1 Introduction

Let p be a fixed odd prime number. Throughout this paper, \(\mathbb {Z}_{p}\), \(\mathbb{Q}_{p}\), and \(\mathbb{C}_{p}\) will, respectively, denote the ring of p-adic rational integers, the field of p-adic rational numbers, and the completions of algebraic closure of \(\mathbb{Q}_{p}\). The p-adic norm is defined by \(|p|_{p} =\frac{1}{p}\).

When one talks of q-extension, q is variously considered as an indeterminate, a complex \(q \in\mathbb{C}\), or p-adic number \(q \in\mathbb{C}_{p}\). If \(q \in\mathbb{C}\), one normally assumes that \(|q|<1\). If \(q \in\mathbb{C}_{p}\), then we assume that \(|q-1|_{p} < p^{- \frac{1}{p-1}}\) so that \(q^{x} = \exp(x\log q)\) for each \(x \in\mathbb{Z}_{p}\). Throughout this paper, we use the notation

$$ [x]_{-q} =\frac{1- (-q)^{x}}{1-(-q)}. $$

Note that \(\lim_{q \rightarrow-1} [x]_{-q} = x \) for each \(x \in \mathbb{Z}_{p}\).

Let \(UD(\mathbb{Z}_{p})\) be the space of uniformly differentiable functions on \(\mathbb{Z}_{p}\). For \(f\in UD({\mathbb{Z}_{p}})\), the p-adic invariant integral on \({\mathbb{Z}_{p}}\) is defined by Kim as follows:

$$ I_{-q}(f)= \int_{\mathbb{Z}_{p}}f(x)\,d\mu_{-q}(x) =\lim_{N \rightarrow\infty} \frac{1}{[p^{N}]_{-q}}\sum _{x=0}^{p^{N}-1} f(x) (-q)^{x}\quad\mbox{(see [1--5])}. $$
(1.1)

Let \(f_{1}\) be the translation of f with \(f_{1} (x )=f (x+1 )\). Then, by (1.1), we get

$$ I_{-q}(f_{1})+I_{-q} (f)=[2]_{q} f(0). $$
(1.2)

As is well known, the Stirling number of the first kind is defined by

$$(x )_{n}=x (x-1 )\cdots (x-n+1 )=\sum _{l=0}^{n}S_{1} (n,l )x^{l}, $$
(1.3)

and the Stirling number of the second kind is given by the generating function:

$$\bigl(e^{t}-1 \bigr)^{m}=m!\sum _{l=m}^{\infty}S_{2} (l,m )\frac {t^{l}}{l!}\quad \mbox{(see [6, 7])}. $$
(1.4)

It is well known that the \((h,q)\) -Euler polynomials are defined by the generating function:

$$ \biggl(\frac{q+1}{q^{h}e^{t}+1} \biggr)e^{xt}=\sum _{n=0}^{\infty}E_{n,q} (x|h)\frac{t^{n}}{n!}\quad\mbox{(see [8])}, $$
(1.5)

where h is an integer. When \(x=0\) and \(h=0\), \(E_{n,q} (0|h) =E_{n,q}(h)\) are called the ordinary q-Euler numbers.

Recently, DS Kim and T Kim introduced the Changhee polynomials of the first kind are defined by the generating function:

$$ \frac{2}{2+t}(1+t)^{x}=\sum _{n=0}^{\infty}Ch_{n}(x)\frac{t^{n}}{n!}\quad\mbox{(see [1, 9--11])}, $$
(1.6)

and T Kim et al. defined the q-Changhee polynomials as follows:

$$ \frac{[2]_{q}}{q(1+t)+1}(1+t)^{x}=\sum _{n=0} ^{\infty} Ch_{n,q}(x)\frac {t^{n}}{n!}\quad\mbox{(see [9, 11, 12])}. $$
(1.7)

As is well known, the Boole polynomials are defined by the generating function:

$$ \sum_{n=0} ^{\infty}Bl_{n}(x|\lambda) \frac{t^{n}}{n!}=\frac {1}{1+(1+t)^{\lambda}}\quad\mbox{(see [7, 13])}. $$

When \(\lambda=1\), \(2Bl_{n}(x|1)=Ch_{n}(x)\) are Changhee polynomials. In [11], Kim et al. consider the q-analog of Boole polynomials, and found some new and interesting identities related to special polynomials, and Y Do and D Lim investigated the properties of \((h,q)\)-Daehee numbers and polynomials, which are defined by

$$ \int _{\mathbb {Z}_{p}}q^{-hy}(x+y)_{n}\,d\mu_{q}(y)\quad\mbox{(see [14])}. $$

In this paper, we consider Witt-type formula for the q-Boole polynomials with weights and derive some new interesting identities and properties of those polynomials and numbers from the Witt-type formula which are related to special polynomials and numbers.

2 q-Analog of Boole polynomials with weight

In this section, we assume that \(t\in{\mathbb{C}}_{p}\) with \(|t|_{p}< p^{-\frac{1}{p-1}}\), \(\lambda\in{\mathbb{Z}}_{p}\) with \(\lambda \neq0\) and \(h\in{\mathbb{Z}}\). From (1.2), we have

$$ \int _{\mathbb {Z}_{p}}q^{(h-1)y}(1+t)^{x+\lambda y}\,d\mu_{-q}(y)=\frac {1+q}{q^{h}(1+t)^{\lambda}+1}(1+t)^{x} =\sum_{n=0} ^{\infty}[2]_{q}Bl_{n,q}(x|h, \lambda)\frac{t^{n}}{n!}, $$
(2.1)

where \(Bl_{n,q} (x|h,\lambda)\) are the \((h,q)\)-Boole polynomials which are defined by

$$ \frac{1}{q^{h}(1+t)^{\lambda}+1}(1+t)^{x}=\sum _{n=0} ^{\infty }Bl_{n,q}(x|h,\lambda) \frac{t^{n}}{n!}. $$
(2.2)

By (2.1), we can derive the following equation:

$$ \int _{\mathbb {Z}_{p}}q^{(h-1)y}\binom{x+\lambda y}{n}\,d\mu_{-q}= \frac{1+q}{n!}Bl_{n,q} (x|h,\lambda). $$
(2.3)

In the special case \(x=0\), \(Bl_{n,q} (0|h,\lambda)=Bl_{n,q}(h,\lambda)\) are called the \((h,q)\) -Boole numbers.

Note that

$$ \begin{aligned}[b] (1+t)^{x+\lambda y}&=e^{(x+\lambda y)\log(1+t)}\\ &=\sum_{n=0} ^{\infty}\frac{(x+\lambda y)^{n}}{n!}\bigl( \log(1+t)\bigr)^{n}\\ &=\sum_{n=0} ^{\infty}\frac{(x+\lambda y)^{n}}{n!}m!\sum _{m=n} ^{\infty }S_{1}(m,n) \frac{t^{m}}{m!}\\ &=\sum_{n=0} ^{\infty} \Biggl\{ \sum _{m=0} ^{n} (x+\lambda y)^{m}S_{1}(n,m) \Biggr\} \frac{t^{n}}{n!}. \end{aligned} $$
(2.4)

The \((h,q)\) -Euler polynomials are defined by the generating function:

$$ \frac{1+q}{q^{h}e^{t}+1}e^{xt}=\sum _{n=0} ^{\infty}E_{n,q}(x|h)\frac{t^{n}}{n!}. $$
(2.5)

Note that \(\lim_{q\rightarrow1}E_{n,q}(x|1)=E_{n}(x)\). When \(x=0\), \(E_{n}(0|h)=E_{n,q}(h)\) are called the \((h,q)\) -Euler numbers.

By (1.2), we can derive easily the following equation:

$$ \int _{\mathbb {Z}_{p}}q^{(h-1)y}e^{(x+y)t}\,d\mu_{-q}(y)=\frac{1+q}{q^{h}e^{t}+1}e^{xt} =\sum_{n=0} ^{\infty}E_{n,q} (x|h) \frac{t^{n}}{n!}. $$
(2.6)

Since

$$ \int _{\mathbb {Z}_{p}}q^{(h-1)y}e^{(x+y)t}\,d\mu_{-q}(y)=\sum _{n=0} ^{\infty} \int _{\mathbb {Z}_{p}}q^{(h-1)y}(x+y)^{n} \,d \mu_{-q}(y)\frac{t^{n}}{n!}, $$

by (2.5), we have

$$ \int _{\mathbb {Z}_{p}}q^{(h-1)y}(x+y)^{n}\,d\mu_{-q}(y)=E_{n,q} (x|h) \quad(n\geq0). $$
(2.7)

From (2.4) and (2.7), we get

$$ \begin{aligned}[b] &\int _{\mathbb {Z}_{p}}q^{(h-1)y}(1+t)^{x+\lambda y}\,d\mu_{-q}(y)\\ &\quad=\sum_{n=0} ^{\infty} \Biggl\{ \sum _{m=0} ^{n} \int _{\mathbb {Z}_{p}}q^{(h-1)y}(x+\lambda y)^{m} \,d\mu_{-q}(y) S_{1}(n,m) \Biggr\} \frac{t^{n}}{n!}\\ &\quad=\sum_{n=0} ^{\infty} \Biggl\{ \sum _{m=0} ^{n}\lambda^{m} E_{m,q} \biggl(\frac{x}{\lambda} \vert h \biggr)S_{1}(n,m) \Biggr\} \frac{t^{n}}{n!}. \end{aligned} $$
(2.8)

Thus, by (2.2), (2.3), and (2.8), we obtain the following theorem.

Theorem 2.1

For \(n \geq0\), we have

$$ Bl_{n,q}(x|h,\lambda)=\frac{1}{[2]_{q}}\sum _{m=0} ^{n} \lambda^{m}E_{m,q} \biggl(\frac{x}{\lambda}\Big| h \biggr)S_{1}(n,m) $$

and

$$ \int _{\mathbb {Z}_{p}}q^{(h-1)y}\binom{x+\lambda y}{n}\,d\mu_{-q}=\frac{[2]_{q}}{n!}Bl_{n,q} (x|h,\lambda). $$

By Theorem 2.1, we note that

$$ Bl_{n,q}(x|h,\lambda)=\frac{1}{[2]_{q}}\int _{\mathbb {Z}_{p}}q^{(h-1)y}(x+\lambda y)_{n}\,d\mu_{-q}(y), $$

where \((x)_{n}=x(x-1)\cdots(x-n+1)\). When \(\lambda=1\) and \(h=0\), we have

$$ Bl_{n,q}(x|0,1)=\frac{1}{[2]_{q}}\int _{\mathbb {Z}_{p}}q^{-1}(x+y)^{n}\,d\mu_{-q}(y). $$
(2.9)

In [13], Arici et al. defined the q-analog of Changhee polynomials by the generating function:

$$ \sum_{n=0} ^{\infty}Ch_{n}(x|q) \frac{t^{n}}{n!}=\frac{[2]_{q}}{[2]_{t}+1}(1+t)^{x}. $$
(2.10)

By (2.10), we have

$$ \int _{\mathbb {Z}_{p}}q^{-y}(1+t)^{x+y}\,d\mu_{-q}(y)= \frac{[2]_{q}}{[2]_{t}+1}(1+t)^{x}=\sum_{n=0} ^{\infty}Ch_{n}(x|q)\frac{t^{n}}{n!}. $$
(2.11)

By (1.6) and (2.10), we note that

$$ \frac{[2]_{q}}{2}Ch_{n}(x)=Ch_{n}(x|q). $$
(2.12)

From (2.11), we get

$$ \int _{\mathbb {Z}_{p}}q^{-1}(x+y)_{n}\,d\mu_{-q}(y)=Ch_{n}(x|q). $$
(2.13)

By (2.9), (2.12), and (2.13), we have

$$ Bl_{n,q} (x|0,1)=\frac{1}{[2]_{q}}Ch_{n}(x|q)= \frac{1}{2}Ch_{n}(x). $$

By replacing t as \(e^{t}-1\) in (2.1), we derive the following equations:

$$ \begin{aligned}[b] \frac{1+q}{q^{h}e^{\lambda t}+1}e^{xt}&=\sum _{n=0} ^{\infty }[2]_{q}Bl_{n,q}(x|h, \lambda)\frac{1}{n!}\bigl(e^{t}-1\bigr)^{n}\\ &=\sum_{n=0} ^{\infty}[2]_{q}Bl_{n,q}(x|h, \lambda)\frac{1}{n!}n!\sum_{m=n} ^{\infty}S_{2}(m,n)\frac{t^{m}}{m!}\\ &=\sum_{n=0} ^{\infty}\sum _{m=0} ^{n} [2]_{q}Bl_{m,q}(x|h, \lambda )S_{2}(n,m)\frac{t^{n}}{n!} \end{aligned} $$
(2.14)

and

$$ \frac{1+q}{q^{h}e^{\lambda t}+1}e^{xt}= \frac{1+q}{q^{h}e^{\lambda t}+1}e^{ (\frac{x}{\lambda} )\lambda t} =\sum_{n=0} ^{\infty}E_{n,q} \biggl( \frac{x}{\lambda}\Big| h \biggr)\lambda^{m}\frac{t^{m}}{m!}. $$
(2.15)

Hence, by (2.14) and (2.15), we obtain the following theorem.

Theorem 2.2

For \(n \geq0\), we have

$$ \sum_{m=0} ^{n}Bl_{m,q}(x|h, \lambda)S_{2}(n,m)=\frac{\lambda ^{m}}{q+1}E_{n,q} \biggl( \frac{x}{\lambda}\Big| h \biggr). $$

From now on, we define the \((h_{1},\ldots,h_{r},q)\) -Boole numbers of the first kind as follows:

$$ \begin{aligned}[b] &[2]_{q} ^{r} Bl_{n,q} ^{(h_{1},\ldots,h_{r})} (\lambda)\\ &\quad=\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}\bigl(\lambda(x_{1}+ \cdots+x_{r})\bigr)_{n}\,d\mu _{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}) \quad(n \geq0). \end{aligned} $$
(2.16)

By (2.16), we have

$$\begin{aligned} &[2]_{q} ^{r} \sum _{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})} (\lambda) \frac{t^{n}}{n!} \\ &\quad=\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}\sum_{n=0} ^{\infty} q^{h_{1}+\cdots+h_{r}-r}\binom{\lambda (x_{1}+\cdots+x_{r})}{n}t^{n}\,d\mu_{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}) \\ &\quad=\sum_{n=0} ^{\infty} \int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}(1+t)^{\lambda (x_{1}+\cdots+x_{k})}\,d\mu_{-q}(x_{1})\cdots \,d \mu_{-q}(x_{r}) \\ &\quad=\prod_{i=1} ^{r} \biggl( \frac{1+q}{q^{h_{i}}(1+t)^{\lambda}+1} \biggr) \\ &\quad=(1+q)^{r}\sum_{n=0} ^{\infty} \biggl(\sum_{l_{1}+\cdots+l_{r}=n}\binom {n}{l_{1},\ldots,l_{r}}B_{i_{1},q} (h, \lambda)\cdots B_{i_{r},q} (h,\lambda ) \biggr)\frac{t^{n}}{n!}. \end{aligned}$$
(2.17)

Thus, by (2.17), we obtain the following corollary.

Corollary 2.3

For \(n \geq0\), we have

$$ Bl_{n,q} ^{(h_{1},\ldots,h_{r})} (\lambda)=\sum_{l_{1}+\cdots+l_{r}=n} \binom {n}{l_{1},\ldots,l_{r}}B_{i_{1},q} (h,\lambda)\cdots B_{i_{r},q} (h,\lambda). $$

The \((h_{1},\ldots,h_{r},q)\) -Euler polynomials are defined by the generating function to be

$$\begin{aligned} &\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}e^{(x_{1}+\cdots+x_{r}+x)t}\,d\mu _{-q}(x_{1})\cdots \,d\mu_{-q}(x_{r}) \\ &\quad=\prod_{i=1} ^{r} \biggl( \frac{1+q}{q^{h_{i}}e^{t}+1} \biggr)e^{xt} \\ &\quad=\sum_{n=0} ^{\infty}E_{n,q}(x|h_{1},\ldots,h_{r}) \frac{t^{n}}{n!}. \end{aligned}$$
(2.18)

By (2.18), we have

$$ \int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}(x_{1}+\cdots+x_{r}+x)^{n}\,d \mu _{-q}(x_{1})\cdots \,d\mu_{-q}(x_{r})=E_{n,q} (x|h_{1},\ldots,h_{r}). $$

In the special case \(x=0\), \(E_{n,q} (0|h_{1},\ldots,h_{r})=E_{n,q} (h_{1},\ldots,h_{r})\) are called the \((h_{1},\ldots,h_{r},q)\) -Euler numbers.

From (1.5) and (2.16), we note that

$$\begin{aligned} &(1+q)^{r}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(\lambda) \\ &\quad=\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}\bigl(\lambda(x_{1}+ \cdots+x_{r})\bigr)_{n}\,d\mu _{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}) \\ &\quad=\sum_{l=0} ^{n} S_{1}(n,l)\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}\lambda ^{l}(x_{1}+ \cdots+x_{r})^{l}\,d\mu_{-q}(x_{1})\cdots \,d \mu_{-q}(x_{r}) \\ &\quad=\sum_{l=0} ^{n} S_{1}(n,l) \lambda^{l}E_{l,q}(h_{1},\ldots,h_{r}). \end{aligned}$$
(2.19)

Therefore, by (2.19), we obtain the following theorem.

Theorem 2.4

For \(n \geq0\), we get

$$ Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(\lambda)=\frac{1}{(1+q)^{r}}\sum _{l=0} ^{n} S_{1}(n,l)\lambda^{l}E_{l,q}(h_{1}, \ldots,h_{r}). $$

By replacing t by \(e^{t}-1\) in (2.17), we have

$$\begin{aligned}{} [2]_{q} ^{r}\sum _{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(\lambda)\frac {(e^{t}-1)^{n}}{n!}&=\prod_{i=1} ^{r} \biggl(\frac{1+q}{q^{h_{i}}e^{\lambda t}+1} \biggr) \\ &=\sum_{n=0} ^{\infty}E_{n,q} (h_{1},\ldots,h_{r})\lambda^{n}\frac{t^{n}}{n!} \end{aligned}$$
(2.20)

and

$$\begin{aligned}{} [2]_{q} ^{r}\sum _{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(\lambda)\frac {1}{n!}\bigl(e^{t}-1 \bigr)^{n} &=[2]_{q} ^{r}\sum_{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(\lambda)\sum _{m=n} ^{\infty}S_{2}(m,n)\frac{t^{m}}{m!} \\ &=[2]_{q} ^{r}\sum_{n=0} ^{\infty} \Biggl\{ \sum_{m=0} ^{n} Bl_{m,q} ^{(h_{1},\ldots,h_{r})}(\lambda)S_{2}(n,m) \Biggr\} \frac{t^{n}}{n!}. \end{aligned}$$
(2.21)

Hence, by (2.20) and (2.21), we obtain the following theorem.

Theorem 2.5

For \(n \geq0\), we have

$$ \frac{\lambda^{n}}{[2]_{q} ^{r}}E_{n,q} (h_{1},\ldots,h_{r})= \sum_{m=0} ^{n} Bl_{m,q} ^{(h_{1},\ldots,h_{r})}(\lambda)S_{2}(n,m). $$

Let us define the \((h_{1},\ldots,h_{r},q)\) -Boole polynomials of the first kind as follows:

$$\begin{aligned} &[2]_{q} ^{r}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda) \\ &\quad=\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}\bigl(\lambda(x_{1}+ \cdots+x_{r})+x\bigr)_{n}\,d\mu _{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}), \end{aligned}$$
(2.22)

where \(n \geq0\) and \(r\in{\mathbb{N}}\). By (2.22), we can derive the generating function of the \((h_{1},\ldots ,h_{r},q)\)-Boole polynomials of the first kind as follows:

$$\begin{aligned} &[2]_{q} ^{r}\sum _{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda )\frac{t^{n}}{n!} \\ &\quad=\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}(1+t)^{\lambda(x_{1}+\cdots+x_{r})+x}\,d\mu _{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}) \\ &\quad=\prod_{i=1} ^{r} \biggl( \frac{1+q}{q^{h_{i}}(1+t)^{\lambda}+1} \biggr) (1+t)^{x}. \end{aligned}$$
(2.23)

By (2.23), we can see easily

$$\begin{aligned} &\prod_{i=1} ^{r} \biggl(\frac{1+q}{q^{h_{i}}(1+t)^{\lambda}+1} \biggr) (1+t)^{x} \\ &\quad=[2]_{q} ^{r} \Biggl(\sum_{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(\lambda )\frac{t^{n}}{n!} \Biggr) \Biggl(\sum_{m=0} ^{\infty} \binom{x}{m}t^{m} \Biggr) \\ &\quad=[2]_{q} ^{r} \sum_{n=0} ^{\infty} \Biggl(\sum_{m=0} ^{n} m! \binom{x}{m}\frac {n!}{(n-m)!m!}Bl_{n-m,q} ^{(h_{1},\ldots,h_{r})}(\lambda) \Biggr)\frac {t^{n}}{n!} \\ &\quad=[2]_{q} ^{r} \sum_{n=0} ^{\infty} \Biggl(\sum_{m=0} ^{n} m! \binom{x}{m}\binom {n}{m}Bl_{n-m,q} ^{(h_{1},\ldots,h_{r})}(\lambda) \Biggr)\frac{t^{n}}{n!} \\ &\quad=[2]_{q} ^{r} \sum_{n=0} ^{\infty} \Biggl(\sum_{m=0} ^{n} \binom {n}{m}Bl_{n-m,q} ^{(h_{1},\ldots,h_{r})}(\lambda) (x)_{m} \Biggr)\frac{t^{n}}{n!}. \end{aligned}$$
(2.24)

By (2.23) and (2.24), we obtain the following theorem.

Theorem 2.6

For \(n \geq0\), we have

$$ Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda)=\sum_{m=0} ^{n} \binom {n}{m}Bl_{n-m,q} ^{(h_{1},\ldots,h_{r})}(\lambda) (x)_{m}. $$

Replacing t as \(e^{t}-1\) in (2.23), we get

$$\begin{aligned}{} [2]_{q} ^{r}\sum _{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda)\frac {1}{n!}\bigl(e^{t}-1 \bigr)^{n}&=\prod_{i=1} ^{n} \biggl(\frac{1+q}{q^{h_{i}}e^{\lambda t}+1} \biggr)e^{xt} \\ &=\sum_{n=0} ^{\infty}E_{n,q} ^{(h_{1},\ldots,h_{r})} \biggl(\frac{x}{\lambda } \biggr)\lambda^{n} \frac{t^{n}}{n!} \end{aligned}$$
(2.25)

and

$$\begin{aligned} &[2]_{q} ^{r}\sum _{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda )\frac{(e^{t}-1)^{n}}{n!} \\ &\quad=[2]_{q} ^{r}\sum_{n=0} ^{\infty} \Biggl(\sum_{m=0} ^{n} Bl_{m,q} ^{(h_{1},\ldots ,h_{r})}(x|\lambda)S_{2}(n,m) \Biggr) \frac{t^{n}}{n!}. \end{aligned}$$
(2.26)

Hence, by (2.25) and (2.26), we obtain the following theorem.

Theorem 2.7

For \(n \geq0\), we have

$$ \sum_{m=0} ^{n} Bl_{m,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda)S_{2}(n,m)=\frac {\lambda^{n}}{[2]_{q} ^{r}}E_{n,q} ^{(h_{1},\ldots,h_{r})} \biggl(\frac{x}{\lambda } \biggr). $$

From (2.23), we get

$$\begin{aligned} &[2]_{q} ^{r}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda) \\ &\quad=\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}\bigl(\lambda(x_{1}+ \cdots+x_{r})+x\bigr)_{n}\,d\mu _{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}) \\ &\quad=\sum_{l=0} ^{n} S_{1}(n,l) \int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{h_{1}+\cdots+h_{r}-r}\bigl(\lambda(x_{1}+\cdots +x_{r})+x\bigr)^{l}\,d\mu_{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}) \\ &\quad=\sum_{l=0} ^{n} S_{1}(n,l) \lambda^{l}E_{n,q} ^{(h_{1},\ldots,h_{r})} \biggl(\frac {x}{\lambda} \biggr). \end{aligned}$$
(2.27)

Thus, by (2.27), we obtain the following theorem.

Theorem 2.8

For \(n\geq0\), we have

$$ Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda)=\frac{1}{[2]_{q} ^{r}}\sum _{l=0} ^{n} S_{1}(n,l)\lambda^{l}E_{n,q} ^{(h_{1},\ldots,h_{r})} \biggl(\frac{x}{\lambda } \biggr). $$

Now, we define the \((h,q)\) -Boole polynomials of the second kind as follows:

$$ {\widehat{Bl}}_{n,q}(x|h,\lambda)=\frac{1}{[2]_{q}}\int _{\mathbb {Z}_{p}}q^{(h-1)y}(-\lambda y+x)_{n}\,d\mu_{-q}(y) \quad(n \geq0). $$
(2.28)

By (2.28), we have

$$\begin{aligned} {\widehat{Bl}}_{n,q}(x|h, \lambda)&=\frac{1}{[2]_{q}}\sum_{l=0} ^{n}(- \lambda )^{l} S_{1}(n,l)\int _{\mathbb {Z}_{p}}\biggl(y-\frac{x}{\lambda} \biggr)^{l}\,d\mu_{-q}(y) \\ &=\frac{1}{[2]_{q}}\sum_{l=0} ^{n}(- \lambda)^{l} S_{1}(n,l)E_{l,q} \biggl(- \frac {x}{\lambda} \biggr). \end{aligned}$$
(2.29)

In the special case \(x=0\), \({\widehat{Bl}}_{n,q} (0|h,\lambda)={\widehat {Bl}}_{n,q}(h,\lambda)\) are called the \((h,q)\) -Boole numbers of the second kind. From (2.29), we can derive the generating function of \({\widehat{Bl}}_{n,q}(x|\lambda)\) as follows:

$$\begin{aligned} \sum_{n=0} ^{\infty}{\widehat{Bl}}_{n,q}(x|h,\lambda)\frac {t^{n}}{n!}&= \frac{1}{[2]_{q}}\int _{\mathbb {Z}_{p}}q^{(h-1)y}(1+t)^{-\lambda y+x}\,d\mu _{-q}(y) \\ &=\frac{(1+t)^{\lambda}}{q^{h}+(1+t)^{\lambda}}(1+t)^{x}. \end{aligned}$$
(2.30)

By replacing t by \(e^{t}-1\) in (2.30), we have

$$\begin{aligned} \sum_{n=0} ^{\infty}{\widehat{Bl}}_{n,q}(x|h,\lambda)\frac {(e^{t}-1)^{n}}{n!}&= \frac{e^{\lambda t}}{q^{h}+e^{\lambda t}}e^{xt} \\ &=\frac{1}{1+q}\sum_{n=0} ^{\infty}(- \lambda)^{n}E_{n,q} \biggl(-\frac {\lambda}{x}\Big| h \biggr)\frac{t^{n}}{n!} \end{aligned}$$
(2.31)

and

$$ \sum_{n=0} ^{\infty}{ \widehat{Bl}}_{n,q}(x|h,\lambda)\frac{(e^{t}-1)^{n}}{n!} =\sum _{n=0} ^{\infty} \Biggl(\sum_{m=0} ^{n} {\widehat {Bl}}_{m,q}(x|h,\lambda)S_{2}(n,m) \Biggr)\frac{t^{n}}{n!}. $$
(2.32)

By (2.31) and (2.32), we obtain the following theorem.

Theorem 2.9

For \(n \geq0\), we have

$$ {\widehat{Bl}}_{n,q}(x|h,\lambda)=\frac{1}{[2]_{q}}\sum _{l=0} ^{n}(-\lambda )^{l} S_{1}(n,l)E_{l,q} \biggl(-\frac{x}{\lambda} \biggr) $$

and

$$ \frac{1}{[2]_{q}}(-\lambda)^{n}E_{n,q} \biggl(- \frac{\lambda}{x}\Big| h \biggr)=\sum_{m=0} ^{n} {\widehat{Bl}}_{m,q}(x|h,\lambda)S_{2}(n,m). $$

For \(h_{1},\ldots,h_{r}\in{\mathbb{Z}}\), we define the \((h_{1},\ldots ,h_{r},q)\) -Boole polynomials of the second kind as follows:

$$\begin{aligned} &{\widehat{Bl}}_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda) \\ &\quad=\frac{1}{q+1}\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}q^{(h_{1}+\cdots+h_{r}-r)y} \bigl(-\lambda (x_{1}+\cdots+x_{r})+x \bigr)_{n}\,d \mu_{-q}(x_{1})\cdots \,d\mu_{-q}(x_{r}). \end{aligned}$$
(2.33)

By (2.33), we can derive the generating function of the \((h_{1},\ldots ,h_{r},q)\)-Boole polynomials of the second kind as follows:

$$\begin{aligned} &\sum_{n=0} ^{\infty}{\widehat{Bl}}_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda ) \frac{t^{n}}{n!} \\ &\quad=\frac{1}{(1+q)^{r}}\int _{\mathbb {Z}_{p}}\cdots \int _{\mathbb {Z}_{p}}(1+t)^{-\lambda x_{1}-\cdots-\lambda x_{r}+x}\,d\mu_{-q}(x_{1}) \cdots \,d\mu_{-q}(x_{r}) \\ &\quad=\prod_{i=1} ^{r} \biggl( \frac{(1+t)^{\lambda}}{q^{h_{i}}+(1+t)^{\lambda }} \biggr) (1+t)^{x} \\ &\quad=\prod_{i=1} ^{r} \biggl( \frac{1}{q^{h_{i}}(1+t)^{-\lambda}+1} \biggr) (1+t)^{x} \\ &\quad=\sum_{n=0} ^{\infty}Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|-\lambda)\frac{t^{n}}{n!}. \end{aligned}$$
(2.34)

Hence, by (2.34), we obtain the following proposition.

Proposition 2.10

For \(n \geq0\), we have

$$ {\widehat{Bl}}_{n,q} ^{(h_{1},\ldots,h_{r})}(x|\lambda)=Bl_{n,q} ^{(h_{1},\ldots,h_{r})}(x|-\lambda). $$

Note that

$$\begin{aligned} \frac{(-1)^{n}[2]_{q}}{n!}Bl_{n,q} (x|h, \lambda)&=(-1)^{n}\int _{\mathbb {Z}_{p}}q^{(h-1)y}\binom {x+\lambda y}{n}\,d\mu_{-q}(y) \\ &=\int _{\mathbb {Z}_{p}}q^{(h-1)y}\binom{-x-\lambda y+n-1}{n}\,d\mu_{-q}(y) \\ &=\int _{\mathbb {Z}_{p}}q^{(h-1)y}\sum_{m=0} ^{n} \binom{-x-\lambda y}{m}\binom{n-1}{n-m}\,d\mu _{-q}(y) \\ &=\sum_{m=0} ^{n} \binom{n-1}{n-m}\int _{\mathbb {Z}_{p}}q^{(h-1)y}\binom{-x-\lambda y}{m}\,d\mu_{-q}(y) \\ &=[2]_{q}\sum_{m=0} ^{n} \binom{n-1}{n-m}\frac{{\widehat{Bl}}_{m,q} (-x|h,\lambda)}{m!}, \end{aligned}$$
(2.35)

and, by a similar method, we get

$$\begin{aligned} \frac{(-1)^{n}[2]_{q}}{n!}{\widehat{Bl}}_{n,q} (x|h,\lambda)&=(-1)^{n}\int _{\mathbb {Z}_{p}}q^{(h-1)y}\binom{x-\lambda y}{n}\,d\mu_{-q}(y) \\ &=[2]_{q}\sum_{m=0} ^{n} \binom{n-1}{n-m}\frac{Bl_{m,q} (-x|h,\lambda)}{m!}. \end{aligned}$$
(2.36)

By (2.35) and (2.36), we obtain the following theorem.

Theorem 2.11

For \(n \geq0\), we have

$$ \frac{(-1)^{n}}{n!}Bl_{n,q} (x|h,\lambda)=\sum _{m=0} ^{n} \binom {n-1}{n-m}\frac{{\widehat{Bl}}_{m,q} (-x|h,\lambda)}{m!} $$

and

$$ \frac{(-1)^{n}}{n!}{\widehat{Bl}}_{n,q} (x|h,\lambda)=\sum _{m=0} ^{n} \binom {n-1}{n-m}\frac{Bl_{m,q} (-x|h,\lambda)}{m!}. $$

By Theorem 2.11, we obtain the following corollary.

Corollary 2.12

For \(n \geq0\), we have

$$ Bl_{n,q} (x|h,\lambda)=\sum_{m=0} ^{n} \sum_{k=0} ^{m} (-1)^{n+m}\binom {n}{n-m,m-k,k}(n-1)_{l-1}Bl_{k,q}(x|h, \lambda) $$

where \(\binom{n}{p,q,r}=\frac{n!}{p!q!r!}\), \(p+q+r=n\).