1 Introduction

The objective of this paper is to extend the work of Liu [14, 15] and Liu and Zeng [17]. These authors have found a q-difference equation related to Rogers–Szegö polynomials [21] which can be used to find interesting transformation formulas. We do the same analysis for the more general Al-Salam–Carlitz polynomials [8]. We apply this approach to provide a generating function for Al-Salam–Carlitz polynomials, generalize Ramanujan’s q-beta integrals and the q-Chu–Vandermonde summation formula. For further information about basic hypergeometric series and q-orthogonal polynomials, see [2, 11, 12, 23].

In this paper, we follow the notations and terminology in [9] and suppose that \(0<q<1\). The basic hypergeometric series \({}_r\phi _s\)

$$\begin{aligned} {}_r\phi _s\biggl [\begin{array}{l} \begin{array}{ccc} a_1,a_2,\ldots ,a_r\\ b_1,b_2,\ldots ,b_s \end{array} \end{array};q,z\biggr ]=\sum _{n=0}^\infty \frac{\bigl (a_1,a_2,\ldots ,a_r;q\bigr )_n}{\bigl (q,b_1,b_2,\ldots ,b_s;q\bigr )_n} \Bigl [(-1)^nq^{n\atopwithdelims ()2}\Bigr ]^{1+s-r}z^n \end{aligned}$$
(1.1)

converges absolutely for all z if \(r\le s\) and for \(\left|z\right|<1\) if \(r=s+1\) and for terminating. The compact factorials of \({}_r\phi _s\) are defined, respectively, by

$$\begin{aligned} (a;q)_0=1,\quad (a;q)_n=\prod _{k=0}^{n-1}(1-aq^k),\quad (a;q)_\infty =\prod _{k=0}^\infty (1-aq^k) \end{aligned}$$
(1.2)

and \((a_1,a_2,\cdots ,a_m;q)_n=(a_1;q)_n(a_2;q)_n\cdots (a_m;q)_n\), where \(m\in \mathbb N:=\{1,2,3,\ldots \}\,\text {and}\, n\in \mathbb N_0:=\mathbb N\cup \{0\}\).

The Rogers–Szegö polynomials were introduced by Szegö in 1926 but were already studied earlier by Rogers in 1894–1895. A good definition can be found in the book by Barry Simon [20, Ex. (1.6.5), pp. 77–87].

The homogeneous Rogers–Szegö polynomials [18, p. 3]

$$\begin{aligned} h_n(b,c\arrowvert q)=\sum _{k=0}^n\left[ \begin{array}{l} n \\ k \\ \end{array}\right] b^kc^{n-k}\quad \text {and}\quad g_n(b,c\arrowvert q)=\sum _{k=0}^n\left[ \begin{array}{l} n \\ k \\ \end{array}\right] q^{k(k-n)}b^kc^{n-k}. \end{aligned}$$
(1.3)

The Al-Salam–Carlitz polynomials were introduced by Al-Salam and Carlitz in 1965 [1, Eqs. (1.11) and (1.15)]

$$\begin{aligned} \phi _n^{(a)}(x\arrowvert q)=\sum _{k=0}^n\left[ \begin{array}{l} n \\ k \\ \end{array}\right] (a;q)_kx^k \quad \text {and}\quad \psi _n^{(a)}(x\arrowvert q)=\sum _{k=0}^n\left[ \begin{array}{l} n \\ k \\ \end{array}\right] q^{k(k-n)}x^k(aq^{1-k};q)_k. \end{aligned}$$
(1.4)

They play important roles in the theory of q-orthogonal polynomials. In fact, there are two families of these polynomials: one with continuous orthogonality and another with discrete orthogonality. They are given explicitly in the book of Koekoek–Swarttouw–Lesky [13, Eqs. (14.24) and (14.25), pp. 534–540].

The generalized Al-Salam–Carlitz polynomials [7, Eq. (4.7)]

$$\begin{aligned} \phi _n^{(a,b,c)}(x,y\arrowvert q)=\sum _{k=0}^n\left[ \begin{array}{l} n \\ k \\ \end{array}\right] \frac{(a,b;q)_k}{(c;q)_k}x^ky^{n-k}\quad \text {and}\nonumber \\ \psi _n^{(a,b,c)}(x,y\arrowvert q)=\sum _{k=0}^n\left[ \begin{array}{l} n \\ k \\ \end{array}\right] \frac{(a,b;q)_k}{(c;q)_k}(-1)^kq^{{{k+1}\atopwithdelims ()2}-nk}x^ky^{n-k}, \end{aligned}$$
(1.5)

whose generating functions are [7, Eqs. (4.10) and (4.11)]

(1.6)
(1.7)

Liu [14, 15] obtained several important results by using the following q-difference equations. Liu and Zeng [17] provide further applications of these q-difference methods to q-orthogonal polynomials.

Proposition 1

([17, Eqs. (1.7) and (1.8)]) Let f(ab) be a two-variable analytic function at \((0,0)\in \mathbb C^2\). Then

  1. (A)

    f can be expanded in terms of \(h_n(a,b\arrowvert q)\) if and only if f satisfies the functional equation

    $$\begin{aligned} bf(aq,b)-af(a,bq)=(b-a)f(a,b). \end{aligned}$$
    (1.8)
  2. (B)

    f can be expanded in terms of \(g_n(a,b\arrowvert q)\) if and only if f satisfies the functional equation

    $$\begin{aligned} af(aq,b)-bf(a,bq)=(a-b)f(aq,bq). \end{aligned}$$
    (1.9)

The method of q-difference equation is an effective way to obtain many results in q-series. For more information, please refer to [6, 7, 14, 15].

Theorem 2

Let f(abcxy) be a five-variable analytic function in a neighbourhood of   \((a,b,c,x,y)=(0,0,0,0,0)\in \mathbb C^5\).

  1. (I)

    If f(abcxy) can be expanded in terms of \(\phi _n^{(a,b,c)}(x,y\arrowvert q)\) if and only if

    $$\begin{aligned}&y\left[ f(a,b,c,x,y)-\Bigl (1+q^{-1}c\Bigr )f(a,b,c,qx,y)+q^{-1}cf\bigl (a,b,c,q^2x,y\bigr )\right] \nonumber \\&\quad =x\Biggl \{\Bigl [f(a,b,c,x,y)-f(a,b,c,x,qy)\Bigr ]\nonumber \\ {}&\qquad \qquad -(a+b)\Bigl [f(a,b,c,qx,y)-f(a,b,c,qx,qy)\Bigr ]\nonumber \\&\qquad \qquad +ab\Bigl [f\bigl (a,b,c,q^2x,y\bigr )-f\bigl (a,b,c,q^2x,qy\bigr )\Bigr ]\Biggr \}. \end{aligned}$$
    (1.10)
  2. (II)

    If f(abcxy) can be expanded in terms of \(\psi _n^{(a,b,c)}(x,y\arrowvert q)\) if and only if

    $$\begin{aligned}&q^{-1}y\left[ f(a,b,c,x,y)-\bigl (1+q^{-1}c\bigr )f\bigl (a,b,c,qx,y\bigr )+q^{-1}cf\bigl (a,b,c,q^2x,y\bigr )\right] \nonumber \\&\quad =x\Biggl \{\Bigl [f(a,b,c,x,y)-f\bigl (a,b,c,x,q^{-1}y\bigr )\Bigr ] \nonumber \\ {}&\qquad \qquad -(a+b)\Bigl [f(a,b,c,qx,y)-f\bigl (a,b,c,qx,q^{-1}y\bigr )\Bigr ]\nonumber \\&\qquad \qquad +ab\Bigl [f\bigl (a,b,c,q^2x,y\bigr )-f\bigl (a,b,c,q^2x,q^{-1}y\bigr )\Bigr ]\Biggr \}. \end{aligned}$$
    (1.11)

Remark 3

For \(a=b=c=0\) in Theorem 2, Eqs. (1.10) and (1.11) reduce to (1.8) and (1.9), respectively.

To determine if a given function is an analytic function in several complex variables, we often use the following Hartogs’s theorem. For more information, please refer to Taylor [22, p. 28] and Liu [16, Theorem 1.8].

Proposition 4

(Hartogs’s theorem [10, p. 15]) If a complex-valued function is holomorphic (analytic) in each variable separately in an open domain \(D\subseteq \mathbb C^n\), then it is holomorphic (analytic) in D.

In order to prove Theorem 2, we need the following fundamental property of several complex variables.

Proposition 5

([19, p. 5, Proposition 1]) If \(f(x_1,x_2,\ldots ,x_k)\) is analytic at the origin \((0,0,\ldots ,0)\in \mathbb C^k\), then, f can be expanded in an absolutely convergent power series,

$$\begin{aligned} f(x_1,x_2,\ldots ,x_k)=\sum _{n_1,n_2,\ldots ,n_k=0}^\infty \alpha _{n_1,n_2,\ldots ,n_k}x_1^{n_1}x_2^{n_2}\cdots x_k^{n_k}. \end{aligned}$$

Proof of Theorem 2

From the Hartogs’s theorem and the theory of several complex variables (see Propositions 4 and 5), we assume that

$$\begin{aligned} f(a,b,c,x,y)=\sum _{k=0}^\infty A_k(a,b,c,y)x^k. \end{aligned}$$
(1.12)

On one hand, substituting Eq. (1.12) into (1.10) yields

$$\begin{aligned}&y\Biggl [\sum _{k=0}^\infty A_k(a,b,c,y)x^k-\Bigl (1+q^{-1}c\Bigr )\sum _{k=0}^\infty A_k(a,b,c,y)(qx)^k\nonumber \\&\qquad +q^{-1}c\sum _{k=0}^\infty A_k(a,b,c,y)\bigl (q^2x\bigr )^k\Biggr ]\nonumber \\&\qquad \quad =x\Biggl \{\Biggl [\sum _{k=0}^\infty A_k(a,b,c,y)x^k-\sum _{k=0}^\infty A_k(a,b,c,qy)x^k\Biggr ]\nonumber \\&\qquad -(a+b)\Biggl [\sum _{k=0}^\infty A_k(a,b,c,y)(qx)^k-\sum _{k=0}^\infty A_k(a,b,c,qy)(qx)^k\Biggr ]\nonumber \\&\qquad +ab\Biggl [\sum _{k=0}^\infty A_k(a,b,c,y)\bigl (q^2x\bigr )^k-\sum _{k=0}^\infty A_k(a,b,c,qy)\bigl (q^2x\bigr )^k\Biggr ]\Biggr \}. \end{aligned}$$
(1.13)

By equating coefficients of \(x^k\) on both sides of Eq. (1.13), we have

$$\begin{aligned} A_k(a,b,c,y)=\frac{\left( 1-aq^{k-1}\right) \left( 1-bq^{k-1}\right) }{\left( 1-q^k\right) \left( 1-cq^{k-1}\right) }D_yA_{k-1}(a,b,c,y). \end{aligned}$$
(1.14)

Iterating, we have

$$\begin{aligned} A_k(a,b,c,y)=\frac{(a,b;q)_k}{(q,c;q)_k}D_y^kA_0(a,b,c,y). \end{aligned}$$
(1.15)

Letting \(f(a,b,c,0,y)=A_0(a,b,c,y)=\sum _{n=0}^\infty \mu _n y^n\), we have

$$\begin{aligned} A_k(a,b,c,y)=\frac{(a,b;q)_k}{(q,c;q)_k}\sum _{n=0}^\infty \mu _n \frac{(q;q)_n}{(q;q)_{n-k}}y^{n-k}. \end{aligned}$$
(1.16)

By Eq. (1.12), we have

$$\begin{aligned}&f(a,b,c,x,y)&=\sum _{k=0}^\infty \frac{(a,b;q)_k}{(q,c;q)_k}\sum _{n=0}^\infty \mu _n \frac{(q;q)_n}{(q;q)_{n-k}}y^{n-k} x^k\\&=\sum _{n=0}^\infty \mu _n\sum _{k=0}^n\left[ \begin{array}{l} n \\ k \\ \end{array}\right] \frac{(a,b;q)_k}{(q,c;q)_k}x^ky^{n-k}=\sum _{n=0}^\infty \mu _n \phi _n^{(a,b,c)}(x,y\arrowvert q). \end{aligned}$$

On the other hand, if f(abcxy) can be expanded in terms of \(\phi _n^{(a,b,c)}(x,y\arrowvert q)\), we can verify that f(abcxy) satisfies Eq. (1.10). The proof of Eq. (1.10) is complete. Similarly, we can deduce Eq. (1.11). The proof of Theorem 2 is complete. \(\square \)

This paper is organized as follows. In Sect. 2, we generalize two generating functions for Andrews–Askey polynomials. In Sect. 3, we deduce generalizations of Ramanujan type q-beta integrals. In Sect. 4, we generalize q-Chu–Vandermonde formula.

2 Two generating functions for generalized Al-Salam–Carlitz polynomials

In this section, we generalize generating functions for Al-Salam–Carlitz polynomials.

Theorem 6

We have

$$\begin{aligned} \sum _{n=0}^\infty \phi _n^{(a,b,c)}(x,y\arrowvert q)\frac{\bigl (s/r;q\bigr )_nr^n}{(q;q)_n}&=\frac{(sy;q)_\infty }{(ry;q)_\infty }{}_3\phi _2\biggl [\begin{array}{l} \begin{array}{lll} a,b,s/r\\ c,sy \end{array} \end{array};q,rx\biggr ],\quad \nonumber \\&\qquad \max \{\left|rx\right|,\left|ry\right|\}<1,\end{aligned}$$
(2.1)
$$\begin{aligned} \sum _{n=0}^\infty \psi _n^{(a,b,c)}(x,y\arrowvert q)\frac{\bigl (r/s;q\bigr )_ns^n}{(q;q)_n}&=\frac{(ry;q)_\infty }{(sy;q)_\infty }{}_3\phi _2\biggl [\begin{array}{l} \begin{array}{lll} a,b,r/s\\ c,q/(sy) \end{array} \end{array};q,\frac{qx}{y}\biggr ],\quad \nonumber \\&\qquad \max \{\left|sy\right|,\left|qx/y\right|\}<1. \end{aligned}$$
(2.2)

Corollary 7

We have

$$\begin{aligned} \sum _{n=0}^\infty \phi _n^{(a,b,c)}(x,y\arrowvert q)(-1)^nq^{n\atopwithdelims ()2}\frac{s^n}{(q;q)_n}&=(sy;q)_\infty {}_2\phi _2\biggl [\begin{array}{l} \begin{array}{lll} a,b\\ c,sy \end{array} \end{array};q,sx\biggr ], \end{aligned}$$
(2.3)
$$\begin{aligned} \sum _{n=0}^\infty \psi _n^{(a,b,c)}(x,y\arrowvert q)\frac{s^n}{(q;q)_n}&=\frac{1}{(sy;q)_\infty }{}_3\phi _2\biggl [\begin{array}{l} \begin{array}{lll} a,b,0\\ c,q/(sy) \end{array} \end{array};q,\frac{qx}{y}\biggr ],\quad \nonumber \\&\qquad \max \{\left|sy\right|,\left|qx/y\right|\}<1. \end{aligned}$$
(2.4)

Remark 8

For \(r=0\) in Theorem 6, Eqs. (2.1) and (2.2) reduce to Eqs. (2.3) and (2.4), respectively. For \(s=0\) in Theorem 6, Eqs. (2.1) and (2.2) reduce to Eqs. (1.6) and (1.7), respectively.

Proof of Theorem 6

By the Weierstrass M-test, series \(\sum _{n=0}^\infty M_n\) is convergent when \(\lim _{n\rightarrow \infty }\left|\frac{M_{n+1}}{M_n}\right|<1\). We check that both sides of Eq. (2.1) are convergent if \(\max \{\left|rx\right|,\left|ry\right|\}<1\), that is,

$$\begin{aligned} \lim _{n\rightarrow \infty }\left|\frac{\phi _{n+1}^{(a,b,c)}(x,y\arrowvert q)(s/r;q)_{n+1}r^{n+1}/(q;q)_{n+1}}{\phi _n^{(a,b,c)}(x,y\arrowvert q)(s/r;q)_nr^n/(q;q)_n}\right|&=\left|ry\right|<1,\\ \lim _{n\rightarrow \infty }\left|\frac{(a,b,s/r;q)_{n+1}(rx)^{n+1}/(q,c,sy;q)_{n+1}}{(a,b,s/r;q)_n(rx)^n/(q,c,sy;q)_n}\right|&=\left|rx\right|<1. \end{aligned}$$

We denote the right-hand side of Eq. (2.1) by f(abcxy), we can verify that f(abcxy) satisfies Eq. (1.10), so we have

$$\begin{aligned} f(a,b,c,x,y)=\sum _{n=0}^\infty \mu _n \phi _n^{(a,b,c)}(x,y\arrowvert q) \end{aligned}$$
(2.5)

and

$$\begin{aligned} f(a,b,c,0,y)=\sum _{n=0}^\infty \mu _n y^n=\frac{(sy;q)_\infty }{(ry;q)_\infty }=\sum _{n=0}^\infty \frac{\bigl (s/r;q\bigr )_n(yr)^n}{(q;q)_n}. \end{aligned}$$
(2.6)

So f(abcxy) is equal to the left-hand side of (2.1). Similarly, we can obtain Eq. (2.2). The proof is complete. \(\square \)

3 Generalizations of two of Ramanujan’s integrals

The following two integrals of Ramanujan [3] are quite famous.

Proposition 9

([3, Eqs. (2) and (3)]) For \(0<q=\exp (-2k^2)<1\) and \(m\in \mathbb R\). Suppose that \(\left|abq\right|<1\), we have

$$\begin{aligned} \int _{-\infty }^{\infty } \frac{e^{-x^2+2mx}}{\left( aq^{1/2}e^{2ikx},bq^{1/2}e^{-2ikx};q\right) _\infty }{{\mathrm{\mathrm{d}}}}x=\sqrt{\pi }e^{m^2}\frac{\left( -aqe^{2mki},-bqe^{-2mki};q\right) _\infty }{(abq;q)_\infty }. \end{aligned}$$
(3.1)

Suppose that \(\max \left\{ \left|aq^{1/2}e^{2mk}\right|,\left|bq^{1/2}e^{-2mk}\right|\right\} <1\), we have

$$\begin{aligned} \int _{-\infty }^{\infty } e^{-x^2+2mx}\left( -aqe^{2kx},-bqe^{-2kx};q\right) _\infty {{\mathrm{\mathrm{d}}}}x\nonumber \\=\sqrt{\pi }e^{m^2}\frac{(abq;q)_\infty }{\left( aq^{1/2}e^{2mk},bq^{1/2}e^{-2mk};q\right) _\infty }. \end{aligned}$$
(3.2)

Derivations of (3.1) and (3.2) for real values of the parameter m have been deduced by Askey [3]. Later on it became clear that these integrals are in fact valid for arbitrary complex values of the parameter m and they are thus instances of the standard Fourier transform with the exponential kernel by Atakishiyev and Feinsilver [5].

In this section, we have the following generalization of Ramanujan’s integrals.

Theorem 10

For \(m\in \mathbb R\), \(0<q=\exp (-2k^2)<1\). Suppose that \(\max \{\left|abq\right|,\left|qc/a\right|\}<1\), we have

$$\begin{aligned}&\int _{-\infty }^{\infty } \frac{e^{-x^2+2mx}}{\Bigl (aq^{1/2}e^{2ikx},bq^{1/2}e^{-2ikx};q\Bigr )_\infty }{}_3\phi _2\left[ \begin{array}{l} \begin{array}{ccc} r,s,0\\ t,q^{1/2}e^{-2ikx}/a \end{array} \end{array};q,\frac{qc}{a}\right] {{\mathrm{\mathrm{d}}}}x\nonumber \\&\quad =\sqrt{\pi }e^{m^2}\frac{\Bigl (-aqe^{2mki},-bqe^{-2mki};q\Bigr )_\infty }{(abq;q)_\infty }{}_3\phi _2\left[ \begin{array}{l} \begin{array}{ccc} r,s,-e^{2mki}/b\\ t,1/(ab) \end{array} \end{array};q,\frac{qc}{a}\right] .\quad \quad \quad \end{aligned}$$
(3.3)

Suppose that \(\max \left\{ \left|aq^{1/2}e^{2mk}\right|,\left|bq^{1/2}e^{-2mk}\right|,\left|cq^{1/2}e^{2mk}\right|\right\} <1\), we have

$$\begin{aligned}&\int _{-\infty }^{\infty } e^{-x^2+2mx}\Bigl (-aqe^{2kx},-bqe^{-2kx};q\Bigr )_\infty {}_2\phi _2\left[ \begin{array}{l} \begin{array}{ccc} r,s\\ t,-aqe^{2kx} \end{array} \end{array};q,-cqe^{2kx}\right] {{\mathrm{\mathrm{d}}}}x\nonumber \\&\quad =\sqrt{\pi }e^{m^2}\frac{(abq;q)_\infty }{\Bigl (aq^{1/2}e^{2mk},bq^{1/2}e^{-2mk};q\Bigr )_\infty }{}_3\phi _2\left[ \begin{array}{l} \begin{array}{lll} r,s,bq^{1/2}e^{-2mk}\\ t,abq \end{array} \end{array};q,cq^{1/2}e^{2mk}\right] .\nonumber \\ \end{aligned}$$
(3.4)

Remark 11

For \(c=0\) in Theorem 10, Eqs. (3.3) and (3.4) reduce to Eqs. (3.1) and (3.2), respectively.

Proof of Theorem 10

It is easily seen that

$$\begin{aligned} \left( \left|q^{\frac{1}{2}}/a\right|;q\right) _n\le \left|\left( q^{\frac{1}{2}}e^{-2ikx}/a;q\right) _n\right| \le \left( -\left|q^{\frac{1}{2}}/a\right|;q\right) _n \end{aligned}$$
(3.5)

and

$$\begin{aligned} \sum _{n=0}^\infty \frac{\bigl (\left|r\right|,\left|s\right|;q\bigr )_n}{\Bigl (-\left|q\right|,-\left|t\right|,-\left|q^{1/2}/a\right|;q\Bigr )_n}\left|\frac{qc}{a}\right|^n\le & {} \left| {}_3\phi _2\biggl [\begin{array}{l} \begin{array}{ccc} r,s,0\\ t,q^{1/2}e^{-2ikx}/a \end{array} \end{array};q,\frac{qc}{a}\biggr ] \right|\nonumber \\\le & {} \sum _{n=0}^\infty \frac{\bigl (-\left|r\right|,-\left|s\right|;q\bigr )_n}{\Bigl (\left|q\right|,\left|t\right|,\left|q^{1/2}/a\right|;q\Bigr )_n}\left|\frac{qc}{a}\right|^n.\nonumber \\ \end{aligned}$$
(3.6)

Thus, we have

$$\begin{aligned}&\left|\int _{-\infty }^{\infty } \frac{e^{-x^2+2mx}}{\Bigl (aq^{1/2}e^{2ikx},bq^{1/2}e^{-2ikx};q\Bigr )_\infty }{{\mathrm{\mathrm{d}}}}x\right|\cdot \sum _{n=0}^\infty \frac{\bigl (\left|r\right|,\left|s\right|;q\bigr )_n}{\Bigl (-\left|q\right|,-\left|t\right|,-\left|q^{1/2}/a\right|;q\Bigr )_n}\left|\frac{qc}{a}\right|^n\nonumber \\&\quad \le \left|\int _{-\infty }^{\infty }\frac{e^{-x^2+2mx}}{\Bigl (aq^{1/2}e^{2ikx},bq^{1/2}e^{-2ikx};q\Bigr )_\infty }{}_3\phi _2\biggl [\begin{array}{l} \begin{array}{lll} r,s,0\\ t,q^{1/2}e^{-2ikx}/a \end{array} \end{array};q,\frac{qc}{a}\biggr ]{{\mathrm{\mathrm{d}}}}x\right|\nonumber \\&\quad \le \left|\int _{-\infty }^{\infty } \frac{e^{-x^2+2mx}}{\Bigl (aq^{1/2}e^{2ikx},bq^{1/2}e^{-2ikx};q\Bigr )_\infty }{{\mathrm{\mathrm{d}}}}x\right| \cdot \sum _{n=0}^\infty \frac{\bigl (-\left|r\right|,-\left|s\right|;q\bigr )_n}{\Bigl (\left|q\right|,\left|t\right|,\left|q^{1/2}/a\right|;q\Bigr )_n}\left|\frac{qc}{a}\right|^n.\nonumber \\ \end{aligned}$$
(3.7)

Denoting the right-hand side of Eq. (3.3) by f(rstca) and utilizing Eqs. (3.1) and (3.7), we have

$$\begin{aligned} \left|f(r,s,t,c,a)\right|&\le \left|\sqrt{\pi }e^{m^2}\frac{\Bigl (-aqe^{2mki},-bqe^{-2mki};q\Bigr )_\infty }{(abq;q)_\infty }\right|\nonumber \\&\quad \times \sum _{n=0}^\infty \frac{\bigl (-\left|r\right|,-\left|s\right|;q\bigr )_n}{\Bigl (\left|q\right|,\left|t\right|,\left|q^{1/2}/a\right|;q\Bigr )_n}\left|\frac{qc}{a}\right|^n\nonumber \\&\le \sqrt{\pi }e^{m^2}\frac{\Bigl (-\left|aq\right|,-\left|bq\right|;q\Bigr )_\infty }{(\left|abq\right|;q)_\infty }\nonumber \\&\quad \times \sum _{n=0}^\infty \frac{\bigl (-\left|r\right|,-\left|s\right|;q\bigr )_n}{\Bigl (\left|q\right|,\left|t\right|,\left|q^{1/2}/a\right|;q\Bigr )_n}\left|\frac{qc}{a}\right|^n. \end{aligned}$$
(3.8)

From the Weierstrass M-test, we know that for \(\max \{\left|abq\right|,\left|qc/a\right|\}<1\), the function f(rstca) is uniformly absolutely convergent, so f(rstca) is an analytic function of rstc and a for \(\max \{\left|abq\right|,\left|qc/a\right|\}<1\) (see also [17], p. 516]). Thus f(rstca) is analytic near \((r,s,t,c,a)=(0,0,0,0,0)\) (see also [4], p. 220] and [17], p. 511]). We can check that f(rstca) satisfies Eq. (1.11), so the left-hand side of Eq. (3.3) equals

$$\begin{aligned} f(r,s,t,c,a)=\sum _{n=0}^\infty \mu _n \psi _n^{(r,s,t)}(c,a\arrowvert q), \end{aligned}$$
(3.9)

where

$$\begin{aligned} f(r,s,t,0,a)&=\sum _{n=0}^\infty \mu _n a^n=\sqrt{\pi }e^{m^2}\frac{\Bigl (-aqe^{2mki},-bqe^{-2mki};q\Bigr )_\infty }{(abq;q)_\infty }\quad by\, (3.1) \\&=\int _{-\infty }^{\infty } \frac{e^{-x^2+2mx}}{\Bigl (aq^{1/2}e^{2ikx},bq^{1/2}e^{-2ikx};q\Bigr )_\infty }{{\mathrm{\mathrm{d}}}}x\\&=\int _{-\infty }^{\infty } \frac{e^{-x^2+2mx}}{\Bigl (bq^{1/2}e^{-2ikx};q\Bigr )_\infty }\left\{ \sum _{n=0}^\infty \frac{\bigl (aq^{1/2}e^{2ikx}\bigr )^n}{(q;q)_n}\right\} {{\mathrm{\mathrm{d}}}}x. \end{aligned}$$

So we have

$$\begin{aligned} f(r,s,t,c,a)=\int _{-\infty }^{\infty } \frac{e^{-x^2+2mx}}{\Bigl (bq^{1/2}e^{-2ikx};q\Bigr )_\infty }\left\{ \sum _{n=0}^\infty \psi _n^{(r,s,t)}(c,a\arrowvert q)\frac{\bigl (q^{1/2}e^{2ikx}\bigr )^n}{(q;q)_n}\right\} {{\mathrm{\mathrm{d}}}}x, \end{aligned}$$
(3.10)

which is equal to the left-hand side of Eq. (3.3) by Eq. (2.4). Similarly, we can gain Eq. (3.4). The proof of Theorem 10 is complete. \(\square \)

4 Generalizations of q-Chu–Vandermonde formula

The q-Chu–Vandermonde formula is [9, Eq. (II.6)]

$$\begin{aligned} {}_2\phi _1\biggl [\begin{array}{l} q^{-n}, a \\ c \end{array} ;q,q\biggr ]=\frac{\bigl (c/a;q\bigr )_n}{(c;q)_n}a^n. \end{aligned}$$
(4.1)

In this section, we now extend the q-Chu–Vandermonde formula.

Theorem 12

For \(n\in \mathbb N_0\), we have

$$\begin{aligned}&\sum _{k=0}^n\frac{\bigl (q^{-n},a;q\bigr )_kq^k}{(q,cd;q)_k}{}_3\phi _2\biggl [\begin{array}{l} \begin{array}{ccc} r,s,aq^k\\ t,qa/(cd) \end{array} \end{array};q,\frac{qg}{d}\biggr ]&\nonumber \\&\quad =\frac{\bigl (cd/a;q\bigr )_n a^n}{(cd;q)_n}{}_3\phi _2\biggl [\begin{array}{l} \begin{array}{ccc} r,s,a\\ t,aq^{1-n}/(cd) \end{array} \end{array};q,\frac{qg}{d}\biggr ],\quad \left|qg/d\right|<1,\end{aligned}$$
(4.2)
$$\begin{aligned}&\sum _{k=0}^n\frac{\bigl (q^{-n},a;q\bigr )_kq^k}{(q,cd;q)_k}{}_3s\phi _2\biggl [\begin{array}{l} \begin{array}{ccc} r,s,aq^k\\ t,cdq^k \end{array} \end{array};q,\frac{cg}{a}\biggr ]&\nonumber \\&\quad =\frac{\bigl (cd/a;q\bigr )_n a^n}{(cd;q)_n}{}_3\phi _2\biggl [\begin{array}{l} \begin{array}{ccc} r,s,aq^n\\ t,cdq^n \end{array} \end{array};q,\frac{cgq^n}{a}\biggr ],\quad \left|cg/a\right|<1. \end{aligned}$$
(4.3)

Remark 13

For \(g=0\) in Theorem 12, Eqs. (4.2) and (4.3) reduce to (4.1), respectively.

Proof of Theorem 12

First, we can rewrite Eq. (4.1) equivalently by

$$\begin{aligned} \sum _{k=0}^n\frac{\bigl (q^{-n},a;q\bigr )_kq^k}{(q;q)_k} \frac{\bigl (cdq^k;q\bigr )_\infty }{\bigl (cd/a;q\bigr )_\infty }=a^n\frac{\bigl (cdq^n;q\bigr )_\infty }{\bigl (cdq^n/a;q\bigr )_\infty }. \end{aligned}$$
(4.4)

We denote the right-hand side of (4.2) by F(rstgd), we can check that F(rstgd) satisfies Eq. (1.11). By Eq. (4.4), we have

$$\begin{aligned} F(r,s,t,g,d)=\sum _{j=0}^\infty \mu _j \psi _j^{(r,s,t)}(g,d\arrowvert q) \end{aligned}$$
(4.5)

and

$$\begin{aligned} F(r,s,t,0,d)&=\sum _{j=0}^\infty \mu _j d^j=a^n\frac{\bigl (cdq^n;q\bigr )_\infty }{\bigl (cdq^n/a;q\bigr )_\infty }\nonumber \\&=\sum _{k=0}^n\frac{\bigl (q^{-n},a;q\bigr )_kq^k}{(q;q)_k}\sum _{j=0}^\infty \frac{\bigl (aq^k;q\bigr )_j\bigl (cd/a\bigr )^j}{(q;q)_j}. \end{aligned}$$

So we have

$$\begin{aligned} F(r,s,t,g,d)=\sum _{k=0}^n\frac{\bigl (q^{-n},a;q\bigr )_kq^k}{(q;q)_k} \sum _{j=0}^\infty \frac{\bigl (aq^k;q\bigr )_j\bigl (c/a\bigr )^j}{(q;q)_j} \psi _j^{(r,s,t)}(g,d\arrowvert q), \end{aligned}$$
(4.6)

which is equal to the left-hand side of (4.2) by Eq. (2.2). Similarly, we can deduce Eq. (4.3). The proof is complete. \(\square \)