$$

Using the relation (2.47), we obtain

$$\begin{aligned} \begin{aligned} C_{m}=\,&\frac{-q^{m}(q^{\alpha +m+1};q^{-\alpha };q)_{\infty }}{(1-q)(q,q;q)_{\infty }}\frac{\pi }{sin(\alpha \pi )}\\&\times \sum _{k=m}^{n}q^{\frac{k(2n-k+1)}{2}}\frac{ (q^{-n};q)_{k} (q^{-k};q)_{m}(q^{-\alpha -k};q)_{k}}{(q;q)_{k}} \beta _{n-k,\alpha }(q), \end{aligned} \end{aligned}$$

and this completes the proof of the theorem. \(\square \)

The little q-Legendre polynomials \(( P_{n}(x\mid q))_{n}\) are defined by

$$\begin{aligned} \begin{aligned} P_{n}(x\mid q)&=\,_{2}\phi _{1} \left( \begin{array}{c} q^{-n},q^{n+1} \\ q \\ \end{array} q;qx \right) \\&=\sum _{k=0}^{n}\frac{(q^{-n};q)_{k}(q^{n+1};q)_{k}}{(q;q)_{k}}\frac{q^{k}x^{k}}{(q;q)_{k}}. \end{aligned} \end{aligned}$$

They satisfy the Rodrigues formula

$$\begin{aligned} P_{n}(x\mid q)=\frac{q^{n(n-1)/2}(1-q)^{n}}{(q;q)_{n}}\,D_{q^{-1}}^{n}(x^{n}(qx;q)_{n}), \quad \text{ for }\,\, n\ge 0, \end{aligned}$$
(4.8)

and the orthogonality relation

$$\begin{aligned} \int _{0}^{1}P_{m}(x\mid q)P_{n}(x \mid q)d_{q}x =\frac{(1-q)}{(1-q^{2n+1})}\delta _{mn},\quad \text{ for }\,\, m,\,n\ge 0, \end{aligned}$$
(4.9)

see [21]. Let \({\mathbb {P}}_{n}=\{g(x):\text{ deg }\, g(x)\le n\}\) with the inner product

$$\begin{aligned} \langle g(x),p(x)\rangle = \int _{0}^{1}g(x)p(x)d_{q}x, \end{aligned}$$

where \(p(x),\,g(x)\in {\mathbb {P}}_{n}\).

Theorem 4.4

Let g(x) \(\in {\mathbb {P}}_{n}\). Then g(x) can be represented by

$$\begin{aligned} g(x)=\sum _{k=0}^{n}C_{k}P_{k}(x\mid q), \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} C_{k}=\frac{q^{k(k-1)/2}(1-q)^{k-1}(1-q^{2k+1})}{(q;q)_{k}}\int _{0}^{1}\,D_{q^{-1}}^{k}(x^{k}(qx;q)_{k})g(x)d_{q}x. \end{aligned} \end{aligned}$$

Proof

Since

$$\begin{aligned} g(x)=\sum _{k=0}^{n}C_{k}P_{k}(x\mid q), \end{aligned}$$

then by the orthogonality relation (4.9), we obtain

$$\begin{aligned} C_{k}=\frac{(1-q^{2k+1})}{(1-q)} \langle g(x),P_{k}(x \mid q)\rangle =\frac{(1-q^{2k+1})}{(1-q)}\int _{0}^{1}P_{k}(x\mid q)g(x)d_{q}x.\nonumber \\ \end{aligned}$$
(4.10)

By using (4.8), we get

$$\begin{aligned} \begin{aligned} C_{k}=\frac{q^{k(k-1)/2}(1-q)^{k-1}(1-q^{2k+1})}{(q;q)_{k}}\int _{0}^{1}\,D_{q^{-1}}^{k}(x^{k}(qx;q)_{k})g(x)d_{q}x, \end{aligned} \end{aligned}$$

which readily gives the result. \(\square \)

Theorem 4.5

For \(n\in {\mathbb {N}}\) and \(x\in {\mathbb {C}}\),

$$\begin{aligned} \begin{aligned} B^{(1)}_{n,\alpha }(x;q)&=\sum _{k=0}^{n}\lambda _{k}\left( \sum _{m=k}^{n}(-1)^{m}q^{\frac{m(2n-m+1)}{2}+1}\,\frac{(q^{-n};q)_{m}(q^{-m};q)_{k}}{(q;q)_{m+k+1}}\beta _{n-m,\alpha }(q)\right) P_{k}(x \mid q),\\ B^{(2)}_{n,\alpha }(x;q)&=\sum _{k=0}^{n}\lambda _{k}\left( \sum _{m=k}^{n}(-1)^{m}q^{nm+1}\, \frac{(q^{-n};q)_{m}(q^{-m};q)_{k}}{(q;q)_{m+k+1}}\beta _{n-m,\alpha }(q)\right) P_{k}(x \mid q),\\ B^{(3)}_{n,\alpha }(x;q)&=\sum _{k=0}^{n}\lambda _{k}\left( \sum _{m=k}^{n}(-1)^{m}q^{\frac{m(4n-m+1)}{4}+1}\, \frac{(q^{-n};q)_{m}(q^{-m};q)_{k}}{(q;q)_{m+k+1}}\beta ^{(3)}_{n-m,\alpha }(q)\right) P_{k}(x\mid q),\end{aligned} \end{aligned}$$

where

$$\begin{aligned} \lambda _{k}=q^{\frac{-k(k-3)}{2}}(1-q^{2k+1}). \end{aligned}$$

Proof

Substitute with \(g(x)=B^{(1)}_{n,\alpha }(x;q)\) in (4.10), we obtain

$$\begin{aligned} \begin{aligned} C_{k} = \dfrac{(1-q^{2k+1})}{(1-q)}\int _{0}^{1}P_{k}(x \mid q) B^{(1)}_{n,\alpha }(x;q)d_{q}x. \end{aligned} \end{aligned}$$
(4.11)

Since the polynomials \(\{P_{k}(x\mid q)\}\) are orthogonal, then \( C_{k}=0\) for \(k>n\), and

$$\begin{aligned} B^{(1)}_{n,\alpha }(x;q) =\sum _{k=0}^{n}C_{k}P_{k}(x\mid q). \end{aligned}$$
(4.12)

Set

$$\begin{aligned} B^{(1)}_{n,\alpha }(x;q)=\sum _{m=0}^{n}\left[ \begin{array}{c} n \\ m \\ \end{array} \right] _{q} \beta _{n-m,\alpha } (q) x^{m}. \end{aligned}$$

From (4.11),

$$\begin{aligned} \begin{aligned} C_{k}&=\dfrac{(1-q^{2k+1})}{(1-q)} \sum _{m=0}^{n}\left[ \begin{array}{c} n \\ m \\ \end{array} \right] _{q} \beta _{n-m,\alpha } (q)\int _{0}^{1}P_{k}(x\mid q) x^{m}d_{q}x\\&=\frac{(1-q^{2k+1})}{(1-q)} \sum _{m=k}^{n}\left[ \begin{array}{c} n \\ m \\ \end{array} \right] _{q} \beta _{n-m,\alpha } (q)\int _{0}^{1}P_{k}(x\mid q) x^{m}d_{q}x, \end{aligned} \end{aligned}$$

since

$$\begin{aligned} \int _{0}^{1} P_{k}(x\mid q) x^{m}d_{q}x=0\quad for\,\,m<k. \end{aligned}$$

Hence, by the Rodrigues formula in (4.8), we obtain

$$\begin{aligned} \begin{aligned} C_{k}&= \dfrac{(1-q^{2k+1})q^{k(k-1)/2}(1-q)^{k-1}}{(q;q)_{k}} \sum _{m=k}^{n}\left[ \begin{array}{c} n \\ m \\ \end{array} \right] _{q} \beta _{n-m,\alpha } (q)\\&\quad \times \int _{0}^{1}\,D_{q^{-1}}^{k}(x^{k}(qx;q)_{k}) x^{m}d_{q}x. \end{aligned} \end{aligned}$$
(4.13)

Using the \(q^{-1}\)-integration by parts

$$\begin{aligned} \int _{0}^{a}f\left( \frac{t}{q}\right) D_{q^{-1}}g(t)d_{q}t =q\left( (fg)\left( \frac{a}{q}\right) -(fg)(0) \right) -\int _{0}^{a}g(t)D_{q^{-1}}f(t)d_{q}t,\nonumber \\ \end{aligned}$$
(4.14)

where f and g are continuous functions at zero, see [5]. This gives

$$\begin{aligned} \begin{aligned} \int _{0}^{1}\,D_{q^{-1}}^{k}(x^{k}(qx;q)_{k}) x^{m}d_{q}x=&q\left[ x^{m}D_{q^{-1}}^{k-1}(x^{k}(qx;q)_{k})\right] _{0}^{\frac{1}{q}} \\&-[m]_{q}q^{1-m}\int _{0}^{1}x^{m-1}D_{q^{-1}}^{k-1}(x^{k}(qx;q)_{k})d_{q}x. \end{aligned} \end{aligned}$$
(4.15)

The first term on the right hand side of (4.15) vanishes because

$$\begin{aligned} D_{q^{-1}}(x^{k}(qx;q)_{k})=[k]_{q^{-1}}x^{k-1}(x;q)_{k}+x^{k}D_{q^{-1}}(qx;q)_{k}, \end{aligned}$$

and

$$\begin{aligned} D_{q^{-1}}^{j}(qx;q)_{k}\mid _{x=\frac{1}{q}}=a^{k} \frac{[k]_{q}!}{[k-j]_{q}!}(1;q)_{k-j}=0,\quad \text{ for }\,\,j=0,1,\ldots ,k-1. \end{aligned}$$

Therefore,

$$\begin{aligned} \int _{0}^{1}\,D_{q^{-1}}^{k}(x^{k}(qx;q)_{k}) x^{m}d_{q}x=-[m]_{q}q^{1-m}\int _{0}^{1}x^{m-1}D_{q^{-1}}^{k-1}(x^{k}(qx;q)_{k})d_{q}x.\nonumber \\ \end{aligned}$$
(4.16)

Now, applying (4.14) \(k-1\) times on the right hand side of (4.16), and using that \(D_{q^{-1}}^{m}(x^{k}(qx;q)_{k}= 0\) at \(x=0,\, x=\frac{1}{q}\) \((m=0,1,\ldots ,k-1)\) yields

$$\begin{aligned} \begin{aligned} \int _{0}^{1}\,D_{q^{-1}}^{k}(x^{k}(qx;q)_{k}) x^{m}d_{q}x&=(-1)^{k} \left( \prod _{j=0}^{k-1}q^{1-m-j}\right) \frac{[m]_{q}!}{[m-k]_{q}!} \int _{0}^{1} x^{m}(qx;q)_{k}\,d_{q}x. \end{aligned} \end{aligned}$$

Since

$$\begin{aligned} \begin{aligned} B_{q}(x,y)&=\int _{0}^{1} t^{x-1}(qt;q)_{y-1}d_{q}t\\ {}&=\int _{0}^{1}t^{x-1}\frac{(tq;q)_{\infty }}{(tq^{y};q)_{\infty }}d_{q}t,\,\,\,Re\,(x)>0,\,Re\,(y)>0,\end{aligned} \end{aligned}$$

see [5, Eq. (1.58), p. 22], then

$$\begin{aligned} \begin{aligned} \int _{0}^{1}D_{q^{-1}}^{k}(x^{k}(qx;q)_{k}) x^{m}d_{q}x&=(-1)^{k}q^{\frac{-k(k-1)}{2}+k}q^{-mk} \frac{[m]_{q}!}{[m-k]_{q}!}B_{q}(m{+}1,k{+}1)\\ {}&= (-1)^{k}q^{\frac{-k(k-3)}{2}}q^{-mk} \frac{[m]_{q}!\Gamma _{q}(m+1)\Gamma _{q}(k+1)}{[m-k]_{q}!\Gamma _{q}(m+k+2)}\\ {}&= (-1)^{k}q^{\frac{-k(k-3)}{2}}q^{-mk} \frac{([m]_{q}!)^{2}[k]_{q}!}{[m-k]_{q}![m+k+1]_{q}!}. \end{aligned}\nonumber \\ \end{aligned}$$
(4.17)

Substituting from (4.17) into (4.13) yields

$$\begin{aligned} \begin{aligned} C_{k}&=(-1)^{k}q^{k}(1-q^{2k+1}) \sum _{m=k}^{n}q^{-mk}\frac{(q;q)_{n}(q;q)_{m}}{(q;q)_{n-m}(q;q)_{m-k}(q;q)_{m+k+1}}\beta _{n-m,\alpha }(q)\\ {}&= q^{\frac{-k(k-3)}{2}}(1-q^{2k+1}) \sum _{m=k}^{n}(-1)^{m} \frac{q^{\frac{m(2n-m+1)}{2}+1}(q^{-n};q)_{m}(q^{-m};q)_{k}}{(q;q)_{m+k+1}}\beta _{n-m,\alpha }(q), \end{aligned}\nonumber \\ \end{aligned}$$
(4.18)

where we used the identity in (2.47). Therefore, from (4.18) and (4.12), we get the required result for \(B^{(1)}_{n,\alpha }(x;q)\). Similarly, we can prove the result for \(B^{(k)}_{n,\alpha }(x;q)\,(k=2,3)\). \(\square \)