\documentclass[12pt,reqno]{article} \usepackage[usenames]{color} \usepackage{amssymb} \usepackage{graphicx} \usepackage{amscd} \usepackage[colorlinks=true, linkcolor=webgreen, filecolor=webbrown, citecolor=webgreen]{hyperref} \definecolor{webgreen}{rgb}{0,.5,0} \definecolor{webbrown}{rgb}{.6,0,0} \usepackage{color} \usepackage{fullpage} \usepackage{float} \usepackage{psfig} \usepackage{graphics,amsmath,amssymb} \usepackage{amsthm} \usepackage{amsfonts} \usepackage{latexsym} \usepackage{epsf} \usepackage{mathtools} \setlength{\textwidth}{6.5in} \setlength{\oddsidemargin}{.1in} \setlength{\evensidemargin}{.1in} \setlength{\topmargin}{-.1in} \setlength{\textheight}{8.4in} \newcommand{\seqnum}[1]{\href{http://oeis.org/#1}{\underline{#1}}} \begin{document} \begin{center} \epsfxsize=4in \leavevmode\epsffile{logo129.eps} \end{center} \theoremstyle{plain} \newtheorem{theorem}{Theorem} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \theoremstyle{definition} \newtheorem{definition}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \newtheorem{conjecture}[theorem]{Conjecture} \theoremstyle{remark} \newtheorem{remark}[theorem]{Remark} \begin{center} \vskip 1cm{\LARGE\bf On $q$-Boson Operators and $q$-Analogues of the\\ \vskip .1in $r$-Whitney and $r$-Dowling Numbers} \vskip 1cm \large Mahid M. Mangontarum\\ Department of Mathematics\\ Mindanao State University -- Main Campus\\ Marawi City 9700\\ Philippines \\ \href{mailto:mmangontarum@yahoo.com}{\tt mmangontarum@yahoo.com} \\ \href{mailto:mangontarum.mahid@msumain.edu.ph}{\tt mangontarum.mahid@msumain.edu.ph} \\ \ \\ Jacob Katriel\\ Department of Chemistry\\ Technion -- Israel Institute of Technology\\ Haifa 32000\\ Israel\\ \href{mailto:jkatriel@technion.ac.il}{\tt jkatriel@technion.ac.il}\\ \end{center} \vskip .2 in \begin{abstract} We define the $(q,r)$-Whitney numbers of the first and second kinds in terms of the $q$-Boson operators, and obtain several fundamental properties such as recurrence formulas, orthogonality and inverse relations, and other interesting identities. As a special case, we obtain a $q$-analogue of the $r$-Stirling numbers of the first and second kinds. Finally, we define the $(q,r)$-Dowling polynomials in terms of sums of $(q,r)$-Whitney numbers of the second kind, and obtain some of their properties. \end{abstract} \newcommand{\fstirling}[2]{\genfrac[]{0pt}{}{#1}{#2}} \newcommand{\sstirling}[2]{\genfrac\{\}{0pt}{}{#1}{#2}} \newcommand{\NN}{\mathbb{N}} \newcommand{\ZZ}{\mathbb{Z}} \newcommand{\QQ}{\mathbb{Q}} \newcommand{\im}{\mbox{\upshape Im}} \newcommand{\stirlings}[2]{\genfrac\{\}{0pt}{}{#1}{#2}} \newcommand{\stirlingf}[2]{\genfrac[]{0pt}{}{#1}{#2}} \newcommand{\rbstirling}[2]{\genfrac\langle\rangle{0pt}{}{#1}{#2}} \newcommand{\braket}[2]{\langle{#1}\lvert{#2}\rangle} \DeclarePairedDelimiter\bra{\langle}{\rvert} \DeclarePairedDelimiter\ket{\lvert}{\rangle} \section{Introduction} \label{sec:Introduction} The investigation of $q$-analogues of combinatorial identities has proven to be a rich source of insight as well as of useful generalizations. Some examples of $q$-analogues are the \textit{$q$-real number}, the \textit{$q$-factorial} and the \textit{$q$-falling factorial of order $r$}, respectively, given by \begin{equation*} [x]_q=\frac{q^x-1}{q-1},\ [n]_q!=\prod_{i=1}^n[i]_q,\ [x]_{q,n}=\prod_{i=0}^{r-1}[x-i]_q, \end{equation*} for any real number $x$ and non-negative integers $n$ and $r$, and the \textit{$q$-binomial coefficients (also known as Gaussian polynomials)} \begin{equation*} \binom{n}{r}_q=\frac{[n]_q!}{[r]_q![n-r]_q!}=\frac{[n]_{q,r}}{[r]_q!}. \end{equation*} The formulation of $q$-analogues is not unique, but some choices appear to allow more productive generalizations than others. In the present paper we apply the properties of the $q$-boson operators as a framework for the generation of $q$-deformations of a family of combinatorial identities involving the Whitney numbers. A lattice $L$ in which every element is the join of elements $x$ and $y$ (in $L$) such that $x$ and $y$ cover the zero element $0$, and is semimodular, is called a \textit{geometric lattice}. Originally, if $L$ is a finite lattice of rank $n$, then the Whitney numbers $w(n,k)$ and $W(n,k)$ of the first and second kinds of $L$ are defined as the coefficients of the characteristic polynomial and as the number of elements of $L$ of corank $k$, respectively. Now, Dowling \cite{Dowl} defined a class of these geometric lattices, called \textit{Dowling lattice}, which is a generalization of the partition lattice. Let $Q_n(G)$ be the Dowling lattice of rank $n$ associated to a finite group $G$ of order $m>0$. Benoumhani \cite{Ben1} defined the \textit{Whitney numbers of the first and second kind of $Q_n(G)$}, denoted by $w_m(n,k)$ and $W_m(n,k)$, respectively, in terms of the relations \begin{equation} \label{w01} m^n(x)_n=\sum_{k=0}^{n}w_m(n,k)(mx+1)^k \end{equation} and \begin{equation} \label{W02} (mx+1)^n=\sum_{k=0}^{n}m^kW_m(n,k)(x)_k, \end{equation} where $(x)_n=x(x-1)\cdots(x-n+1)$ is the falling factorial of $x$ of order $n$. Notice that if the group $G$ is the trivial group ($m=1$), multiplication of both equations \eqref{w01} and \eqref{W02} by $(x+1)$ yields the horizontal generating functions for the well-known Stirling numbers of the first and second kind \cite{Stir}, denoted by $\fstirling{n}{k}$ and $\sstirling{n}{k}$, respectively. Hence, \begin{equation*} w_1(n,k)=\fstirling{n+1}{k+1},\ W_1(n,k)=\sstirling{n+1}{k+1}. \end{equation*} We note that Benoumhani \cite{Ben1,Ben2} already established the fundamental properties of the numbers $w_m(n,k)$ and $W_m(n,k)$ while Dowling \cite{Dowl} gave a detailed discussion of geometric lattices. Other generalizations of the Stirling numbers $\fstirling{n}{k}$ and $\sstirling{n}{k}$ were already considered by several authors. For instance, Broder \cite{Bro} defined the $r$-Stirling numbers $\widehat{\fstirling{n+r}{k+r}}_r$ and $\widehat{\sstirling{n+r}{k+r}}_r$ of the first and second kind whose relation to the Whitney numbers is stated in equations \eqref{w7} and \eqref{w8} below. Belbachir and Bousbaa \cite{Bel} recently introduced the translated Whitney numbers $\widetilde{w}_{(\alpha)}(n,k)$ and $\widetilde{W}_{(\alpha)}(n,k)$ of the first and second kind, which are related to the Stirling numbers via \begin{equation*} \widetilde{w}_{(\alpha)}(n,k)=\alpha^{n-k}\fstirling{n}{k},\ \widetilde{W}_{(\alpha)}(n,k)=\alpha^{n-k}\sstirling{n}{k}. \end{equation*} Furthermore, Mez\H{o} \cite{Mez1} defined the $r$-Whitney numbers $w_{m,r}(n,k)$ and $W_{m,r}(n,k)$ of the first and second kind as the coefficients in the expressions \begin{equation} m^n(x)_n=\sum_{k=0}^{n}w_{m,r}(n,k)(mx+r)^k\label{w1} \end{equation} and \begin{equation} (mx+r)^n=\sum_{k=0}^{n}m^kW_{m,r}(n,k)(x)_k.\label{w2} \end{equation} respectively. Further developement of the numbers $w_{m,r}(n,k)$ and $W_{m,r}(n,k)$ were due to Cheon and Jung \cite{Cheon}, Merca \cite{Merca}, Corcino et al.\ \cite{Tina2}, Corcino et al.\ \cite{CorMonSus}, C. B. Corcino and R. B. Corcino \cite{Tina1}, and R. B. Corcino and C. B Corcino \cite{CorTin1,CorTin2}. Corcino and Hererra \cite{CorHer} introduced the \textit{limit of the differences of the generalized factorial} \begin{equation} F_{\alpha,\gamma}(n,k)=\lim_{\beta\rightarrow 0}\frac{\left[\Delta_t^k\left(\beta t+\gamma|\alpha\right)_n\right]_{t=0}}{k!\beta^k},\label{ldgf} \end{equation} where \begin{equation} \left(\beta t+\gamma|\alpha\right)_n=\prod_{j=0}^{n-1}\left(\beta t+\gamma-j\alpha\right),\ \left(\beta t+\gamma|\alpha\right)_0=1, \end{equation} which is a generalization of the Stirling numbers of the first kind. The numbers $F_{\alpha,-\gamma}(n,k)$ are actually the $r$-Whitney numbers of the first kind in \eqref{w1}. That is, \begin{equation*} F_{\alpha,-\gamma}(n,k)=w_{\alpha,\gamma}(n,k). \end{equation*} Similarly, Corcino \cite{CorRB} defined the \textit{$(r,\beta)$-Stirling numbers} $\rbstirling{n}{k}_{r,\beta}$ as coefficients in \begin{equation} t^n=\sum_{k=0}^n\binom{\frac{t-r}{\beta}}{k}\beta^{k}k!\rbstirling{n}{k}_{r,\beta}.\label{rbeta} \end{equation} The numbers $\rbstirling{n}{k}_{r,\beta}$ are found to be equivalent to the $r$-Whitney numbers of the second kind in \eqref{w2}. To be precise, \begin{equation*} \rbstirling{n}{k}_{r,\beta}=W_{\beta,r}(n,k). \end{equation*} Corcino et al.\ \cite{CorTinAl}, and Corcino and Aldema \cite{CorAl} further studied the numbers $\rbstirling{n}{k}_{r,\beta}$. Recall that the classical \textit{Boson operators} $a$ and $a^\dagger$ satisfy the commutation relation \begin{equation} [a,a^\dagger]\equiv aa^\dagger-a^\dagger a=1.\label{bos} \end{equation} If we define the Fock space by the basis $\{\ket{s} \; ; \; s=0,1,2,\ldots\}$, to be referred to as Fock states, the relations $a\ket{s}=\sqrt{s}\ket{s-1}$ and $a^{\dagger}\ket{s}=\sqrt{s+1}\ket{s+1}$ form a representation that satisfies the commutation relation \eqref{bos}. The operator $\hat{n}\equiv a^\dagger a$, when acting on $\ket{s}$, yields \begin{equation*} a^\dagger a\ket{s}=s\ket{s}, \end{equation*} and the operator $(a^\dagger)^k a^k$, when acting on the same state, yields \begin{equation*} (a^\dagger)^k a^k\ket{s}=(s)_k\ket{s}. \end{equation*} Let $\{\bra{s}\equiv (\ket{s})^{\dagger} \; ; \; s=0,1,2,\ldots\}$ denote the Fock basis of the dual space. Requiring the normalization of the scalar product $\braket{0}{0}=1$ we note that \begin{equation*} \braket{s+1}{s+1}=\frac{1}{s+1}\bra{s}aa^{\dagger}\ket{s}=\frac{1}{s+1}\Big(\bra{s}a^{\dagger}a\ket{s}+\braket{s}{s}\Big)=\braket{s}{s}. \end{equation*} Hence, from the normalization of $|0>$ it follows that all the Fock states are normalized. Moreover, since $\bra{s+1}a^{\dagger}\ket{s}=\sqrt{s+1}\braket{s+1}{s+1}$ and $(a\ket{s+1})^{\dagger}\ket{s}=\sqrt{s+1}\braket{s}{s}$, it follows that $a^{\dagger}$ is the Hermitian conjugate of $a$. That is, $a^{\dagger}a$ is Hermitian. Orthogonality follows from the fact that the Fock states are eigenstates of $a^{\dagger}a$ with distinct eigenvalues. Hence, the \textit{horizontal generating functions} of the Stirling numbers $\fstirling{n}{k}$ and $\sstirling{n}{k}$, \begin{equation} \label{standard} (x)_n=\sum_{k=0}^n (-1)^{n-k}\fstirling{n}{k}x^k \end{equation} and \begin{equation*} x^n=\sum_{k=0}^n\sstirling{n}{k}(x)_k, \end{equation*} can be expressed as \begin{equation*} (a^\dagger)^n a^n=\sum_{k=0}^n (-1)^{n-k}\fstirling{n}{k}(a^\dagger a)^k \end{equation*} and \begin{equation*} (a^\dagger a)^n=\sum_{k=0}^n\sstirling{n}{k}(a^\dagger)^k a^k, \end{equation*} respectively \cite{Kat1}. Now, the defining relations for the $r$-Whitney numbers, \eqref{w1} and \eqref{w2}, can be expressed as \begin{equation} m^n(a^\dagger)^n a^n=\sum_{k=0}^{n}w_{m,r}(n,k)(ma^\dagger a+r)^k\label{w3} \end{equation} and \begin{equation} (ma^\dagger a+r)^n=\sum_{k=0}^{n}m^kW_{m,r}(n,k)(a^\dagger)^k a^k.\label{w4} \end{equation} Making use of the $q$-Boson operators \cite{Arik} that satisfy \begin{equation} [a,a^\dagger]_q\equiv aa^\dagger-qa^\dagger a=1,\label{qbos} \end{equation} we have \begin{equation*} a\ket{s}=\sqrt{[s]_q}\ket{s-1}, \ a^{\dagger}\ket{s}=\sqrt{[s+1]_q}\ket{s+1}, \end{equation*} hence, \begin{equation*} a^\dagger a\ket{s}=[s]_q\ket{s}, \end{equation*} and \begin{equation*} (a^\dagger)^k a^k\ket{s}=[s]_{q,k}\ket{s}. \end{equation*} \begin{remark} Although we use the same notation for the boson and for the $q$-boson operators, no confusion should arise because the meaning of these symbols should be clear from the context. \end{remark} In line with this, the defining relations for Carlitz's \cite{Car3} $q$-Stirling numbers of the first and second kind, $\fstirling{n}{k}_q$ and $\sstirling{n}{k}_q$, can be written in the form \cite{Kat1} \begin{equation} (a^\dagger)^n a^n=\sum_{k=1}^n(-1)^{n-k}\fstirling{n}{k}_q(a^\dagger a)^k\label{c1} \end{equation} and \begin{equation} (a^\dagger a)^n=\sum_{k=1}^n\sstirling{n}{k}_q(a^\dagger)^k a^k,\label{c2} \end{equation} respectively. We define $q$-analogues for the Whitney numbers $w_{m,r}(n,k)$ and $W_{m,r}(n,k)$ via the same pattern as in \eqref{c1} and \eqref{c2}. \section{$(q,r)$-Whitney numbers} \label{sec:Whitney Numbers} \begin{definition} For non-negative integers $n$ and $k$ and complex numbers $r$ and $m$, the $(q,r)$-Whitney numbers of the first and second kind, denoted by $w_{m,r,q}(n,k)$ and $W_{m,r,q}(n,k)$, respectively, are defined by \begin{equation} m^n(a^\dagger)^n a^n=\sum_{k=0}^{n}w_{m,r,q}(n,k)(ma^\dagger a+r)^k\label{qw1} \end{equation} and \begin{equation} (ma^\dagger a+r)^n=\sum_{k=0}^{n}m^kW_{m,r,q}(n,k)(a^\dagger)^k a^k\label{qw2} \end{equation} with initial conditions $w_{m,r,q}(0,0)=W_{m,r,q}(0,0)=1$ and $w_{m,r,q}(n,k)=W_{m,r,q}(n,k)=0$ for $k>n$ and for $k<0$, where the operators $a^\dagger$ and $a$ satisfy the relation in \eqref{qbos}. \end{definition} Before proceeding we note that from \eqref{qw1} and \eqref{qw2}, \begin{eqnarray} w_{m,0,q}(n,k) &=& (-m)^{n-k}\fstirling{n}{k}_q,\label{w5} \\ W_{m,0,q}(n,k) &=& m^{n-k}\sstirling{n}{k}_q.\label{w6} \end{eqnarray} Similarly, the $r$-Stirling numbers $\widehat{\fstirling{n+r}{k+r}}_r$ and $\widehat{\sstirling{n+r}{k+r}}_r$ are specified by the horizontal generating functions \begin{equation*} (x-r)_n=\sum_{k=0}^n(-1)^{n-k}\widehat{\fstirling{n+r}{k+r}}_rx^k, \end{equation*} or, equivalently, \begin{equation*} (x)_n=\sum_{k=0}^n(-1)^{n-k}\widehat{\fstirling{n+r}{k+r}}_r (x+r)^k, \end{equation*} and \begin{equation*} (x+r)^n=\sum_{k=0}^n\widehat{\sstirling{n+r}{k+r}}_r(x)_r. \end{equation*} Hence, $\widehat{\fstirling{n+r}{k+r}}_{q,r}$ and $\widehat{\sstirling{n+r}{k+r}}_{q,r}$, the $q$-analogues of $\widehat{\fstirling{n+r}{k+r}}_r$ and $\widehat{\sstirling{n+r}{k+r}}_r$, respectively, are specified by the horizontal generating functions \begin{eqnarray} (a^{\dagger})^na^n &=& \sum_{k=0}^n(-1)^{n-k}\widehat{\fstirling{n+r}{k+r}}_{q,r}(a^{\dagger}a+r)^k,\label{qr1}\\ (a^{\dagger}a+r)^n &=& \sum_{k=0}^n\widehat{\sstirling{n+r}{k+r}}_{q,r}(a^{\dagger})^ka^k.\label{qr2} \end{eqnarray} It follows that \begin{eqnarray} w_{1,r,q}(n,k) &=& (-1)^{n-k}\widehat{\fstirling{n+r}{k+r}}_{q,r},\label{w7} \\ W_{1,r,q}(n,k) &=& \widehat{\sstirling{n+r}{k+r}}_{q,r}.\label{w8} \end{eqnarray} We will refer to the $q$-analogues in \eqref{qr1} and \eqref{qr2} as the $(q,r)$-Stirling numbers of the first and second kind, respectively. \begin{theorem} The $(q,r)$-Whitney numbers $w_{m,r,q}(n,k)$ and $W_{m,r,q}(n,k)$ satisfy the following identities: \begin{equation} w_{m,r,q}(n,k)=(-1)^{n-k}\sum_{i=k}^n \binom{i}{k}r^{i-k}m^{n-i}\fstirling{n}{i}_q, \label{identity1} \end{equation} \begin{equation} W_{m,r,q}(n,k)=\sum_{i=k}^n\binom{n}{i}r^{n-i}m^{i-k}\sstirling{i}{k}_q.\label{identity2} \end{equation} \end{theorem} \begin{proof} From Eq.~\eqref{c1}, we get \begin{eqnarray*} m^n(a^\dagger)^n a^n&=&m^n\sum_{i=0}^n(-1)^{n-i}\fstirling{n}{i}_q(a^\dagger a)^i\\ &=&m^n\sum_{i=0}^n(-1)^{n-i}\fstirling{n}{i}_q\left(\frac{\hat{z}-r}{m }\right)^i\\ &=&m^n\sum_{i=0}^n(-1)^{n-i}\fstirling{n}{i}_q\frac{1}{m^i}\sum_{k=0}^i\binom{i}{k}\hat{z}^k(-r)^{i-k}\\ &=&\sum_{k=0}^n (-1)^{n-k}\left\{\sum_{i=k}^n m^{n-i}\fstirling{n}{i}_q\binom{i}{k}r^{i-k}\right\}\hat{z}^k, \end{eqnarray*} where $\hat{z}=ma^\dagger a+r$. Furthermore, comparing the coefficient of $\hat{z}^k$ with that in equation \eqref{qw1} yields equation \eqref{identity1}. To prove equation \eqref{identity2}, we write \begin{eqnarray*} (ma^\dagger a+r)^n&=&\sum_{i=0}^n\binom{n}{i}r^{n-i}m^i(a^\dagger a)^i\\ &=&\sum_{i=0}^n\binom{n}{i}r^{n-i}m^i\sum_{k=0}^i\sstirling{i}{k}_q(a^\dagger)^k a^k\\ &=&\sum_{k=0}^n\left\{\sum_{i=k}^nr^{n-i}m^i\sstirling{i}{k}_q\binom{n}{i}\right\}(a^\dagger)^k a^k. \end{eqnarray*} Comparing the coefficient of $(a^\dagger)^k a^k$ with that in equation \eqref{qw2} gives us \eqref{identity2}. \end{proof} \begin{remark} \noindent (a) As $q\rightarrow 1$, we have \begin{equation*} w_{m,r}(n,k)=\sum_{i=k}^n(-1)^{n-k}\binom{i}{k}r^{i-k}m^{n-i}\fstirling{n}{i}; \end{equation*} \begin{equation*} W_{m,r}(n,k)=\sum_{i=k}^n\binom{n}{i}r^{n-i}m^{i-k}\sstirling{i}{k}. \end{equation*} \noindent (b) Note that of all the factors in equations \eqref{identity1} and \eqref{identity2} only the Stirling numbers are $q$-deformed. \end{remark} The following corollary is a direct consequence of the previous theorem. \begin{corollary}The $(q,r)$-Stirling numbers are given by \begin{equation} \widehat{\fstirling{n+r}{k+r}}_{q,r}=\sum_{i=k}^n\binom{i}{k}r^{i-k}\fstirling{n}{i}_q; \end{equation} \begin{equation} \widehat{\sstirling{n+r}{k+r}}_{q,r}=\sum_{i=k}^n\binom{n}{i}r^{n-i}\sstirling{i}{k}_q. \end{equation} \end{corollary} \section{Some recurrence relations} \label{sec:Recurrence} In this section, we present some recurrence relations involving the $(q,r)$-Whitney numbers. We recall the $q$-boson identities \begin{equation*} [a,(a^{\dagger})^n]_{q^n}=[n]_q(a^{\dagger})^{n-1} \end{equation*} and \begin{equation*} [a^n,a^{\dagger}]_{q^n}=[n]_qa^{n-1}, \end{equation*} that can be easily established by induction. The latter can also be written in the form \begin{equation*} a^{\dagger}a^n=q^{-n}(a^na^{\dagger}-[n]_qa^{n-1}). \end{equation*} \begin{theorem} The $(q,r)$-Whitney numbers $w_{m,r,q}(n,k)$ and $W_{m,r,q}(n,k)$ satisfy the following triangular recurrence relations: \begin{equation} w_{m,r,q}(n+1,k)=q^{-n}\Big(w_{m,r,q}(n,k-1)-(m[n]_q+r)w_{m,r,q}(n,k)\Big), \label{identity4} \end{equation} \begin{equation} W_{m,r,q}(n+1,k)=q^{k-1}W_{m,r,q}(n,k-1)+(m[k]_q+r)W_{m,r,q}(n,k).\label{identity5} \end{equation} \end{theorem} \begin{proof} From equation \eqref{qw1}, $\displaystyle\sum_{k=0}^{n+1}w_{m,r,q}(n+1,k)(ma^{\dagger}a+r)^k=m^{n+1}(a^{\dagger})^n(a^{\dagger}a^n)a$ \begin{eqnarray*} &=&m^{n+1}(a^{\dagger})^nq^{-n}(a^na^{\dagger}-[n]_qa^{n-1})a\\ &=&m^{n+1}q^{-n}\left((a^{\dagger})^na^n\right)(a^{\dagger}a)-m^{n+1}q^{-n}[n]_q(a^{\dagger})^na^n\\ &=&q^{-n}\sum_{k=0}^nw_{m,r,q}(n,k)(ma^{\dagger}a+r)^k(ma^{\dagger}a+r-r)-mq^{-n}[n]_q\sum_{k=0}^nw_{m,r,q}(n,k)(ma^{\dagger}a+r)^k\\ &=&q^{-n}\sum_{k=1}^{n+1}w_{m,r,q}(n,k-1)(ma^{\dagger}a+r)^k-q^{-n}(m[n]_q+r)\sum_{k=0}^nw_{m,r,q}(n,k)(ma^{\dagger}a+r)^k\\ &=&q^{-n}\sum_{k=0}^{n+1} \left\{w_{m,r,q}(n,k-1)-(m[n]_q+r)w_{m,r,q}(n,k)\right\}(ma^{\dagger}a+r)^k. \end{eqnarray*} Equating coefficients of $(ma^{\dagger}a+r)^k$ gives us \eqref{identity4} and a similar derivation yields equation \eqref{identity5}. \end{proof} Equations \eqref{identity4} and \eqref{identity5} are useful in computing the first few values of $w_{m,r,q}(n,k)$ and $W_{m,r,q}(n,k)$, using the initial values specified above. \begin{remark} \noindent (a) From \eqref{identity4} we obtain the explicit expression $$w_{m,r,q}(n,0)=(-1)^n q^{-\frac{n(n-1)}{2}} \prod_{i=0}^{n-1}(m[i]_q+r).$$ On the other hand, the relation \eqref{identity1} yields $$w_{m,r,q}(n,0)=(-1)^n\sum_{i=0}^n r^im^{n-i}\fstirling{n}{i}_q.$$ Equating these expressions and substituting $x=\frac{r}{m}$ we obtain $$\sum_{i=0}^n\fstirling{n}{i}_q x^i = q^{-\frac{n(n-1)}{2}}\prod_{i=0}^{n-1}([i]_q+x).$$ This is a horizontal generating function for the $q$-Stirling numbers of the first kind in terms of a $q$-analogue of the rising factorial. Indeed, replacing $x$ by $-[s]_q$, and noting that $$[s]_q-[i]_q=q^{-i}[s-i]_q$$ and $$\prod_{i=0}^{n-1}q^i = q^{\binom{n}{2}},$$ we obtain $$\sum_{i=0}^n \fstirling{n}{i}_q(-1)^i[s]_q^i =(-1)^n\prod_{i=0}^{n-1}[s-i]_q.$$ \noindent (b) From \eqref{identity5} $W_{m,r,q}(n+1,0)=rW_{m,r,q}(n,0)$, hence $W_{m,r,q}(n,0)=r^n$. The same result is obtained from \eqref{identity2}. That is, $$W_{m,r,q}(n,0)=\sum_{i=0}^n\binom{n}{i}r^{n-i}m^i\delta_{i,0}=r^n.$$ \noindent (c) As $q\rightarrow 1$, we have \begin{equation*} w_{m,r}(n+1,k)=w_{m,r}(n,k-1)-(mn+r)w_{m,r}(n,k); \end{equation*} \begin{equation*} W_{m,r}(n+1,k)=W_{m,r}(n,k-1)+(mk+r)W_{m,r}(n,k). \end{equation*} This confirms that $w_{m,r,q}(n,k)$ and $W_{m,r,q}(n,k)$ are proper $q$-analogues of $w_{m,r}(n,k)$ and $W_{m,r}(n,k)$, respectively. \end{remark} As a consequence of the previous theorem, when $m=1$ we have \begin{corollary} The $(q,r)$-Stirling numbers satisfy the following triangular recurrence relations: \begin{eqnarray*} \widehat{\fstirling{n+1+r}{k+r}}_{q,r}&=&q^{-n}\widehat{\fstirling{n+r}{k-1+r}}_{q,r}+([n]_q+r)q^{-n}\widehat{\fstirling{n+r}{k+r}}_{q,r},\\ \widehat{\sstirling{n+1+r}{k+r}}_{q,r}&=&q^{k-1}\widehat{\sstirling{n+r}{k-1+r}}_{q,r}+([k]_q+r)\widehat{\sstirling{n+r}{k+r}}_{q,r}. \end{eqnarray*} \end{corollary} We can use these recurrence relations to compute the first few values of the $(q,r)$-Stirling numbers of the first and second kind, respectively. \begin{theorem} The $(q,r)$-Whitney numbers satisfy the following recurrence relations \begin{equation} \label{T1} w_{m,r+1,q}(n,\ell) = \sum_{k=\ell}^n\binom{k}{\ell}(-1)^{k-\ell}w_{m,r,q}(n,k), \end{equation} \begin{equation} \label{T2} W_{m,r+1,q}(n,k) = \sum_{\ell=k}^n\binom{n}{\ell}W_{m,r,q}(\ell,k). \end{equation} \end{theorem} \begin{proof} From equation \eqref{qw1}, we have \begin{eqnarray*} m^n(a^{\dagger})^na^n &=& \sum_{k=0}^n w_{m,r,q}(n,k)(ma^{\dagger}a+r)^k \\ &=& \sum_{k=0}^n w_{m,r,q}(n,k)\Big( (ma^{\dagger}a+r+1)-1\Big)^k \\ &=& \sum_{k=0}^n w_{m,r,q}(n,k)\sum_{\ell=0}^k\binom{k}{\ell}(-1)^{k-\ell}(ma^{\dagger}a+r+1)^{\ell} \\ &=& \sum_{\ell=0}^n(ma^{\dagger}a+r+1)^{\ell}\sum_{k=\ell}^n\binom{k}{\ell}(-1)^{k-\ell}w_{m,r,q}(n,k). \end{eqnarray*} On the other hand, $$m^n(a^{\dagger})^na^n = \sum_{\ell=0}^n w_{m,r+1,q}(n,\ell)(ma^{\dagger}a+r+1)^{\ell}.$$ Hence, by comparing the coefficients of $(ma^{\dagger}a+r+1)^{\ell}$ we obtain equation \eqref{T1}. Similarly, from equation \eqref{qw2} \begin{equation*} (ma^{\dagger}a+r+1)^n=\sum_{k=0}^n m^k W_{m,r+1,q}(n,k)(a^{\dagger})^ka^k, \end{equation*} and since \begin{eqnarray*} (ma^{\dagger}a+r+1)^n &=& \sum_{\ell=0}^n\binom{n}{\ell} (ma^{\dagger}a+r)^{\ell}\\ &=& \sum_{\ell=0}^n\binom{n}{\ell}\sum_{k=0}^{\ell}m^kW_{m,r,q}(\ell,k)(a^{\dagger})^ka^k \\ &=& \sum_{k=0}^n m^k(a^{\dagger})^ka^k \sum_{\ell=k}^n\binom{n}{\ell}W_{m,r,q}(\ell,k), \end{eqnarray*} we obtain equation \eqref{T2}. \end{proof} When $m=1$, the theorem reduces to the recursion formulas for $(q,r)$-Stirling numbers. That is, \begin{corollary} \begin{eqnarray*} \widehat{\fstirling{n+r+1}{l+r+1}}_{q,r+1}&=&\sum_{k=l}^n(-1)^{l-k}\binom{k}{l}\widehat{\fstirling{n+r}{k+r}}_{q,r},\\ \widehat{\sstirling{n+r+1}{k+r+1}}_{q,r+1}&=&\sum_{l=k}^n\binom{n}{l}\widehat{\sstirling{l+r}{k+r}}_{q,r}. \end{eqnarray*} \end{corollary} \section{Orthogonality and inverse relations} \begin{theorem} The $(q,r)$-Whitney numbers $w_{m,r,q}(n,k)$ and $W_{m,r,q}(k,j)$ satisfy the following orthogonality relations: \begin{equation} \label{O1} \sum_{k=j}^{n}W_{m,r,q}(n,k)w_{m,r,q}(k,j)=\delta_{jn}, \end{equation} and \begin{equation} \label{O2} \sum_{k=j}^{n}w_{m,r,q}(n,k)W_{m,r,q}(k,j)=\delta_{jn}, \end{equation} where $\delta_{jn}$ is the Kronecker delta. \end{theorem} \begin{proof} Using equation \eqref{qw1} we substitute $m^k(a^\dagger)^k a^k$ in \eqref{qw2}, obtaining \begin{eqnarray*} (ma^\dagger a+r)^n&=&\sum_{k=0}^{n}W_{m,r,q}(n,k)\sum_{j=0}^{k}w_{m,r,q}(k,j)(ma^\dagger a+r)^j\\ &=&\sum_{j=0}^{n}\left\{\sum_{k=j}^{n}W_{m,r,q}(n,k)w_{m,r,q}(k,j)\right\}(ma^\dagger a+r)^j. \end{eqnarray*} Comparing the coefficients of $(ma^\dagger a+r)^j$ yields equation \eqref{O1}. Equation \eqref{O2} is obtained similarly. \end{proof} The classical \textit{binomial inversion formula} given by \begin{equation} f_k=\sum_{j=0}^k\binom{k}{j}g_j\Leftrightarrow g_k=\sum_{j=0}^k(-1)^{k-j}\binom{k}{j}f_j\label{binominv} \end{equation} can be a useful tool in deriving the explicit formula of the classical Stirling numbers of the second kind. The $q$-analogue of \eqref{binominv} is given by \cite{Comt} \begin{equation} f_k=\sum_{j=0}^k\binom{k}{j}_qg_j\Leftrightarrow g_k=\sum_{j=0}^k(-1)^{k-j}q^{\binom{k-j}{2}}\binom{k}{j}_qf_j,\label{qbinominv} \end{equation} The next theorem presents an inverse relation for the $(q,r)$-Whitney numbers $w_{m,r,q}(n,k)$ and $W_{m,r,q}(k,j)$. \begin{theorem} The $(q,r)$-Whitney numbers $w_{m,r,q}(n,\ell)$ and $W_{m,r,q}(n,\ell)$ satisfy the following inverse relation: \begin{equation} f_n=\sum_{\ell=0}^nw_{m,r,q}(n,\ell)g_{\ell}\Leftrightarrow g_n=\sum_{\ell=0}^nW_{m,r,q}(n,\ell)f_{\ell}. \end{equation} \end{theorem} \begin{proof} By the hypothesis, \begin{eqnarray*} \sum_{\ell=0}^n W_{m,r,q}(n,\ell)f_{\ell}&=&\sum_{\ell=0}^n W_{m,r,q}(n,\ell)\sum_{k=0}^{\ell} w_{m,r,q}(\ell,k)g_k\\ &=&\sum_{k=0}^n\left\{\sum_{\ell=k}^n W_{m,r,q}(n,\ell)w_{m,r,q}(\ell,k)\right\}g_k\\ &=&\sum_{k=0}^n\left\{\delta_{kn}\right\}g_k\\ &=&g_n. \end{eqnarray*} The converse can be shown similarly. \end{proof} The next theorem can be deduced in a similar way, from the orthogonality relations \begin{theorem} The $(q,r)$-Whitney numbers $w_{m,r,q}(n,\ell)$ and $W_{m,r,q}(n,\ell)$ satisfy the following inverse relation: \begin{equation} f_{\ell}=\sum_{n=\ell}^{\infty}w_{m,r,q}(n,\ell)g_n\Leftrightarrow g_{\ell}=\sum_{n=\ell}^{\infty}W_{m,r,q}(n,\ell)f_n. \end{equation} \end{theorem} \section{$(q,r)$-Dowling polynomials and numbers} Cheon and Jung \cite{Cheon} defined the \textit{$r$-Dowling polynomials}, denoted by $D_{m,r}(n,x)$, in terms of sums of $r$-Whitney numbers of the second kind. That is, \begin{equation} D_{m,r}(n,x)=\sum_{k=0}^nW_{m,r}(n,k)x^k.\label{rdowling} \end{equation} When $x=1$, we obtain the \textit{$r$-Dowling numbers} $D_{m,r}(n)\equiv D_{m,r}(n,1)$. The polynomials \eqref{rdowling} are actually equivalent to the $(r,\beta)$-Bell polynomials $G_{n,\beta,r}(x)$ of R. B. Corcino and C. B. Corcino \cite{Cor}. That is, \begin{equation*} D_{\beta,r}(n,x)=G_{n,\beta,r}(x). \end{equation*} Moreover, \begin{itemize} \item when $m=1$ and $r=1$, we recover the classical \textit{Dowling polynomials} $D(n,x)\equiv D_{1,1}(n,x)$; \item when $m=1$ and $r=0$, we recover the classical \textit{Bell polynomials} $B_n(x)\equiv D_{1,0}(n,x)$; \item when $m=1$, we recover Mez\H{o}'s \cite{Mez} \textit{$r$-Bell polynomials} $B_{n,r}(x)$. That is, $D_{1,r}(n,x)=B_{n,r}(x)$; and \item when $m=\alpha$ and $r=0$, we recover the \textit{translated Dowling polynomials} $\widetilde{D}_{(\alpha)}(n;x)$ by Mangontarum et al.\ \cite{Mah1}. That is, $D_{\alpha,0}(n,x)=\widetilde{D}_{(\alpha)}(n;x)$. \end{itemize} \smallskip Taking these into consideration, the next definition seems to be natural. \begin{definition} For non-negative integers $n$ and $k$, and complex numbers $m$ and $r$, the $(q,r)$-Dowling polynomials, denoted by $D_{m,r,q}(n,x)$, are defined by \begin{equation} D_{m,r,q}(n,x)=\sum_{k=0}^{n}W_{m,r,q}(n,k)x^k\label{qDP} \end{equation} and the $(q,r)$-Dowling numbers, denoted by $D_{m,r,q}(n)$, are defined by \begin{equation} D_{m,r,q}(n)=D_{m,r,q}(n,1).\label{qDN} \end{equation} \end{definition} \smallskip The \textit{coherent states} \begin{equation} \ket{\gamma}=\hbox{exp}\left(-\frac{|\gamma|^2}{2}\right)\sum_{k\geq 0}\frac{\gamma^k}{\sqrt{k!}}\ket{k}, \label{coherent} \end{equation} where $\gamma$ is an arbitrary (complex-valued) constant, satisfy $a\ket{\gamma}=\gamma\ket{\gamma}$ and $\braket{\gamma}{\gamma}=1$. Katriel \cite{Kat2} gave an illustration on how \eqref{coherent} can be a very useful tool in the derivation of certain Dobinski-type formulas. The \textit{q-coherent states} corresponding to the $q$-Boson operators were defined as \begin{equation} \ket{\gamma}_q=\left(\widehat{e}_q(-|\gamma|^2)\right)^{\frac{1}{2}}\sum_{k\geq 0}\frac{\gamma^k}{\sqrt{[k]_q!}}\ket{k}\label{qCS} \end{equation} which satisfy $a\ket{\gamma}=\gamma\ket{\gamma}$. Here, $\widehat{e}_q(x)$ is the type 2 $q$-exponential function given by \begin{equation} \widehat{e}_q(x)=\prod_{i=1}^{\infty}(1+(1-q)q^{i-1}x)=\sum_{i\geq 0}q^{\binom{i}{2}}\frac{x^i}{[i]_q!},\label{type2qexp} \end{equation} which is the inverse of the type 1 $q$-exponential function \begin{equation} e_q(x)=\prod_{i=1}^{\infty}(1-(1-q)q^{i-1}x)^{-1}=\sum_{i\geq 0} \frac{x^i}{[i]_q!}.\label{type1qexp} \end{equation} That is, $e_q(x)\widehat{e}_q(-x)=1$. Taking the expectation value of both sides of \eqref{qw2} with respect to $\ket{\gamma}$ yields \begin{equation} \bra{\gamma}(ma^\dagger a+r)^n\ket{\gamma}=\sum_{k=0}^{n}m^kW_{m,r,q}(n,k)|\gamma|^{2k}. \end{equation} The left-hand-side can be evaluated using the $q$-coherent states in \eqref{qCS}, yielding \begin{equation} \bra{\gamma}(ma^\dagger a+r)^n\ket{\gamma}=\widehat{e}_q\left(-|\gamma|^2\right)\sum_{k\geq 0}\frac{|\gamma|^{2k}}{[k]_q!}(m[k]_q+r)^n. \end{equation} Defining $x=m|\gamma|^2$ we obtain \begin{equation} \sum_{k=0}^{n}W_{m,r,q}(n,k)x^k=\widehat{e}_q\left(-\frac{x}{m}\right)\sum_{k\geq 0}\left(\frac{x}{m}\right)^k\frac{(m[k]_q+r)^n}{[k]_q!}. \end{equation} Using \eqref{qDP}, the following theorem is easily observed. \begin{theorem} The $(q,r)$-Dowling polynomials $D_{m,r,q}(n,x)$ and the $(q,r)$-Dowling numbers $D_{m,r,q}(n)$ have the following explicit formulas: \begin{equation} D_{m,r,q}(n,x)=\widehat{e}_q\left(-\frac{x}{m}\right)\sum_{k\geq 0}\left(\frac{x}{m}\right)^k\frac{(m[k]_q+r)^n}{[k]_q!},\label{qDPe} \end{equation} and \begin{equation} D_{m,r,q}(n)=\widehat{e}_q\left(-m^{-1}\right)\sum_{k\geq 0}\frac{(m[k]_q+r)^n}{m^k[k]_q!}.\label{qDNe} \end{equation} \end{theorem} \begin{proof} \eqref{qDNe} can be obtained by letting $x=1$ in \eqref{qDPe}. \end{proof} Katriel \cite{Kat2} defined the $q$-Bell polynomial as \begin{equation} \sum_{\ell=0}^k\sstirling{k}{\ell}_qx^{\ell}=\widehat{e}_q(x)\sum_{m=1}^{\infty}x^m\frac{[m]_q^k}{[m]_q!}. \end{equation} Expanding the right-hand side using \eqref{type2qexp} yields \begin{equation} \sum_{\ell=0}^k\sstirling{k}{\ell}_qx^{\ell} =\sum_{\ell=0}^{\infty}\frac{x^{\ell}}{[\ell]_q!}\sum_{j=0}^{\ell}(-1)^{\ell-j}q^{\binom{\ell-j}{2}}\binom{\ell}{j}_q[j]_q^k. \end{equation} Equating coefficients of equal powers of $x$ gives us \begin{equation} \sstirling{k}{\ell}_q=\frac{1}{[\ell]_q!}\sum_{j=0}^{\ell}(-1)^{\ell-j}q^{\binom{\ell-j}{2}}\binom{\ell}{j}_q[j]^k_q.\label{c3} \end{equation} Notice that as $q\rightarrow 1$, \eqref{c3} reduces to the well-known explicit formula of $\sstirling{k}{j}$. That is \begin{equation} \lim_{q\rightarrow 1}\sstirling{k}{\ell}_q=\frac{1}{{\ell}!}\sum_{j=0}^{\ell}(-1)^{\ell-j}\binom{\ell}{j}j^k. \end{equation} In the following theorem, we will present an expression analogous to \eqref{c3} for the $q$-analogue $W_{m,r,q}(n,k)$. \begin{theorem} The $(q,r)$-Whitney numbers of the second kind, $W_{m,r,q}(n,k)$, have the following explicit formula: \begin{equation} W_{m,r,q}(n,\ell)=\frac{1}{m^{\ell}[\ell]_q!}\sum_{k=0}^{\ell}(-1)^{\ell-k}q^{\binom{\ell-k}{2}}\binom{\ell}{k}_q(m[k]_q+r)^n.\label{identity3} \end{equation} \end{theorem} \begin{proof} Substituting $y=\frac{x}{m}$ in \eqref{qDPe} gives us \begin{eqnarray*} \sum_{k=0}^nm^kW_{m,r,q}(n,k)y^k&=&\sum_{i\geq 0}q^{\binom{i}{2}}\frac{(-y)^i}{[i]_q!}\sum_{k\geq 0}y^k\frac{(m[k]_q+r)^n}{[k]_q!}\\ &=&\sum_{\ell\geq 0}\frac{y^{\ell}}{[\ell]_q!}\sum_{k=0}^{\ell}(-1)^{\ell-k}q^{\binom{\ell-k}{2}}\binom{\ell}{k}_q(m[k]_q+r)^n. \end{eqnarray*} Equating the coefficients of equal powers of $y$ on both sides of this equation we obtain equation \eqref{identity3}. \end{proof} Note that as $q\rightarrow 1$, we have \begin{eqnarray*} \lim_{q\rightarrow 1}W_{m,r,q}(n,\ell)&=&\frac{1}{m^{\ell}{\ell}!}\sum_{k=0}^{\ell}(-1)^{\ell-k}\binom{\ell}{k}(mk+r)^n\\ &=&W_{m,r}(n,\ell). \end{eqnarray*} Furthermore, \begin{equation*} \lim_{q\rightarrow 1}W_{m,1,q}(n,l)=W_m(n,l). \end{equation*} \begin{remark} We can also prove \eqref{identity3} in the following manner: First, we write \eqref{qw2} as \begin{eqnarray*} (m[\ell]_q+r)^n&=&\sum_{k=0}^{n}m^kW_{m,r,q}(n,k)[\ell]_{q,k}\\ &=&\sum_{k=0}^{\ell}\binom{\ell}{k}_q\left\{\frac{m^kW_{m,r,q}(n,k)[\ell]_{q,k}}{\binom{\ell}{k}_q}\right\}. \end{eqnarray*} Next, we apply the $q$-binomial inversion formula in \eqref{qbinominv} which gives us \begin{equation*} \frac{m^{\ell}W_{m,r,q}(n,\ell)[\ell]_{q,\ell}}{\binom{k}{k}_q} =\sum_{k=0}^{\ell}(-1)^{\ell-k}q^{\binom{\ell-k}{2}}\binom{l}{k}_q(m[k]_q+r)^n. \end{equation*} This is precisely the explicit formula obtained in the previous theorem. \end{remark} Now, using \eqref{identity3}, \begin{eqnarray*} \sum_{n\geq 0}W_{m,r,q}(n,k)\frac{t^n}{[n]_q!}&=&\sum_{n\geq 0}\sum_{j=0}^k\frac{(-1)^{k-j}}{m^k[k]_q!}q^{\binom{k-j}{2}}\binom{k}{j}_q(m[j]_q+r)^n\frac{t^n}{[n]_q!}\\ &=&\frac{1}{m^k[k]_q!}\sum_{j=0}^k(-1)^{k-j}q^{\binom{k-j}{2}}\binom{k}{j}_qe_q\left[(m[j]_q+r)t\right], \end{eqnarray*} where $e_q(x)$ is the type 1 $q$-exponential function in \eqref{type1qexp}. Thus, we have the following theorem. \begin{theorem} The $(q,r)$-Whitney numbers of the second kind satisfy the following exponential generating function: \begin{equation} \sum_{n\geq 0}W_{m,r,q}(n,k)\frac{t^n}{[n]_q!}=\frac{1}{m^k[k]_q!}\sum_{j=0}^n(-1)^{k-j}q^{\binom{k-j}{2}}\binom{k}{j}_{\! q} e_q\left[(m[j]_q+r)t\right].\label{identity6} \end{equation} \end{theorem} \begin{remark} As $q\rightarrow 1$, we have \begin{equation*} \lim_{q\rightarrow 1}\sum_{n\geq 0}W_{m,r,q}(n,k)\frac{t^n}{[n]_q!}=\frac{e^{rt}}{k!}\left(\frac{e^{mt}-1}{m}\right)^k, \end{equation*} which is the exponential generating function of the $r$-Whitney numbers of the second kind. \end{remark} The $q$-difference operator \cite{Kim} can be written in the form \begin{equation} \Delta^k_{q}f(x)=\sum_{j=0}^k(-1)^{k-j}q^{\binom{k-j}{2}}\binom{k}{j}_{\! q}f(x+j).\label{qDiffOp} \end{equation} We are now ready to state the next theorem. \begin{theorem} The $(q,r)$-Whitney numbers of the second kind satisfy the following identity: \begin{equation} \sum_{n\geq 0}W_{m,r,q}(n,k)\frac{t^n}{[n]_q!}=\left\{\Delta^k_q\left(\frac{e_q[(m[x]_q+r)t]}{m^k[k]_q!}\right)\right\}_{x=0}.\label{identity7} \end{equation} \end{theorem} \begin{proof} \eqref{identity7} follows directly from \eqref{identity6} and \eqref{qDiffOp}. \end{proof} The next corollary is easily verified. \begin{corollary} The $(q,r)$-Whitney numbers of the second kind can be expressed explicitly as \begin{equation} W_{m,r,q}(n,k)=\left\{\Delta^k_q\left(\frac{(m[x]_q+r)^n}{m^k[k]_q!}\right)\right\}_{x=0}. \end{equation} \end{corollary} \section{Further identities for the $(q,r)$-Whitney numbers} Graham et al.\ \cite{Graham} presented a useful set of Stirling number identities while Katriel \cite{Kat1} presented the $q$-analogues of all but two of them. Three of these identities are generalized to the $(q,r)$-Whitney numbers using appropriate modifications of the procedures presented by Katriel \cite{Kat1}. Their derivation requires the following. \begin{lemma} For $f(x)$ a polynomial, the operator identity \begin{equation}\label{L1} a^{\dagger}f(1+qa^{\dagger}a)a=a^{\dagger}af(a^{\dagger}a), \end{equation} holds. \end{lemma} \begin{proof} We write the $q$-commutation relation, equation \eqref{qbos}, in the form $aa^{\dagger}=1+qa^{\dagger}a$. It follows that $$(a^{\dagger}a)(a^{\dagger}a)^k=a^{\dagger}(aa^{\dagger})^ka=a^{\dagger}(1+qa^{\dagger}a)^ka.$$ For $f(x)=\sum_{k}c_kx^k$ we obtain \begin{eqnarray*} a^{\dagger}af(a^{\dagger}a) &=& \sum_{k}c_k(a^{\dagger}a)(a^{\dagger}a)^k \\ &=& \sum_k c_ka^{\dagger}(1+qa^{\dagger}a)^ka = a^{\dagger}\left(\sum_k c_k(1+qa^{\dagger}a)^k\right)a \\ &=& a^{\dagger}f(1+qa^{\dagger}a)a. \end{eqnarray*} \end{proof} \begin{remark} The lemma can also be written in the form \begin{equation} \label{L2} a^{\dagger}g(a^{\dagger}a)a=a^{\dagger}a g\left(\frac{1}{q}(a^{\dagger}a-1)\right), \end{equation} where $g(x)$ is a polynomial. \end{remark} \begin{theorem}[Identity 1] The $(q,r)$-Whitney numbers of the second kind satisfy \begin{equation*} W_{m,r,q}(n+1,k)-rW_{m,r,q}(n,k)=\sum_{\ell=k-1}^n\binom{n}{\ell}q^{\ell}(m+r(1-q))^{n-\ell}W_{m,r,q}(\ell,k-1). \end{equation*} \end{theorem} \begin{proof} In terms of the identity \eqref{L1} and with the aid of \eqref{qw2} \begin{eqnarray*} a^{\dagger}\Big(m(1+qa^{\dagger}a)+r\Big)^na &=& a^{\dagger}a(ma^{\dagger}a+r)^n \\ &=& \frac{1}{m}(ma^{\dagger}a+r-r)(ma^{\dagger}a+r)^n\\ &=& \frac{1}{m}(ma^{\dagger}a+r)^{n+1}-\frac{r}{m}(ma^{\dagger}a+r)^n \\ &=& \sum_{k=0}^{n+1}m^{k-1}(a^{\dagger})^ka^k\Big(W_{m,r,q}(n+1,k)-rW_{m,r,q}(n,k)\Big). \end{eqnarray*} On the other hand, defining $\alpha=m+r(1-q)$ (which will hold throught the present section), \begin{eqnarray*} a^{\dagger}\Big(m(1+qa^{\dagger}a)+r\Big)^na &=& a^{\dagger}\Big(q(ma^{\dagger}a+r)+\alpha\Big)^na \\ &=& a^{\dagger}\left(\sum_{\ell=0}^n\binom{n}{\ell}q^{\ell}\alpha^{n-\ell}(ma^{\dagger}a+r)^{\ell}\right)a \\ &=& \sum_{\ell=0}^n\binom{n}{\ell}q^{\ell}\alpha^{n-{\ell}}\sum_{k=0}^{\ell}m^kW_{m,r,q}(\ell,k)(a^{\dagger})^{k+1}a^{k+1} \\ &=& \sum_{k=1}^{n+1}m^{k-1}(a^{\dagger})^ka^k\sum_{\ell=k-1}^n\binom{n}{\ell}q^{\ell}\alpha^{n-\ell}W_{m,r,q}(\ell,k-1). \end{eqnarray*} Equating coefficients of $m^{k-1}(a^{\dagger})^ka^k$ the theorem follows. \end{proof} For $r=0$ this identity reduces to the $q$-Stirling numbers identity \cite[identity\ 1]{Kat1} \begin{equation*} W_{m,0,q}(n+1,k)=\sum_{\ell=k-1}^n\binom{n}{\ell}q^{\ell}m^{n-\ell}W_{m,0,q}(\ell,k-1). \end{equation*} The following corollary is an immediate consequence of the previous theorem. \begin{corollary} As $q\rightarrow 1$, \begin{equation*} W_{m,r}(n+1,k)-rW_{m,r}(n,k)=\sum_{\ell=k-1}^n\binom{n}{\ell}m^{n-\ell}W_{m,r}(\ell,k-1). \end{equation*} \end{corollary} \begin{theorem}[Identity 2] The $(q,r)$-Whitney numbers of the first kind satisfy \begin{eqnarray*} w_{m,r,q}(n+1,\ell) &=& \sum_{k=\ell-1}^n\frac{1}{q^k}w_{m,r,q}(n,k) \Big(-(m+r(1-q)\Big)^{k-\ell}\\ & &\ \cdot\left(\binom{k}{\ell-1}(-(m+r(1-q)))-r\binom{k}{\ell}\right). \end{eqnarray*} \end{theorem} \begin{proof} We note that from \eqref{qw1}, \begin{equation*} m^{n+1}(a^{\dagger})^{n+1}a^{n+1} = \sum_{\ell=0}^{n+1}w_{m,r,q}(n+1,\ell)(ma^{\dagger}a+r)^{\ell}. \end{equation*} On the other hand, using \eqref{L2}, \begin{eqnarray*} m^{n+1}(a^{\dagger})^{n+1}a^{n+1} &=& ma^{\dagger}\Big(m^n(a^{\dagger})^na^n\Big)a\\ &=& ma^{\dagger}\Big(\sum_{k=0}^n w_{m,r,q}(n,k)(ma^{\dagger}a+r)^k\Big)a \\ &=& ma^{\dagger}a\sum_{k=0}^n w_{m,r,q}(n,k)\Big(\frac{m}{q}(a^{\dagger}a-1)+r \Big)^k \\ &=& ma^{\dagger}a\sum_{k=0}^n w_{m,r,q}(n,k)\frac{1}{q^k}\Big((ma^{\dagger}a+r)-\alpha \Big)^k\\ &=& ((ma^{\dagger}a+r)-r)\sum_{k=0}^n w_{m,r,q}(n,k)\frac{1}{q^k} \sum_{\ell=0}^k\binom{k}{\ell}(ma^{\dagger}a+r)^{\ell}(-\alpha)^{k-\ell} \\ &=& \sum_{k=0}^n w_{m,r,q}(n,k)\frac{1}{q^k}\sum_{\ell=0}^k\binom{k}{\ell}(ma^{\dagger}a+r)^{\ell+1}(-\alpha)^{k-\ell}\\ & & -r\sum_{k=0}^n w_{m,r,q}(n,k)\frac{1}{q^k}\sum_{\ell=0}^n\binom{k}{\ell}(ma^{\dagger}a+r)^{\ell}(-\alpha)^{k-\ell}\\ &=& \sum_{\ell=0}^{n+1}(ma^{\dagger}a+r)^{\ell}\sum_{k=\ell-1}^n\frac{1}{q^k}w_{m,r,q}(n,k)(-\alpha)^{k-\ell}\cdot \\ & & \;\;\;\;\; \cdot \left(\binom{k}{\ell-1}(-\alpha)-r\binom{k}{\ell}\right). \end{eqnarray*} Equating the coefficients of equal powers of $ma^{\dagger}a+r$ we obtain the theorem. \end{proof} For $r=0$, we recover the $q$-Stirling numbers identity \cite[identity\ 2]{Kat1} \begin{equation*} w_{m,0,q}(n+1,\ell)=\sum_{k=\ell-1}^n\frac{1}{q^k}w_{m,0,q}(n,k)(-m)^{k-\ell+1}\binom{k}{\ell-1}, \end{equation*} Moreover, we have the following corollary: \begin{corollary} As $q\rightarrow 1$, \begin{equation*} w_{m,r}(n+1,\ell)=-\sum_{k=\ell-1}^nw_{m,r}(n,k)(-m)^{k-\ell}\left(m\binom{k}{\ell-1}+r\binom{k}{\ell}\right). \end{equation*} \end{corollary} \begin{theorem}[Identity 3] The $(q,r)$-Whitney numbers of the second kind satisfy \begin{equation*} W_{m,r,q}(n,k-1) = \frac{1}{q^n}\sum_{\ell=k}^{n+1} (-m-r(1-q))^{n-\ell}\left(\binom{n}{\ell-1}(-m-r(1-q))-\binom{n}{\ell}r\right)W_{m,r,q}(\ell,k). \end{equation*} \end{theorem} \begin{proof} Note that \begin{eqnarray*} a^{\dagger}(ma^{\dagger}a+r)^na&=&\sum_{k=0}^n m^kW_{m,r,q}(n,k)(a^{\dagger})^{k+1}a^{k+1}\\ &=&\sum_{k=1}^{n+1} m^{k-1}W_{m,r,q}(n,k-1)(a^{\dagger})^ka^k, \end{eqnarray*} and on the other hand, using \eqref{L2}, \begin{eqnarray*} a^{\dagger}(ma^{\dagger}a+r)^na &=& a^{\dagger}a\Big(\frac{m}{q}(a^{\dagger}a-1)+r\Big)^n =a^{\dagger}a\frac{1}{q^n}(ma^{(\dagger}a+r)-\alpha)^n \\ &=& \frac{1}{m} ((ma^{\dagger}a+r)-r)\frac{1}{q^n}\sum_{\ell=0}^n\binom{n}{\ell}(ma^{\dagger}a+r)^{\ell}(-\alpha))^{n-\ell} \\ &=& \frac{1}{mq^n} \sum_{\ell=1}^{n+1} (ma^{\dagger}a+r)^{\ell}(-\alpha)^{n-\ell}\cdot\left( \binom{n}{\ell-1}(-\alpha) -r\binom{n}{\ell}\right) \\ &=& \frac{1}{mq^n}\sum_{k=0}^{n+1} m^k(a^{\dagger})^ka^k\sum_{\ell=k}^{n+1} (-\alpha)^{n-\ell}\left(\binom{n}{\ell-1}(-\alpha)-\binom{n}{\ell}r\right)W_{m,r,q}(\ell,k). \end{eqnarray*} Equating the coefficients of $(a^{\dagger})^ka^k$ we obtain the theorem. \end{proof} For $r=0$ this theorem reduces to \begin{equation*} W_{m,0,q}(n,k-1)=\frac{1}{q^n}\sum_{\ell=k}^{n+1}(-m)^{n+1-\ell}\binom{n}{\ell-1}W_{m,0,q}(\ell,k). \end{equation*} Using equation \eqref{w6}, we can verify that this is just the $q$-Stirling numbers identity \cite[identity\ 3]{Kat1}. The next corollary is easily verified. \begin{corollary} As $q\rightarrow 1$, \begin{equation*} W_{m,r}(n,k-1)=\sum_{\ell=k}^{n+1}(-m)^{n-\ell}\left[\binom{n}{\ell-1}(-m)-\binom{n}{\ell}r\right]W_{m,r}(\ell,k). \end{equation*} \end{corollary} Presently, much is yet to be learnt regarding the $(q,r)$-Whitney numbers. The classical $r$-Whitney and Stirling numbers are known to have various applications in different fields. It is tempting to explore applications for the $(q,r)$-Whitney numbers. To close this section, Corcino and Hererra \cite{CorHer} defined the $q$-analogue of the limit of the differences of the generalized factorial $F_{\alpha,\gamma}(n,k)$ in \eqref{ldgf}, denoted by $\phi_{\alpha,\gamma}[n,k]_q$. $\phi_{\alpha,\gamma}[n,k]_q$ can be defined in terms of the relation \begin{equation} \sum_{k=0}^n\phi_{\alpha,\gamma}[n,k]_qt^k=\left\langle t+[\gamma]_q|[\alpha]_q\right\rangle^q_n, \end{equation} where \begin{equation} \left\langle t+[\gamma]_q|[\alpha]_q\right\rangle^q_n=\prod_{j=0}^{n-1}\left(t+[\gamma]_q-[j\alpha]_q\right). \end{equation} The numbers $\phi_{\alpha,\gamma}[n,k]_q$ are actually $q$-analogues of the numbers $w_{m,r}(n,k)$. Similarly, Corcino and Montero \cite{CorMon} defined the $q$-analogue $\sigma[n,k]^{\beta,r}_q$ of the Rucinski-Voigt numbers in terms of the reccurence relation \begin{equation} \sigma[n,k]^{\beta,r}_q=\sigma[n-1,k-1]^{\beta,r}_q+\left([k\beta]_q+[r]_q\right)\sigma[n-1,k]^{\beta,r}_q. \end{equation} $\sigma[n,k]^{\beta,r}_q$ is also a $q$-analogue of the numbers $\rbstirling{n}{k}_{r,\beta}$ and $W_{m,r}(n,k)$. However, by comparing the defining relations for $\phi_{\alpha,\gamma}[n,k]_q$ and $\sigma[n,k]^{\beta,r}_q$ with those of the$(q,r)$-Whitney numbers $w_{m,r,q}(n,k)$ and $W_{m,r,q}(n,k)$, respectively, we note that they represent distinctly motivated $q$-analogues that cannot be simply related to one another. \section{Acknowledgments} The authors would like to thank the editor-in-chief for his helpful comments and suggestions, and the referee for carefully reading the manuscript and giving invaluable recommendations which helped improve this paper. \begin{thebibliography}{99} \bibitem{Arik} M. Arik and D. Coon, Hilbert spaces of analytic functions and generalized coherent states, \textit{J. Math. Phys.}, \textbf{17} (1976), 524--527. DOI: \href{http://dx.doi.org/10.1063/1.522937}{doi:10.1063/1.522937}. \bibitem{Bel} H. Belbachir and I. Bousbaa, Translated Whitney and $r$-Whitney numbers: a combinatorial approach, \textit{J. Integer Seq.}, \textbf{16} (2013), \href{https://cs.uwaterloo.ca/journals/JIS/VOL16/Belbachir/belbachir32.html}{Article 13.8.6}. \bibitem{Ben1} M. Benoumhani, On Whitney numbers of Dowling lattices, \textit{Discrete Math.}, \textbf{159} (1996), 13--33. \bibitem{Ben2} M. Benoumhani, On some numbers related to the Whitney numbers of Dowling lattices, \textit{Adv. Appl. Math.}, \textbf{19} (1997), 106--116. \bibitem{Bro} A. Broder, The $r$-Stirling numbers, \textit{Discrete Math.} \textbf{49} (1984), 241--259. \bibitem{Car3} L. Carlitz, $q$-Bernoulli numbers and polynomials, \textit{Duke Math. J.} \textbf{15} (1948), 987--1000. \bibitem{Cheon} G.-S. Cheon and J.-H. Jung, The $r$-Whitney numbers of Dowling lattices, \textit{Discrete Math.}, \textbf{15} (2012), 2337--2348. \bibitem{Comt} L. Comtet, \textit{Advanced Combinatorics}, D. Reidel Publishing Co., 1974. \bibitem{Tina1} C. B. Corcino and R. B. Corcino, Asymptotic estimates for second kind generalized Stirling numbers, \textit{J. Appl. Math.}, \textbf{2013}, Article ID 918513, 7 pages, (2013). DOI: \href{http://dx.doi.org/10.1155/2013/918513}{doi:10.1155/2013/918513}. \bibitem{Tina2} C. B. Corcino, R. B. Corcino, and N. Acala, Asymptotic estimates for $r$-Whitney numbers of the second kind, \textit{J. Appl. Math.}, \textbf{2014}, Article ID 354053, 7 pages, (2014). DOI: \href{http://dx.doi.org/10.1155/2014/354053}{doi:10.1155/2014/354053}. \bibitem{CorRB} R. B. Corcino, The $(r,\beta)$-Stirling numbers, \textit{The Mindanao Forum}, \textbf{14} (1999), 91--99. \bibitem{CorAl} R. B. Corcino and R. M. Aldema, Some combinatorial and statistical applications of $(r,\beta)$-Stirling numbers, \textit{Mati\-my\-\'as Mat.}, \textbf{25} (2002), 19--29. \bibitem{Cor} R. B. Corcino and C. B. Corcino, On generalized Bell polynomials, \textit{Discrete Dyn. Nat. Soc.}, \textbf{2011}, Article ID 623456, 21 pages, (2011). DOI: \href{http://dx.doi.org/10.1155/2011/623456}{doi:10.1155/2011/623456}. \bibitem{CorTin1} R. B. Corcino and C. B Corcino, On the maximum of generalized Stirling numbers, \textit{Util. Math.}, \textbf{86} (2011), 241--256. \bibitem{CorTin2} R. B. Corcino and C. B Corcino, The Hankel transform of the generalized Bell numbers and its $q$-analogue, \textit{Util. Math.}, \textbf{89} (2012), 297--309. \bibitem{CorTinAl} R. B. Corcino, C. B. Corcino, and R. M. Aldema, Asymptotic normality of the $(r,\beta)$-Stirling numbers, \textit{Ars Combin.}, \textbf{81} (2006), 81--96. \bibitem{CorHer} R. B. Corcino and M. Hererra, The limit of the differences of the generalized factorial, its $q$- and $p,q$-analogue, \textit{Util. Math.}, \textbf{72} (2007), 33--49. \bibitem{CorMon} R. B. Corcino and C. B. Montero, A $q$-analogue of Rucinski-Voigt numbers, \textit{ISRN Discrete Math.}, \textbf{2012}, Article ID 592818, 18 pages, (2012). DOI: \href{http://dx.doi.org/10.5402/2012/592818}{doi:10.5402/2012/592818}. \bibitem{CorMonSus} R. B. Corcino, M. B. Montero, and S. Ballenas, Schlomilch-type formula for $r$-Whitney numbers of the first kind, \textit{Mati\-my\-\'as Mat.}, \textbf{37} (2014), 1--10. \bibitem{Dowl} T. A. Dowling, A class of geometric lattices based on finite groups, \textit{J. Combin. Theory, Ser. B}, \textbf{15} (1973), 61--86. \bibitem{Graham} R. L. Graham, D. E. Knuth, and O. Patashnik, \textit{Concrete Mathematics}, Addison-Wesley, Reading, 1990. \bibitem{Kat1} J. Katriel, Stirling number identities: interconsistency of $q$-analogues, \textit{J. Phys. A: Math. Gen.}, \textbf{31} (1998), 3559--3572. \bibitem{Kat2} J. Katriel, Bell numbers and coherent states, \textit{Phys. Letters A}, \textbf{273} (2000), 159--161. \bibitem{Kim} M.-S. Kim and J.-W. Son, A note on $q$-difference operators, \textit{Commun. Korean Math. Soc.}, \textbf{17} (2002), 432--430. \bibitem{Mah1} M. M. Mangontarum, A. P.-M. Ringia, and N. S. Abdulcarim, The translated Dowling polynomials and numbers, \textit{International Scholarly Research Notices}, \textbf{2014}, Article ID 678408, 8 pages, (2014). DOI: \href{http://dx.doi.org/10.1155/2014/678408}{doi:10.1155/2014/678408}. \bibitem{Merca} M. Merca, A new connection between $r$-Whitney numbers and Bernoulli polynomials, \textit{Integral Transforms Spec. Funct.}, \textbf{25} (2014), 937--942. \bibitem{Mez1} I. Mez\H{o}, A new formula for the Bernoulli polynomials, \textit{Results Math.}, \textbf{58} (2010), 329--335. \bibitem{Mez} I. Mez\H{o}, The $r$-Bell Numbers, \textit{J. Integer Seq.}, \textbf{14} (2011), \href{https://cs.uwaterloo.ca/journals/JIS/VOL14/Mezo/mezo9.html}{Article 11.1.1}. \bibitem{Stir} J. Stirling, {\it Methodus Differentialissme Tractus de Summatione et Interpolatione Serierum Infinitarum}, \textit{London}, 1730. \end{thebibliography} \bigskip \hrule \bigskip \noindent 2010 {\it Mathematics Subject Classification}: Primary 11B83; Secondary 11B73, 05A30. \noindent \emph{Keywords:} Whitney number, Stirling number, Boson operator, $q$-analogue. \bigskip \hrule \bigskip \noindent (Concerned with sequences \seqnum{A000110}, \seqnum{A003575}, \seqnum{A008275}, and \seqnum{A008277}.) \bigskip \hrule \bigskip \vspace*{+.1in} \noindent Received May 13 2015; revised version received, September 1 2015. Published in {\it Journal of Integer Sequences}, September 7 2015. \bigskip \hrule \bigskip \noindent Return to \htmladdnormallink{Journal of Integer Sequences home page}{http://www.cs.uwaterloo.ca/journals/JIS/}. \vskip .1in \end{document} .