Skip to main content

Theory and Modern Applications

Theory of nth-order linear general quantum difference equations

Abstract

In this paper, we derive the solutions of homogeneous and non-homogeneous nth-order linear general quantum difference equations based on the general quantum difference operator \(D_{\beta }\) which is defined by \(D_{\beta }{f(t)}= (f(\beta (t))-f(t) )/ (\beta (t)-t )\), \(\beta (t)\neq t\), where β is a strictly increasing continuous function defined on an interval \(I\subseteq \mathbb{R}\) that has only one fixed point \(s_{0}\in {I}\). We also give the sufficient conditions for the existence and uniqueness of solutions of the β-Cauchy problem of these equations. Furthermore, we present the fundamental set of solutions when the coefficients are constants, the β-Wronskian associated with \(D_{\beta }\), and Liouville’s formula for the β-difference equations. Finally, we introduce the undetermined coefficients, the variation of parameters, and the annihilator methods for the non-homogeneous β-difference equations.

1 Introduction

Quantum difference operator allows us to deal with sets of non-differentiable functions. Its applications are used in many mathematical fields such as the calculus of variations, orthogonal polynomials, basic hypergeometric functions, quantum mechanics, and the theory of scale relativity; see, e.g., [3, 5, 7, 13, 14].

The general quantum difference operator \(D_{\beta }\) generalizes the Jackson q-difference operator \(D_{q}\) and the Hahn difference operator \(D_{q,\omega }\), see [1, 2, 4, 8, 12]. It is defined, in [10, p. 6], by

$$ {D}_{\beta }f(t)=\textstyle\begin{cases} \frac{f(\beta (t))-f(t)}{\beta (t)-t},& {t}\neq {s_{0}}, \\ {{f'}(s_{0})},& {t}={s_{0}}, \end{cases} $$

where \(f:I\rightarrow \mathbb{X}\) is a function defined on an interval \(I\subseteq {\mathbb{R}}\), \(\mathbb{X}\) is a Banach space, and \(\beta:I\rightarrow I\) is a strictly increasing continuous function defined on I that has only one fixed point \(s_{0}\in {I}\) and satisfies the inequality \((t-s_{0})(\beta (t)-t)\leq 0\) for all \(t\in I\). The function f is said to be β-differentiable on I if the ordinary derivative \({f'}\) exists at \(s_{0}\). The general quantum difference calculus was introduced in [10]. The exponential, trigonometric, and hyperbolic functions associated with \(D_{\beta }\) were presented in [9]. The existence and uniqueness of solutions of the first-order β-initial value problem were established in [11]. In [6], the existence and uniqueness of solutions of the β-Cauchy problem of the second-order β-difference equations were proved. Also, a fundamental set of solutions for the second-order linear homogeneous β-difference equations when the coefficients are constants was constructed, and the different cases of the roots of their characteristic equations were studied. Moreover, the Euler–Cauchy β-difference equation was derived.

The organization of this paper is briefly summarized in the following. In Sect. 2, we present the needed preliminaries of the β-calculus from [6, 911]. In Sect. 3, we give the sufficient conditions for the existence and uniqueness of solutions of the β-Cauchy problem of the nth-order β-difference equations. Also, we construct the fundamental set of solutions for the homogeneous linear β-difference equations when the coefficients \(a_{j}\) (\(0\leq j \leq n\)) are constants. Furthermore, we introduce the β-Wronskian which is an effective tool to determine whether the set of solutions is a fundamental set or not and prove its properties. Finally, we study the undetermined coefficients, the variation of parameters, and the annihilator methods for the non-homogeneous linear β-difference equations.

Throughout this paper, J is a neighborhood of the unique fixed point \(s_{0}\) of β, \(S(y_{0}, b)=\{y\in \mathbb{X}:\Vert y-y_{0}\Vert \leq b \}\), and \(R=\{(t,y)\in {{I}\times \mathbb{X}}:|t-s_{0}|\leq {a},\Vert y-y _{0}\Vert \leq {b}\}\) is a rectangle, where a, b are fixed positive real numbers, \(\mathbb{X}\) is a Banach space. Furthermore, \(D_{\beta }^{n}f=D _{\beta }(D_{\beta }^{n-1}f)\), \(n\in \mathbb{N}_{0}=\mathbb{N} \cup \{0\}\), where f is β-differentiable n times over I, \(\mathbb{N}\) is the set of natural numbers. We use the symbol T for the transpose of the vector or the matrix.

2 Preliminaries

Lemma 2.1

([10])

The following statements are true:

  1. (i)

    The sequence of functions \(\{\beta^{k}(t)\}_{k=0}^{\infty }\) converges uniformly to the constant function \(\hat{\beta }(t):=s _{0}\) on every compact interval \(V \subseteq I\) containing \(s_{0}\).

  2. (ii)

    The series \(\sum_{k=0}^{\infty }|\beta^{k}(t)-\beta^{k+1}(t)|\) is uniformly convergent to \(|t-s_{0}| \) on every compact interval \(V \subseteq I\) containing \(s_{0}\).

Lemma 2.2

([10])

If \(f:I\rightarrow \mathbb{X} \) is a continuous function at \(s_{0}\), then the sequence \(\{f(\beta^{k}(t))\}_{k=0}^{\infty }\) converges uniformly to \(f(s_{0})\) on every compact interval \(V\subseteq I\) containing \(s_{0}\).

Theorem 2.3

([10])

If \(f:I \rightarrow \mathbb{X} \) is continuous at \(s_{0}\), then the series \(\sum_{k=0}^{\infty }\| (\beta^{k}(t)- \beta^{k+1}(t) ) f(\beta^{k}(t))\|\) is uniformly convergent on every compact interval \(V \subseteq I\) containing \(s_{0}\).

Theorem 2.4

([10])

Assume that \(f:{I}\rightarrow \mathbb{X}\) and \(g:{I} \rightarrow \mathbb{R}\) are β-differentiable at \(t\in {I}\). Then:

  1. (i)

    The product \(fg:I\rightarrow \mathbb{X}\) is β-differentiable at t and

    $$\begin{aligned} {D}_{\beta }(fg) (t) &=\bigl({D}_{\beta }f(t) \bigr)g(t)+f\bigl(\beta (t)\bigr){D}_{\beta }g(t) \\ & =\bigl({D}_{\beta }f(t)\bigr)g\bigl(\beta (t)\bigr)+f(t){D}_{\beta }g(t), \end{aligned} $$
  2. (ii)

    \(f/g\) is β-differentiable at t and

    $$ {D}_{\beta } ({f}/{g} ) (t)=\frac{({D}_{\beta }f(t))g(t)-f(t) {D}_{\beta }g(t)}{g(t)g(\beta (t))}, $$

    provided that \(g(t)g(\beta (t))\neq {0}\).

Theorem 2.5

([10])

Assume that \(f:{I}\to \mathbb{X}\) is continuous at \(s_{0}\). Then the function F defined by

$$ F(t)=\sum_{k=0}^{\infty } \bigl( \beta^{k}(t)-\beta^{k+1}(t) \bigr)f\bigl(\beta ^{k}(t) \bigr), \quad t\in {I} $$
(2.1)

is a β-antiderivative of f with \(F(s_{0})=0\). Conversely, a β-antiderivative F of f vanishing at \(s_{0}\) is given by (2.1).

Definition 2.6

([10])

The β-integral of \(f:{I}\rightarrow {\mathbb{X}}\) from a to b, \(a,b\in {I}\), is defined by

$$ \int^{b}_{a}f(t)\,d_{\beta }{t}= \int^{b}_{s_{0}}f(t)\,d_{\beta }{t}- \int ^{a}_{s_{0}}f(t)\,d_{\beta }{t}, $$

where

$$ \int^{x}_{s_{0}}f(t)\,d_{\beta } {t}=\sum ^{\infty }_{k=0} \bigl(\beta^{k}(x)- \beta^{k+1}(x) \bigr)f\bigl(\beta ^{k}(x)\bigr),\quad x\in {I}, $$

provided that the series converges at \(x=a\) and \(x=b\). f is called β-integrable on I if the series converges at a and b for all \(a,b\in {I}\). Clearly, if f is continuous at \(s_{0}\in {I}\), then f is β-integrable on I.

Definition 2.7

([9])

The β-exponential functions \(e_{p,\beta }(t)\) and \(E_{p,\beta }(t)\) are defined by

$$\begin{aligned} e_{p,\beta }(t)=\frac{1}{\prod_{k=0}^{\infty }[1-p(\beta^{k} (t))( \beta^{k}(t)-\beta^{k+1}(t))]} \end{aligned}$$
(2.2)

and

$$\begin{aligned} E_{p,\beta }(t)=\prod_{k=0}^{\infty } \bigl[1+ p\bigl(\beta^{k}(t)\bigr) \bigl(\beta ^{k} (t) - \beta^{k+1}(t) \bigr) \bigr], \end{aligned}$$
(2.3)

where \(p:I \rightarrow \mathbb{C}\) is a continuous function at \(s_{0}\), \(e_{p,\beta }(t)=\frac{1}{E_{-p,\beta }(t)}\).

The both products in (2.2) and (2.3) are convergent to a non-zero number for every \(t\in I\) since \(\sum_{k=0}^{\infty } | p( \beta^{k}(t)) (\beta^{k}(t)-\beta^{k+1}(t) ) |\) is uniformly convergent.

Definition 2.8

([9])

The β-trigonometric functions are defined by

$$\begin{aligned}& \cos_{p,\beta }(t) =\frac{ e_{ip,\beta }(t)+e_{-ip,\beta }(t)}{2}, \\& \sin_{p,\beta } (t) =\frac{ e_{ip,\beta }(t)-e_{-ip,\beta }(t)}{2i}, \\& \operatorname{Cos}_{p,\beta } (t) =\frac{E_{ip,\beta }(t)+E_{-ip,\beta }(t)}{2}, \\& \text{and} \quad \operatorname{Sin}_{p,\beta }(t) =\frac{E_{ip,\beta }(t)-E_{-ip,\beta } (t)}{2i}. \end{aligned}$$

Theorem 2.9

([9])

The β-exponential functions \(e_{p,_{\beta }}(t)\) and \(E_{p,_{\beta }}(t)\) are the unique solutions of the first-order β-difference equations

$$\begin{aligned}& D_{\beta }y(t)=p(t)y(t), \qquad y(s_{0})=1, \\& D_{\beta }y(t)=p(t)y\bigl(\beta (t)\bigr), \qquad y(s_{0})=1, \end{aligned}$$

respectively.

Theorem 2.10

([9])

Assume that \(f:I\rightarrow \mathbb{X}\) is continuous at \(s_{0}\). Then the solution of the following equation \(D_{\beta }y(t)= p(t)y(t)+f(t)\), \(y(s_{0})=y_{0}\in \mathbb{X}\), has the form

$$ y(t)=e_{p,\beta }(t) \biggl[y_{0}+ \int_{s_{0}}^{t}f(\tau)E_{-p,\beta }\bigl( \beta ( \tau)\bigr)\,d_{\beta }\tau \biggr]. $$

Theorem 2.11

([11])

Let \(z\in \mathbb{C}\) be a constant. Then the function \(\phi (t)\) defined by

$$ \phi (t)=\sum_{k=0}^{\infty }z^{k} \alpha_{k}(t) $$

is the unique solution of the β-IVP

$$ D_{\beta }{y(t)}=zy(t),\qquad y(s_{0})=1, $$

where

$$\alpha_{k}(t) =\textstyle\begin{cases} \sum_{i_{1},i_{2},i_{3},\dots,i_{k-1}=0}^{\infty } (\prod_{l=1} ^{k-1}(\beta,\beta)_{\sum_{j=1}^{l}{i_{j}}} ) ( \beta^{\sum_{j=1}^{k-1}{i_{j}}}(t)-s_{0} ),&\textit{if } k\geq 2, \\ t-s_{0},&\textit{if } k=1, \\ 1,&\textit{if } k=0, \end{cases} $$

with \((\beta,\beta)_{i}=\beta^{i}(t)-\beta^{i+1}(t)\).

Proposition 2.12

([11])

Let \(z\in \mathbb{C}\). The β-exponential function \(e_{z,\beta }(t)\) has the expansion

$$ e_{z,\beta }(t)=\sum_{k=0}^{\infty }z^{k} \alpha_{k}(t). $$

Theorem 2.13

([11])

Assume that \(f:R\rightarrow {\mathbb{X}}\) is continuous at \((s_{0},y_{0})\in {R}\) and satisfies the Lipschitz condition (with respect to y)

$$ \bigl\Vert f(t,y_{1})-f(t,y_{2})\bigr\Vert \leq {L} \Vert y_{1}-y_{2}\Vert \quad \textit{for all } (t,y_{1}),(t,y_{2})\in {R}, $$

where L is a positive constant. Then the sequence defined by

$$ \phi_{k+1}(t)=y_{0}+ \int_{s_{0}}^{t}f \bigl(\tau,\phi_{k}(\tau) \bigr)\,d_{\beta }{\tau },\qquad \phi_{0}(t)=y_{0}, \quad \vert t-s_{0}\vert \leq {\delta }, k\geq {0} $$
(2.4)

converges uniformly on the interval \(|t-s_{0}|\leq {\delta }\) to a function ϕ, the unique solution of the β-IVP

$$ D_{\beta }{y(t)}=f(t,y),\qquad y(s_{0})=y_{0}, \quad t\in {I}, $$
(2.5)

where \(\delta =\min \{a,\frac{b}{Lb+M}, \frac{\rho }{L}\}\) with \(\rho \in (0,1)\) and \(M=\sup_{(t,y)\in {R}}\Vert f(t,y)\Vert <\infty \), \(\rho \in (0,1)\).

Theorem 2.14

([6])

Let \(f_{i}(t,y_{1},y_{2}):I \times \prod_{i=1}^{2} S_{i}(x _{i}, b_{i})\rightarrow {\mathbb{X}}\), \(s_{0}\in I\) such that the following conditions are satisfied:

  1. (i)

    For \(y_{i}\in S_{i}(x_{i},b_{i})\), \(i=1,2\), \(f_{i}(t,y_{1},y _{2})\) are continuous at \(t=s_{0}\).

  2. (ii)

    There is a positive constant A such that, for \(t\in I\), \(y_{i}, \tilde{y}_{i}\in S_{i}(x_{i},b_{i})\), \(i=1,2\), the following Lipschitz condition is satisfied:

$$ \bigl\Vert f_{i}(t,y_{1},y_{2})-f_{i}(t, \tilde{y}_{1},\tilde{y}_{2}) \bigr\Vert \leq A \sum _{i=1}^{2} \Vert y_{i}- \tilde{y}_{i} \Vert . $$

Then there exists a unique solution of the β-initial value problem β-IVP

$$\begin{aligned} D_{\beta }y_{i}(t)=f_{i}\bigl(t,y_{1}(t),y_{2}(t) \bigr), \qquad y_{i}(s_{0})=x_{i}\in { \mathbb{X}},\quad i =1,2, t \in I. \end{aligned}$$

Corollary 2.15

([6])

Let \(f(t,y_{1},y_{2})\) be a function defined on \(I\times \prod_{i=1}^{2} S_{i}(x_{i},b_{i})\) such that the following conditions are satisfied:

  1. (i)

    For any values of \(y_{i}\in S_{i}(x_{i},b_{i})\), \(i=1,2\), f is continuous at \(t=s_{0}\).

  2. (ii)

    f satisfies the Lipschitz condition

    $$ \bigl\Vert f(t,y_{1},y_{2})-f(t,\tilde{y}_{1}, \tilde{y}_{2}) \bigr\Vert \leq A\sum_{i=1}^{2} \Vert y_{i} -\tilde{y}_{i} \Vert , $$

where \(A>0\), \(y_{i},\tilde{y}_{i}\in S_{i}(x_{i},b_{i})\), \(i=1,2\), and \(t \in I\). Then

$$\begin{aligned} D_{\beta }^{2}y(t)=f\bigl(t,y(t),D_{\beta }y(t)\bigr), \qquad D_{\beta }^{i-1}y(s_{0})=x_{i},\quad i=1,2, \end{aligned}$$

has a unique solution on \([s_{0},s_{0} +\delta ]\).

Corollary 2.16

([6])

Assume that the functions \(a_{j}(t):I\rightarrow \mathbb{C}\), \(j=0,1,2\), and \(b(t):I\rightarrow {\mathbb{X}}\) satisfy the following conditions:

  1. (i)

    \(a_{j}(t)\), \(j=0,1,2\), and \(b(t)\) are continuous at \(s_{0}\) with \(a_{0}(t)\neq 0\) for all \(t \in I\),

  2. (ii)

    \(a_{j}(t)/a_{0}(t)\) is bounded on I, \(j=1,2\). Then

$$ a_{0}(t)D_{\beta }^{2}y(t)+ a_{1}(t)D_{\beta }y(t)+a_{2}(t)y(t)=b(t), \qquad D_{\beta }^{i-1}y(s_{0})= x_{i},\quad x_{i} \in {\mathbb{X}}, i=1,2, $$

has a unique solution on a subinterval \(J\subseteq I\), \(s_{0}\in J\).

3 Main results

In this section, we give the sufficient conditions for the existence and uniqueness of solutions of the β-Cauchy problem of the nth-order β-difference equations. We also present the fundamental set of solutions for the homogeneous linear β-difference equations when the coefficients \(a_{j}\) (\(0\leq j \leq n \)) are constants. Furthermore, we introduce the β-Wronskian. Finally, we study the undetermined coefficients, the variation of parameters, and the annihilator methods for the non-homogeneous linear β-difference equations.

3.1 Existence and uniqueness of solutions

Theorem 3.1

Let I be an interval containing \(s_{0}\), \(f_{i}(t,y_{1},\ldots,y _{n}):I \times \prod_{i=1}^{n}S_{i}(x_{i},b_{i})\rightarrow \mathbb{X}\), such that the following conditions are satisfied:

  1. (i)

    For \(y_{i}\in S_{i}(x_{i},b_{i})\), \(i=1,\ldots,n\), \(f_{i}(t,y_{1},\ldots,y_{n})\) are continuous at \(t=s_{0}\).

  2. (ii)

    There is a positive constant A such that, for \(t \in I\), \(y_{i},\tilde{y}_{i}\in S_{i}(x_{i},b_{i})\), \(i=1,\ldots,n\), the following Lipschitz condition is satisfied:

$$ \bigl\Vert f_{i}(t,y_{1},\ldots,y_{n})-f_{i}(t, \tilde{y}_{1},\ldots,\tilde{y}_{n}) \bigr\Vert \leq A \sum _{i=1}^{n} \Vert y_{i}- \tilde{y}_{i} \Vert . $$

Then there exists a unique solution of the β-initial value problem β-IVP

$$\begin{aligned} D_{\beta }y_{i}(t)=f_{i}\bigl(t,y_{1}(t), \ldots,y_{n}(t)\bigr),\qquad y_{i}(s_{0})=x_{i} \in \mathbb{X},\quad i=1,\ldots,n, t \in I. \end{aligned}$$

Proof

See the proof of Theorem 2.14. □

The proof of the following two corollaries is the same as the proof of Corollaries 2.15, 2.16.

Corollary 3.2

Let \(f(t,y_{1},\ldots,y_{n})\) be a function defined on \(I\times \prod_{i=1}^{n} S_{i}(x_{i},b_{i})\) such that the following conditions are satisfied:

  1. (i)

    For any values of \(y_{r}\in S_{r}(x_{r},b_{r})\), f is continuous at \(t=s_{0}\).

  2. (ii)

    f satisfies the Lipschitz condition

    $$ \bigl\Vert f (t,y_{1},\ldots,y_{n})-f(t, \tilde{y}_{1},\ldots,\tilde{y}_{n}) \bigr\Vert \leq A\sum _{i=1}^{n} \Vert y_{i}- \tilde{y}_{i} \Vert , $$

where \(A>0\), \(y_{i},\tilde{y}_{i}\in S_{i}(x_{i},b_{i})\), \(i=1, \ldots,n\), and \(t \in I\). Then

$$\begin{aligned}& \begin{aligned} &D_{\beta }^{n} y(t) =f \bigl(t,y(t),D_{\beta }y(t),\ldots,D_{\beta }^{n-1}y(t)\bigr), \\ &D_{\beta }^{i-1}y(s_{0}) =x_{i},\quad i=1, \ldots,n, \end{aligned} \end{aligned}$$
(3.1)

has a unique solution on \([s_{0},s_{0} +\delta ]\).

The following corollary gives us the sufficient conditions for the existence and uniqueness of solutions of the β-Cauchy problem (3.1).

Corollary 3.3

Assume that the functions \(a_{j}(t):I\rightarrow \mathbb{C}\), \(j=0,1, \ldots,n\), and \(b(t):I\rightarrow \mathbb{X}\) satisfy the following conditions:

  1. (i)

    \(a_{j}(t)\), \(j=0,1,\ldots,n\), and \(b(t)\) are continuous at \(s_{0}\) with \(a_{0}(t)\neq 0\) for all \(t \in I \),

  2. (ii)

    \(a_{j}(t)/a_{0}(t)\) is bounded on I, \(j=1,\ldots,n\). Then

$$\begin{aligned}& a_{0}(t)D_{\beta }^{n}y(t)+a_{1}(t)D_{\beta }^{n-1}y(t)+ \cdots +a_{n}(t)y(t)=b(t), \\& D_{\beta }^{i-1}y(s_{0})=x_{i},\quad i=1, \ldots,n, \end{aligned}$$

has a unique solution on a subinterval \(J\subset I\) containing \(s_{0}\).

3.2 Homogeneous linear β-difference equations

Consider the nth-order homogeneous linear β-difference equation

$$ a_{0}(t)D_{\beta }^{n}y(t)+a_{1}(t)D_{\beta }^{n-1}y(t)+ \cdots + a _{n-1}(t)D_{\beta }y(t)+a_{n}(t)y(t)=0, $$
(3.2)

where the coefficients \(a_{j}(t)\), \(0\leq j\leq n\), are assumed to satisfy the conditions of Corollary 3.3. Equation (3.2) may be written as \(L_{n}y=0\), where

$$ L_{n}=a_{0}(t)D_{\beta }^{n}+a_{1}(t)D_{\beta }^{n-1}+ \cdots +a_{n-1}(t)D _{\beta }+a_{n}(t). $$

The following lemma is an immediate consequence of Corollary 3.3.

Lemma 3.4

If y is a solution of equation (3.2) such that \(D_{\beta } ^{i-1}y(s_{0})=0\), \(1\leq i\leq n\), then \(y(t)=0\) for all \(t\in J\).

Theorem 3.5

The nth-order homogeneous linear scalar β-difference equation (3.2) is equivalent to the first-order homogeneous linear system of the form

$$ D_{\beta }Y(t)=A(t)Y(t), $$

where

$$\begin{aligned} Y= \left ( \begin{matrix} y_{1} \\ \vdots \\ y_{n} \end{matrix} \right ) \quad \textit{and}\quad A= \left ( \begin{matrix} 0 & 1 & \ldots & 0 \\ \vdots & \vdots & \ldots & \vdots \\ 0 & 0 & & 1 \\ -\frac{a_{n}}{a_{0}} & -\frac{a_{n-1}}{a_{0}} & \ldots & -\frac{a_{1}}{a _{0}} \end{matrix} \right ) . \end{aligned}$$

Proof

Let

$$\begin{aligned}& \begin{aligned} &y_{1} =y, \\ &y_{2} = D_{\beta }y, \\ &\vdots \\ &y_{n-1} =D_{\beta }^{n-2}y, \\ &y_{n} =D_{\beta }^{n-1}y. \end{aligned} \end{aligned}$$
(3.3)

β-differentiating (3.3), we have

$$\begin{aligned} D_{\beta }y=D_{\beta }y_{1} ,\qquad D_{\beta }^{2}y=D_{\beta }y_{2}, \qquad \ldots,\qquad D_{\beta }^{n-1}y=D_{\beta }y_{n-1}, \qquad D_{\beta }^{n}y=D_{\beta }y_{n}. \end{aligned}$$
(3.4)

Then

$$\begin{aligned} D_{\beta }y_{1}=y_{2}, \qquad D_{\beta }y_{2}=y_{3}, \qquad \ldots,\qquad D_{\beta }y _{n-1}=y_{n}. \end{aligned}$$
(3.5)

Since \(a_{0}(t) \neq 0\) on J, (3.2) is equivalent to

$$ D_{\beta }^{n}y=-\frac{a_{n}(t)}{a_{0}(t)}y- \frac{a_{n-1}(t)}{a_{0}(t)}D_{\beta }y- \cdots - \frac{a_{1}(t)}{a_{0}(t)}D_{\beta }^{n-1} y, $$

from (3.3) and (3.4), we have

$$\begin{aligned} D_{\beta }y_{n}= -\frac{a_{n}(t)}{a_{0}(t)}y_{1}- \frac{a_{n-1}(t)}{a _{0}(t)}y_{2}-\cdots -\frac{a_{1}(t)}{a_{0}(t)}y_{n}. \end{aligned}$$
(3.6)

Combining (3.5) and (3.6), we get

$$\begin{aligned}& \begin{aligned} &D_{\beta }y_{1}= y_{2}, \\ &\vdots \\ &D_{\beta }y_{n-1}= y_{n}, \\ &D_{\beta }y_{n}= -\frac{a_{n}(t)}{a_{0}(t)}y_{1}- \frac{a_{n-1}(t)}{a _{0}(t)}y_{2}-\cdots -\frac{a_{1}(t)}{a_{0}(t)}y_{n}. \end{aligned} \end{aligned}$$
(3.7)

This is equivalent to the homogeneous linear vector β-difference equation

$$\begin{aligned} D_{\beta }Y(t)=A(t)Y(t), \end{aligned}$$
(3.8)

where

$$\begin{aligned} Y= \left ( \begin{matrix} y_{1} \\ \vdots \\ y_{n} \end{matrix} \right ) \quad \textit{and}\quad A= \left ( \begin{matrix} 0 & 1 & \ldots & 0 \\ \vdots & \vdots & \ldots & \vdots \\ 0 & 0 & & 1 \\ -\frac{a_{n}}{a_{0}} & -\frac{a_{n-1}}{a_{0}} & \ldots & -\frac{a_{1}}{a _{0}} \end{matrix} \right ). \end{aligned}$$

 □

Theorem 3.6

Consider equation (3.2) and the corresponding system (3.8). If f is a solution of (3.2) on J, then \(\phi = (f,D_{\beta }f,\ldots,D_{\beta }^{n-1}f )^{T}\) is a solution of (3.8) on J. Conversely, if \(\phi = (\phi_{1}, \ldots,\phi_{n} )^{T}\) is a solution of (3.8) on J, then its first component \(\phi_{1}\) is a solution f of (3.2) on J and \(\phi = (f,D_{\beta }f,\ldots,D_{\beta }^{n-1}f )^{T}\).

Proof

Suppose that f satisfies equation (3.2). Then

$$\begin{aligned} a_{0}(t)D_{\beta }^{n}f(t)+\cdots +a_{n-1}(t)D_{\beta }f(t)+a_{n}(t)f(t)=0,\quad t\in J. \end{aligned}$$
(3.9)

Consider

$$\begin{aligned} \phi (t)= \bigl(\phi_{1}(t),\ldots, \phi_{n}(t) \bigr)^{T}= \bigl(f(t),D_{\beta }f(t), \ldots,D_{\beta }^{n-1}f(t) \bigr)^{T}. \end{aligned}$$
(3.10)

From (3.9) and (3.10), we have

$$\begin{aligned} \begin{aligned}&D_{\beta }\phi_{1}(t) =\phi_{2}(t), \\ &\vdots \\ &D_{\beta }\phi_{n-1}(t) =\phi_{n}(t), \\ &D_{\beta }\phi_{n}(t) = -\frac{a_{n}(t)}{a_{0}(t)} \phi_{1}(t)-\frac{a _{n-1}(t)}{a_{0}(t)}\phi_{2}(t)-\cdots - \frac{a_{1}(t)}{a_{0}(t)}\phi _{n}(t). \end{aligned} \end{aligned}$$
(3.11)

Comparing (3.11) with (3.7), ϕ defined by (3.10) satisfies system (3.7). Conversely, suppose that \(\phi (t)= (\phi_{1}(t),\ldots,\phi_{n}(t) )^{T}\) satisfies system (3.7) on J. Then (3.11) holds for all \(t \in J\). The first \(n - 1\) equations of (3.11) give

$$\begin{aligned} \begin{aligned}&\phi_{2}(t) =D_{\beta }\phi_{1}(t), \\ &\phi_{3}(t) =D_{\beta }\phi_{2}(t)=D_{\beta }^{2} \phi_{1}(t), \\ &\vdots \\ &\phi_{n}(t) = D_{\beta }\phi_{n-1}(t)=D_{\beta }^{2} \phi_{n-2}(t)= \cdots =D_{\beta }^{n-1} \phi_{1}(t), \end{aligned} \end{aligned}$$
(3.12)

and so \(D_{\beta }\phi_{n}(t)=D_{\beta }^{n}\phi_{1}(t)\). The last equation of (3.11) becomes

$$ a_{0}(t)D_{\beta }^{n}\phi_{1}(t)+a_{1}(t)D_{\beta }^{n-1} \phi_{1}(t)+ \cdots +a_{n-1}(t)D_{\beta } \phi_{1}(t) +a_{n}(t)\phi_{1}(t)=0. $$

Thus \(\phi_{1}\) is a solution f of equation (3.2); and moreover, (3.12) shows that \(\phi (t)= (f(t),D_{\beta }f(t), \ldots, D_{\beta }^{n-1}f(t) )^{T}\). □

The following corollary is an immediate consequence of Theorem 3.6.

Corollary 3.7

If f is the solution of equation (3.2) on J satisfying the initial condition \(D_{\beta }^{i-1}f(s_{0})=x_{i}\), \(1\leq i\leq n\), then \(\phi = (f,D_{\beta }f,\ldots,D_{\beta }^{n-1}f )^{T}\) is the solution of system (3.8) on J satisfying the initial condition \(\phi (s_{0})=(x_{1},\dots,x_{n})^{T}\). Conversely, if \(\phi =(\phi_{1},\dots,\phi_{n})^{T}\) is the solution of (3.8) on J satisfying the initial condition \(\phi (s_{0})=(x _{1},\dots,x_{n})^{T}\), then \(\phi_{1}\) is the solution f of (3.2) on J satisfying the initial condition \(D_{\beta }^{i-1}f(s _{0} ) =x_{i}\), \(1\leq i\leq n\).

Theorem 3.8

A linear combination \(y=\sum_{k=1}^{m}c_{k}y_{k}\) of m solutions \(y_{1},\ldots,y_{m}\) of the homogeneous linear β-difference equation (3.2) is also a solution of it, where \(c_{1},\ldots,c _{m}\) are arbitrary constants.

Proof

The proof is straightforward. □

Definition 3.9

(A fundamental set)

A set of n linearly independent solutions of the nth-order homogeneous linear β-difference equation (3.2) is called a fundamental set of equation (3.2).

By the theory of differential equations, we can easily prove the following theorems.

Theorem 3.10

If the solutions \(y_{1},\ldots,y_{n}\) of the homogeneous linear β-difference equation (3.2) are linearly independent on J, then the corresponding solutions

$$ \phi_{1}= \bigl(y_{1},D_{\beta }y_{1}, \ldots,D_{\beta }^{n-1}y_{1} \bigr)^{T},\qquad \ldots,\qquad \phi_{n}= \bigl(y_{n}, D_{\beta }y_{n}, \ldots, D_{ \beta }^{n-1}y_{n} \bigr)^{T} $$

of system (3.8) are linearly independent on J; and conversely.

Theorem 3.11

Any arbitrary solution y of homogeneous linear β-difference equation (3.2) on J can be represented as a suitable linear combination of a fundamental set of solutions \(y_{1},\ldots,y_{n}\) of (3.2).

Now, we are concerned with constructing the fundamental set of solutions of equation (3.2) when the coefficients are constants. Equation (3.2) can be written as

$$\begin{aligned} L_{n} y(t)=a_{0} D_{\beta }^{n}y(t)+a_{1}D_{\beta }^{n-1}y(t)+ \cdots +a_{n} y(t)=0, \end{aligned}$$
(3.13)

where \(a_{j}\), \(0\leq j \leq n\), are constants. From Theorem 3.5, equation (3.13) is equivalent to the system

$$ D_{\beta }{Y(t)}=AY(t), $$
(3.14)

where

$$ A= \left ( \begin{matrix} 0 & 1 & \ldots & 0 \\ \vdots & \vdots & \ldots & \vdots \\ 0 & 0 & & 1 \\ -\frac{a_{n}}{a_{0}} & -\frac{a_{n-1}}{a_{0}} & \ldots &-\frac{a_{1}}{a _{0}} \end{matrix} \right ) . $$

The characteristic polynomial of equation (3.13) is given by

$$ P(\lambda)=\det (\lambda \mathcal{I}-A)=a_{0} \lambda^{n}+a_{1} \lambda^{n-1}+\cdots +a_{n}, $$
(3.15)

where \(\mathcal{I}\) is the unit square matrix of order n, \(\lambda_{i}\), \(1\leq i \leq k\), are distinct roots of \(p(\lambda)=0\) of multiplicity \(m_{i}\), so that \(\sum_{i=1}^{k}m_{i}=n\).

Theorem 3.12

Let A be a constant \(n\times n\) matrix. Then the function \(\Phi (t)\) defined by

$$ \Phi (t)=\sum_{r=0}^{\infty }A^{r} \alpha_{r}(t) $$

is the unique solution of the β-IVP

$$ D_{\beta }{Y(t)}=AY(t),\quad Y(s_{0})= \mathcal{I}, $$

where \(\mathcal{I}\) is the unit square matrix of order n and

$$\begin{aligned} \alpha_{r}(t)= \textstyle\begin{cases} \sum_{i_{1},i_{2},i_{3},\ldots,i_{r-1}=0}^{\infty }(\prod_{l=1}^{r-1}( \beta,\beta)_{\sum_{j=1}^{l}i_{j} })(\beta^{\sum_{j=1}^{r-1}i_{j}} {(t)-s_{0}}),& \textit{if } r \geq 2, \\ t-s_{0} & \textit{if } r=1, \\ \mathcal{I},& \textit{if } r=0, \end{cases}\displaystyle \end{aligned}$$

with \((\beta;\beta)_{i}=\beta_{i}(t)-\beta_{i+1}(t)\).

Proof

By using the successive approximations, with choosing \(\Phi_{0}(t)= \mathcal{I}\), we have the desired result. See the proof of Theorem 2.11. □

Corollary 3.13

Let A be a constant \(n\times n\) matrix with characteristic polynomial (3.15), then \(\Phi (t)=e_{A,\beta }(t)=\sum_{r=0}^{\infty }A^{r} \alpha_{r}(t)\) is the unique solution of (3.13) satisfying the initial conditions

$$ \Phi (s_{0})=\mathcal{I},\qquad D_{\beta }\Phi (s_{0})=A,\qquad \ldots,\qquad D_{\beta } ^{n-1}\Phi (s_{0})=A^{n-1}. $$

Proof

The proof is straightforward. □

We have from the previous that

$$ y_{i}(t)=e_{\lambda_{i},\beta }(t)=\sum_{r=0}^{\infty } \lambda_{i} ^{r}\alpha_{r}(t),\quad 1\leq i \leq k, $$

forms a fundamental set of solutions of equation (3.13).

Example 3.14

Consider the homogeneous linear system

$$\begin{aligned} D_{\beta }Y(t)= \left ( \begin{matrix} 3 &1 & -1 \\ 1& 3 & -1 \\ 3 & 3& -1 \end{matrix} \right ) Y(t). \end{aligned}$$
(3.16)

Let \(Y(t)=\gamma e_{\lambda,\beta }(t)\), where \(\gamma=(\gamma_{1},\ldots,\gamma_{n})^{T}\) is a constant vector. The characteristic equation is

$$ \lambda^{3}-5\lambda^{2}+8\lambda -4=0, $$

where \(\lambda_{1}=1\), \(\lambda_{2}=\lambda_{3}=2\). Then

$$ y_{1}(t)= \left ( \begin{matrix} 1 \\ 1 \\ 3 \end{matrix} \right ) e_{1,\beta }(t),\qquad y_{2}(t)= \left ( \begin{matrix} 1 \\ -1 \\ 0 \end{matrix} \right ) e_{2,\beta }(t) \quad \mbox{and}\quad y_{3}(t)= \left ( \begin{matrix} 1 \\ 0 \\ 1 \end{matrix} \right ) e_{2,\beta }(t) $$

are the solutions of (3.16). The general solution of system (3.16) is

$$ Y(t)=c_{1} \left ( \begin{matrix} e_{1,\beta }(t) \\ e_{1,\beta }(t) \\ 3 e_{1,\beta }(t) \end{matrix} \right ) +c_{2} \left ( \begin{matrix} e_{2,\beta }(t) \\ -e_{2,\beta }(t) \\ 0 \end{matrix} \right ) +c_{3} \left ( \begin{matrix} e_{2,\beta }(t) \\ 0 \\ e_{2,\beta }(t) \end{matrix} \right ) , $$

where \(c_{1}\), \(c_{2}\), and \(c_{3}\) are arbitrary constants.

Example 3.15

Consider the homogeneous linear system

$$\begin{aligned} D_{\beta }Y(t)= \left ( \begin{matrix} 4 & 3 & 1 \\ -4 & -4 & -2 \\ 8 & 12& 6 \end{matrix} \right ) Y(t). \end{aligned}$$
(3.17)

Assume that \(Y=\gamma e_{\lambda,\beta }(t)\). The characteristic equation is

$$ \lambda^{3}-6\lambda^{2}+12\lambda -8=0, $$

where \(\lambda_{1}=\lambda_{2}=\lambda_{3}=2\). Then

$$ y_{1}(t)= \left ( \begin{matrix} 1 \\ 0 \\ - 2 \end{matrix} \right ) e_{2,\beta }(t) \quad \mbox{and}\quad y_{2}(t)= \left ( \begin{matrix} 0 \\ 1 \\ - 3 \end{matrix} \right ) e_{2,\beta }(t). $$

Let \(y_{3}(t)=(\gamma t+\nu)e_{2,\beta }(t)\),

$$ \gamma= \left ( \begin{matrix} k_{1} \\ k_{2} \\ -2k_{1} - 3k_{2} \end{matrix} \right ) \quad \mbox{and}\quad \nu = \left ( \begin{matrix} \nu_{1} \\ \nu_{2} \\ \nu_{3} \end{matrix} \right ) , $$

where \(k_{1}\) and \(k_{1}\) are constants, and also γ and ν satisfy

$$ (A-\lambda \mathcal{I})\gamma =0 $$

and

$$ (A-\lambda \mathcal{I})\nu =\gamma. $$

Therefore,

$$ y_{3}(t)=\left [ \left ( \begin{matrix} 1 \\ -2 \\ 4 \end{matrix} \right ) t+ \left ( \begin{matrix} 0 \\ 0 \\ 1 \end{matrix} \right ) \right ] e_{2,\beta }(t). $$

The general solution of system (3.17) is

$$ Y(t)=c_{1} \left ( \begin{matrix} e_{2,\beta }(t) \\ 0 \\ - 2e_{2,\beta }(t) \end{matrix} \right ) +c_{2} \left ( \begin{matrix} 0 \\ e_{2,\beta }(t) \\ - 3e_{2,\beta }(t) \end{matrix} \right ) +c_{3} \left ( \begin{matrix} te_{2,\beta }(t) \\ -2te_{2,\beta }(t) \\ (4t + 1)e_{2,\beta }(t) \end{matrix} \right ) , $$

where \(c_{1}\), \(c_{2}\), \(c_{3}\) are arbitrary constants.

3.3 β-Wronskian

Definition 3.16

Let \(y_{1},\dots,y_{n}\) be β-differentiable functions \((n-1)\) times defined on I, then we define the β-Wronskian of the functions \(y_{1},\ldots,y_{n}\) by

$$ W_{\beta }(y_{1},\ldots,y_{n}) (t)=\left \vert \begin{matrix} y_{1}(t) & \ldots & y_{n}(t) \\ D_{\beta }y_{1}(t)& \ldots & D_{\beta }y_{n} (t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{n-1}y_{1}(t) & \ldots & D_{\beta }^{n-1}y_{n}(t) \end{matrix} \right \vert . $$

Throughout this paper, we write \(W_{\beta }\) instead of \(W_{\beta }(y _{1},\ldots,y_{n})\).

Lemma 3.17

Let \(y_{1}(t),\ldots,y_{n}(t)\) be n-times β-differentiable functions defined on I. Then, for any \(t\in I\), \(t \neq s_{0}\),

$$ D_{\beta }W_{\beta }(y_{1},\ldots,y_{n}) (t)= \left \vert \begin{matrix} y_{1}(\beta (t)) & \ldots & y_{n}(\beta (t)) \\ D_{\beta }y_{1} (\beta (t)) & \ldots & D_{\beta }y_{n} (\beta (t)) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{n-2}y_{1}(\beta (t)) & \ldots & D_{\beta }^{n-2} y_{n}( \beta (t)) \\ D_{\beta }^{n} y_{1}(t) & \ldots & D_{\beta }^{n} y_{n} (t) \end{matrix} \right \vert . $$
(3.18)

Proof

We prove by induction on n. The lemma is trivial when \(n =1\). Then suppose that it is true for \(n=k\). Our objective is to show that it holds for \(n=k+1\).

We expand \(W_{\beta }(y_{1},\ldots,y_{k+1})\) in terms of the first row to obtain

$$ W_{\beta }(y_{1},\ldots,y_{k+1})=\sum _{j=1}^{k+1}(-1)^{j+1} y_{j}(t) W_{\beta }^{(j)}(t), $$

where

$$ W_{\beta }^{(j)}=\textstyle\begin{cases} W_{\beta }(D_{\beta }y_{2},\ldots,D_{\beta }y_{k+1}),& j=1, \\ W_{\beta }(D_{\beta }y_{1},\ldots,D_{\beta }y_{j-1},D_{\beta }y_{j+1}, \ldots,D_{\beta }y_{k+1} ),& 2\leq j\leq k, \\ W_{\beta }(D_{\beta }y_{1},\ldots,D_{\beta }y_{k}),& j=k+1. \end{cases} $$

Consequently,

$$ D_{\beta }W_{\beta }(y_{1},\ldots,y_{k+1}) (t)= \sum_{j=1}^{k+1}(-1)^{j+1}D _{\beta }y_{j}(t)W_{\beta }^{(j)} (t)+\sum _{j=1}^{k+1}(-1)^{j+1}y_{j} \bigl( \beta (t)\bigr) D_{\beta }W_{\beta }^{(j)}(t). $$

We have

$$ \sum_{j=1}^{k+1}(-1)^{j+1}D_{\beta }y_{j} (t)W_{\beta }^{(j)}(t)= \left \vert \begin{matrix} D_{\beta }y_{1}(t) &\ldots &D_{\beta }y_{k+1}(t) \\ D_{\beta }y_{1}(t) & \ldots & D_{\beta }y_{k+1}(t) \\ D_{\beta }^{2} y_{1}(t) & \ldots & D_{\beta }^{2} y_{k+1}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{k-1} y_{1}(t) & \ldots & D_{\beta }^{k-1}y_{k+1}(t) \\ D_{\beta }^{k} y_{1}(t) & \ldots & D_{\beta }^{k}y_{k+1}(t) \end{matrix} \right \vert =0, $$

and from the induction hypothesis we have

$$\begin{aligned}& \sum_{j=1}^{k+1}(-1)^{j+1}y_{j} \bigl(\beta (t)\bigr)D_{\beta }W_{\beta }^{(j)}(t) \\& \quad = \sum_{j=1}^{k+1}(-1)^{j+1}y_{j} \bigl(\beta (t)\bigr) \\& \qquad {}\times \left \vert \begin{matrix} D_{\beta }y_{1}(\beta (t)) & \ldots & D_{\beta }y_{j-1}(\beta (t)) & D _{\beta }y_{j+1}(\beta (t)) & \ldots &D_{\beta }y_{k+1}(\beta (t)) \\ D_{\beta }^{2} y_{1}(\beta (t)) & \ldots &D_{\beta }^{2} y_{j-1}( \beta (t)) &D_{\beta }^{2} y_{j+1}(\beta (t)) & \ldots & D_{\beta } ^{2}y_{k+1}(\beta (t)) \\ \vdots & \ddots & \vdots & \ddots & \vdots & \vdots \\ D_{\beta }^{k-1} y_{1}(\beta (t)) & \ldots &D_{\beta }^{k-1} y_{j-1}( \beta (t)) & D_{\beta }^{k-1}y_{j+1}(\beta (t)) & \ldots & D_{\beta } ^{k-1} y_{k+1}(\beta (t)) \\ D_{\beta }^{k+1}y_{1}(t) & \ldots & D_{\beta }^{k+1}y_{j-1}(t) &D_{ \beta }^{k+1}y_{j+1}(t) & \ldots & D_{\beta }^{k+1} y_{k+1}(t) \end{matrix} \right \vert , \end{aligned}$$
(3.19)

where at \(j=1\) the determinant of (3.19) starts with \(D_{\beta }y_{2}(\beta (t))\) and at \(j= k+1\) the determinant ends with \(D_{\beta }^{k+1}y_{k}(t)\). So,

$$ \sum_{j=1}^{k+1}(-1)^{j+1}y_{j} \bigl(\beta (t)\bigr)D_{\beta }W_{\beta }^{(j)}(t)= \left \vert \begin{matrix} y_{1}(\beta (t)) & \ldots & y_{k+1} (\beta (t)) \\ D_{\beta } y_{1} (\beta (t)) & \ldots & D_{\beta } y_{k+1} (\beta (t)) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{k-1} y_{1}(\beta (t)) & \ldots & D_{\beta }^{k-1} y_{k+1}( \beta (t)) \\ D_{\beta }^{k+1}y_{1} (t) & \ldots &D_{\beta }^{k+1} y_{k+1}(t) \end{matrix} \right \vert . $$

Thus, we have

$$ D_{\beta }W_{\beta }(y_{1},\ldots,y_{k+1}) (t)= \left \vert \begin{matrix} y_{1}(\beta (t)) & \ldots & y_{k+1} (\beta (t)) \\ D_{\beta }y_{1} (\beta (t)) & \ldots & D_{\beta }y_{k+1} (\beta (t)) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{k-1} y_{1}(\beta (t)) & \ldots & D_{\beta }^{k-1} y_{k+1}( \beta (t)) \\ D_{\beta }^{k+1}y_{1} (t) & \ldots &D_{\beta }^{k+1} y_{k+1}(t) \end{matrix} \right \vert $$

as required. □

Theorem 3.18

If \(y_{1}(t),\ldots,y_{n}(t)\) are solutions of equation (3.2) in J, then their β-Wronskian satisfies the first-order β-difference equation

$$ D_{\beta }W_{\beta }(t)=-P(t)W_{\beta }(t),\quad \forall t\in J \backslash \{s_{0}\}, $$
(3.20)

where

$$ P(t)= \sum_{k=0} ^{n-1} \bigl(t-\beta (t) \bigr)^{k}a_{k+1}(t)/a_{0}(t). $$

Proof

First, we show by induction that the following relation

$$ D_{\beta }W_{\beta }(y_{1},\ldots,y_{n})=\sum _{k=1}^{n}(-1)^{k-1}\bigl(t- \beta (t)\bigr)^{k-1} \left \vert \begin{matrix} y_{1}(t)& \ldots & y_{n}(t) \\ D_{\beta }y_{1}(t) & \ldots & D_{\beta }y_{n}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{n-k-1}y_{1}(t) & \ldots & D_{\beta }^{n-k-1}y_{n}(t) \\ D_{\beta }^{n-k+1}y_{1}(t) & \ldots & D_{\beta }^{n-k+1}y_{n}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{n}y_{1}(t) & \ldots & D_{\beta }^{n}y_{n}(t) \end{matrix} \right \vert $$
(3.21)

holds. Indeed, clearly (3.21) is true at \(n=1\). Assume that (3.21) is true for \(n=m\). From Lemma 3.17,

$$\begin{aligned} D_{\beta }W_{\beta }(y_{1},\ldots,y_{m+1}) (t) =& \left \vert \begin{matrix} y_{1}(\beta (t)) & \ldots & y_{m+1}(\beta (t)) \\ D_{\beta }y_{1}(\beta (t)) & \ldots & D_{\beta }y_{m+1}(\beta (t)) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{m-1}y_{1}(\beta (t)) & \ldots & D_{\beta }^{m-1}y_{m+1}( \beta (t)) \\ D_{\beta }^{m+1}y_{1}(t) & \ldots & D_{\beta }^{m+1}y_{m+1}(t) \end{matrix} \right \vert \\ =&\sum_{j=1}^{m+1}(-1)^{j+1}y_{j} \bigl(\beta (t)\bigr)W_{\beta }^{\ast (j)}(t), \end{aligned}$$

where

$$ W_{\beta }^{\ast (j)}= \textstyle\begin{cases} D_{\beta }W_{\beta }(D_{\beta }y_{2},\ldots,D_{\beta }y_{m+1}),& j=1, \\ D_{\beta }W_{\beta }(D_{\beta }y_{1},D_{\beta }y_{j-1},D_{\beta }y _{j+1},\ldots,D_{\beta }y_{m+1}),& 2\leq j\leq m, \\ D_{\beta }W _{\beta }(D_{\beta }y_{1},\ldots,D_{\beta }y_{m}),& j=m+1. \end{cases} $$

One can see that \(W_{\beta }^{\ast (j)}(t)=\sum_{k=1}^{m}(-1)^{k-1} (t-\beta (t) )^{k-1}R_{jk}\), where

$$\begin{aligned}& R_{jk}= \left \vert \begin{matrix} D_{\beta }y_{1}(t) & \ldots & D_{\beta }y_{j-1}(t) & D_{\beta }y_{j+1}(t) & \ldots & D_{\beta }y_{m+1}(t) \\ D_{\beta }^{2} y_{1}(t) & \ldots & D_{\beta }^{2} y_{j-1}(t) & D_{ \beta }^{2} y_{j+1}(t)& \ldots & D_{\beta }^{2} y_{m+1}(t) \\ \vdots & \ddots & \vdots & \vdots & \ddots & \vdots \\ D_{\beta }^{m-k} y_{1}(t) & \ldots & D_{\beta }^{m-k} y_{j-1}(t) & D _{\beta }^{m-k} y_{j+1}(t) & \ldots & D_{\beta }^{m-k} y_{m+1}(t) \\ D_{\beta }^{m-k+2} y_{1}(t) & \ldots & D_{\beta }^{m-k+2} y_{j-1}(t) & D_{\beta }^{m-k+2} y_{j+1}(t) & \ldots & D_{\beta }^{m-k+2} y_{m+1}(t) \\ \vdots & \ddots & \vdots & \vdots & \ddots & \vdots \\ D_{\beta }^{m+1} y_{1}(t) & \ldots & D_{\beta }^{m+1} y_{j-1}(t) & D _{\beta }^{m+1} y_{j+1}(t)& \ldots & D_{\beta }^{m+1} y_{m+1}(t) \end{matrix} \right \vert , \\& \quad 2\leq j\leq m, \\& R_{jk}= \left \vert \begin{matrix} D_{\beta }y_{2}(t) & \ldots & D_{\beta }y_{m+1}(t) \\ D_{\beta }^{2} y_{2}(t)& \ldots & D_{\beta }^{2} y_{m+1}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{m-k} y_{2}(t) & \ldots & D_{\beta }^{m-k} y_{m+1}(t) \\ D_{\beta }^{m-k+2} y_{2}(t) & \ldots & D_{\beta }^{m-k+2} y_{m+1}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{m+1} y_{2}(t) & \ldots & D_{\beta }^{m+1} y_{m+1}(t) \end{matrix} \right \vert ,\quad j=1, \\& R_{jk}= \left \vert \begin{matrix} D_{\beta }y_{1}(t) & \ldots & D_{\beta }y_{m}(t) \\ D_{\beta }^{2} y_{1}(t)& \ldots & D_{\beta }^{2} y_{m}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{m-k} y_{1}(t) & \ldots & D_{\beta }^{m-k} y_{m}(t) \\ D_{\beta }^{m-k+2} y_{1}(t) & \ldots & D_{\beta }^{m-k+2} y_{m}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{m+1} y_{1}(t) & \ldots & D_{\beta }^{m+1} y_{m}(t) \end{matrix} \right \vert ,\quad j=m+1. \end{aligned}$$

It follows that

$$\begin{aligned} D_{\beta } W_{\beta }(y_{1},\ldots,y_{m+1}) =& \sum_{j=1}^{m+1}(-1)^{j+1}\bigl[y _{j}(t)-\bigl(t-\beta (t)\bigr)D_{\beta }y_{j}(t) \bigr] \\ &{} \times \sum_{k=1}^{m}(-1)^{k-1} \bigl(t- \beta (t)\bigr)^{k-1}R_{jk} \\ =&\sum_{k=1}^{m}(-1)^{k-1} \bigl(t-\beta (t)\bigr)^{k-1}\sum_{j=1}^{m+1}(-1)^{j+1}y _{j}(t)R_{jk} \\ &{}+ \sum_{k=1}^{m}(-1)^{k} \bigl(t-\beta (t)\bigr)^{k}\sum_{j=1}^{m+1}(-1)^{j+1} D_{\beta }y_{j}(t)R_{jk} \\ =&\sum_{k=1}^{m}(-1)^{k-1} \bigl(t-\beta (t)\bigr)^{k-1}M(k)+ \sum_{k=1}^{m}(-1)^{k} \bigl(t-\beta (t)\bigr)^{k}L(k), \end{aligned}$$
(3.22)

where

$$\begin{aligned}& M(k)=\sum_{j=1}^{m+1}(-1)^{j+1}y_{j}(t)R_{jk}= \left \vert \begin{matrix} y_{1}(t) & \ldots & y_{m+1}(t) \\ D_{\beta }y_{1}(t) & \ldots & D_{\beta }y_{m+1}(t) \\ D_{\beta }^{2} y_{1}(t)& \ldots & D_{\beta }^{2} y_{m+1}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{m-k} y_{1}(t) & \ldots & D_{\beta }^{m-k} y_{m+1}(t) \\ D_{\beta }^{m-k+2} y_{1}(t) & \ldots & D_{\beta }^{m-k+2} y_{m+1}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{m+1} y_{1}(t) & \ldots & D_{\beta }^{m+1} y_{m+1}(t) \end{matrix} \right \vert , \end{aligned}$$
(3.23)
$$\begin{aligned}& L(k)=\sum_{j=1}^{m+1}(-1)^{j+1} D_{\beta }y_{j}(t)R_{jk}=\textstyle\begin{cases} 0,& \mbox{if } (k=1,\ldots,m-1), \\ \left \vert {\scriptsize\begin{matrix}{} D_{\beta }y_{1}(t) & \ldots & D_{\beta }y_{m+1}(t) \cr D_{\beta }^{2} y_{1}(t)& \ldots & D_{\beta }^{2} y_{m+1}(t) \cr \vdots & \ddots & \vdots \cr D_{\beta }^{m+1} y_{1}(t) & \ldots & D_{\beta }^{m+1} y_{m+1}(t) \end{matrix}} \right \vert ,& \mbox{if } k=m. \end{cases}\displaystyle \end{aligned}$$
(3.24)

Using relations (3.23) and (3.24) and substituting in (3.22), we obtain relation (3.21) at \(n = m+1\). Since \(D_{\beta }^{n}y_{j}(t)=-\sum_{i=1}^{n} (a_{i}(t)/a_{0}(t) )D _{\beta }^{n-i}y_{j}(t)\), it follows that

$$\begin{aligned} D_{\beta }W_{\beta }(t) =&\sum_{k=1}^{n}(-1)^{k-1} \bigl(t-\beta (t)\bigr)^{k-1} \biggl(\frac{-a_{k}(t)}{a_{0}(t)} \biggr) \left \vert \begin{matrix} y_{1}(t) & \ldots & y_{n}(t) \\ D_{\beta }y_{1}(t) & \ldots & D_{\beta }y_{n}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{n-k-1} y_{1}(t)& \ldots & D_{\beta }^{n-k-1} y_{n}(t) \\ D_{\beta }^{n-k+1} y_{1}(t) & \ldots & D_{\beta }^{n-k+1} y_{n}(t) \\ \vdots & \ddots & \vdots \\ D_{\beta }^{n-1} y_{1}(t) & \ldots & D_{\beta }^{n-1} y_{n}(t) \\ D_{\beta }^{n-k} y_{1}(t) & \ldots & D_{\beta }^{n-k} y_{n}(t) \end{matrix} \right \vert \\ =&\sum_{k=1}^{n}(-1)^{2(k-1)} \bigl(t-\beta (t) \bigr)^{k-1} \biggl(\frac{-a _{k}(t)}{a_{0}(t)} \biggr)W_{\beta }(t) \\ =&-\sum_{k=0}^{n-1} \bigl(t-\beta (t) \bigr)^{k} \biggl(\frac{a_{k+1}(t)}{a _{0}(t)} \biggr)W_{\beta }(t) =-P(t)W_{\beta }(t), \end{aligned}$$

which is the desired result. □

The following theorem gives us Liouville’s formula for β-difference equations.

Theorem 3.19

Assume that \((\beta (t)-t)P(t)\neq 1\), \(t\in J\). Then the β-Wronskian of any set of solutions \(\{y_{i}(t)\}_{i=1}^{n}\), valid in J, is given by

$$ W_{\beta }(t)=\frac{W_{\beta }(s_{0})}{\prod_{k=0}^{\infty } [1+ P( \beta^{k}(t)) (\beta^{k} (t)-\beta^{k+1}(t) ) ]}, \quad t\in J. $$
(3.25)

Proof

Relation (3.20) implies that

$$ W_{\beta }\bigl(\beta (t)\bigr)= \bigl[1+\bigl(t-\beta (t)\bigr)P(t) \bigr]W_{\beta }(t),\quad t \in J\backslash \{s_{0}\}. $$

Hence,

$$\begin{aligned} W_{\beta }(t) =&\frac{W_{\beta }(\beta (t))}{1+(t-\beta (t))P(t)} \\ =&\frac{W_{\beta }(\beta^{m}(t))}{\prod_{k=0}^{m-1} [1+ P(\beta ^{k}(t)) (\beta^{k} (t) - \beta^{k+1}(t) ) ]},\quad m \in \mathbb{N}. \end{aligned}$$

Taking \(m\rightarrow \infty \), we get

$$ W_{\beta }(t)=\frac{W_{\beta }(s_{0})}{\prod_{k=0}^{\infty } [1+ P( \beta^{k}(t)) (\beta^{k} (t) - \beta^{k+1}(t) ) ]},\quad t\in J. $$

 □

Example 3.20

We calculate the β-Wronskian of the β-difference equation

$$ D_{\beta }^{2}y(t)+y(t)=0. $$
(3.26)

The functions \(y_{1}(t)=\cos_{1,\beta }(t)\) and \(y_{2}(t)=\sin_{1, \beta }(t)\) are solutions of equation (3.26) subject to the initial conditions \(y_{1}(s_{0})=1\), \(D_{\beta }y_{1}(s_{0})=0\), \(y _{2}(s_{0})=0\), \(D_{\beta }y_{2}(s_{0})=1\), respectively. Here, \(P(t)=(t-\beta (t))\). So, \((\beta (t)- t)P(t)\neq 1\) for all \(t\neq s_{0} \). Since

$$ W_{\beta }(s_{0})= \left \vert \begin{matrix} \cos_{1,\beta }(s_{0}) & \sin_{1,\beta }(s_{0}) \\ \sin_{1,\beta }(s_{0})& \cos_{1,\beta }(s_{0}) \end{matrix} \right \vert =\left \vert \begin{matrix} 1 & 0 \\ 0 & 1 \end{matrix} \right \vert =1. $$

Therefore, \(W_{\beta }(t)=\frac{1}{\prod_{k=0}^{\infty } [1+ (\beta^{k}(t)-\beta ^{k+1}(t) )^{2} ]}\).

The following corollary can be deduced directly from Theorem 3.19.

Corollary 3.21

Let \(\{y_{i}\}_{i=1}^{n}\) be a set of solutions of equation (3.2) in J. Then \(W_{\beta }(t)\) has two possibilities:

  1. (i)

    \(W_{\beta }(t)\neq 0\) in J if and only if \(\{y_{i} \} _{i=1}^{n}\) is a fundamental set of equation (3.2) valid in J.

  2. (ii)

    \(W_{\beta }(t)=0\) in J if and only if \(\{y_{i}\}_{i=1} ^{n}\) is not a fundamental set of equation (3.2) valid in J.

3.4 Non-homogeneous linear β-difference equations

The nth-order non-homogeneous linear β-difference equation has the form

$$ a_{0}(t)D_{\beta }^{n}y(t)+a_{1}(t)D_{\beta }^{n-1}y(t)+ \cdots +a_{n-1}(t)D _{\beta }y(t)+a_{n}(t)y(t)=b(t), $$
(3.27)

where the coefficients \(a_{j}(t)\), \(0\leq j\leq n\), and \(b(t)\) are assumed to satisfy the conditions of Corollary 3.3. We may write this as

$$ L_{n}y=b(t), $$
(3.28)

where, as before, \(L_{n}=a_{0}(t)D_{\beta }^{n}+a_{1}(t)D_{\beta } ^{n-1}+\cdots +a_{n-1}(t)D_{\beta }+a_{n}(t)\).

By the theory of differential equations, if \(y_{1}(t)\) and \(y_{2}(t)\) are two solutions of the non-homogeneous equation (3.28), then \(y_{1}\pm y_{2}\) is a solution of the corresponding homogeneous equation (3.2). Also, by Theorem 3.11, if \(y_{1}(t),\ldots,y _{n}(t)\) form a fundamental set for equation (3.2) and \(\varphi (t)\) is a particular solution of equation (3.27), then for any solution of equation (3.27), there are constants \(c_{1},\ldots,c_{n}\) such that

$$ y (t)=\varphi (t)+c_{1}y_{1}(t)+\cdots +c_{n}y_{n}(t). $$
(3.29)

Therefore, if we can find any particular solution \(\varphi (t)\) of equation (3.27), then (3.29) gives a general formula for all solutions of equation (3.27).

Theorem 3.22

Let \(\varphi_{i}\) be a particular solution of \(L_{n}y=b_{i}(t)\), \(i=1, \ldots,m\). Then \(\sum_{i=1}^{m}\zeta_{i}\varphi_{i}\) is a particular solution of the equation \(L_{n}y=\sum_{i=1}^{m}\zeta_{i}b_{i}(t)\), where \(\zeta_{1},\ldots,\zeta_{m}\) are constants.

Proof

The proof is straightforward. □

3.4.1 Method of undetermined coefficients

We will illustrate the method of undetermined coefficients when the coefficients \(a_{j}\) (\(0\leq j \leq n \)) of the non-homogeneous linear β-difference equation (3.27) are constants by simple examples.

Example 3.23

Find a particular solution of

$$ D_{\beta }^{2}y(t)-3D_{\beta }y(t)-4y(t)=3e_{2,\beta }(t). $$
(3.30)

Assume that

$$ \varphi (t)=\zeta e_{2,\beta }(t), $$
(3.31)

where the coefficient ζ is a constant to be determined. To find ζ, we calculate

$$ D_{\beta }\varphi (t)=2\zeta e_{2,\beta }(t),\qquad D_{\beta }^{2}\varphi (t)=4\zeta e_{2,\beta }(t) $$
(3.32)

by substituting with equations (3.31), (3.32) in equation (3.30). Thus a particular solution is

$$ \varphi (t)=-1/2e_{2,\beta }(t). $$

In the following example, we refer the reader to see the different cases of the roots of the characteristic equation of second-order linear homogeneous β-difference equation when the coefficients are constants, see [6].

Example 3.24

Find the general solution of

$$ D_{\beta }^{2}y-3D_{\beta }y-4y=2\sin_{1,\beta }(t). $$
(3.33)

The corresponding homogeneous equation of (3.33) is

$$ D_{\beta }^{2}y-3D_{\beta }y-4y=0. $$
(3.34)

Then the characteristic polynomial of (3.34) is

$$ P(\lambda)=\lambda^{2}-3\lambda -4=0. $$
(3.35)

Therefore,

$$ y_{h}(t)=c_{1}e_{4,\beta }(t)+c_{2}e_{-1,\beta }(t). $$

Now, assume that

$$ \varphi (t)=\zeta_{1}\sin_{1,\beta }(t)+\zeta_{2} \cos_{1,\beta }(t), $$
(3.36)

where \(\zeta_{1}\) and \(\zeta_{2}\) are to be determined. Then

$$\begin{aligned}& \begin{aligned}D_{\beta }\varphi (t) =\zeta_{1} \cos_{1,\beta }(t)-\zeta_{2} \sin_{1,\beta }(t), \\ D_{\beta }^{2}\varphi (t) =-\zeta_{1} \sin_{1,\beta }(t)-\zeta_{2} \cos_{1,\beta }(t). \end{aligned} \end{aligned}$$
(3.37)

By substituting with equations (3.36), (3.37) in equation (3.33), we get a particular solution

$$ \varphi (t)=-5/17\sin_{1,\beta }(t)+3/17\cos_{1,\beta }(t). $$

Then the general solution of (3.33) is

$$ y(t)=c_{1}e_{4,\beta }(t)+c_{2}e_{-1,\beta }(t)-5/17 \sin_{1,\beta }(t)+3/17\cos _{1,\beta }(t). $$

In the following example, we show the solution in the case of \(b(t)\) being a linear combination of exponential and trigonometric functions.

Example 3.25

Find the general solution of

$$ D_{\beta }^{2}y-2D_{\beta }y-3y=2e_{1,\beta }(t)-10 \sin_{1,\beta }(t). $$
(3.38)

The corresponding homogeneous equation of (3.38) has the solution

$$ {y}_{h}(t)=c_{1}e_{3,\beta }(t)+c_{2}e_{-1,\beta }(t). $$

The non-homogeneous term is the linear combination \(2e_{1,\beta }(t)-10 \sin_{1,\beta }(t)\) of the two functions given by \(e_{1,\beta }(t)\) and \(\sin_{1,\beta }(t)\).

Let

$$\begin{aligned} \varphi (t)=c_{1}e_{1,\beta }(t)+ c_{2} \sin_{1,\beta }(t)+c_{3} \cos_{1,\beta }(t) \end{aligned}$$
(3.39)

be a particular solution of (3.38). Then

$$\begin{aligned}& \begin{aligned}D_{\beta }\varphi (t) =c_{1}e_{1,\beta }(t)+c_{2} \cos_{1,\beta }(t)-c _{3}\sin_{1,\beta }(t), \\ D_{\beta }^{2}\varphi (t) =c_{1}e_{1,\beta }(t)-c_{2} \sin_{1,\beta }(t)-c _{3}\cos_{1,\beta }(t), \end{aligned} \end{aligned}$$
(3.40)

where \(c_{1}\), \(c_{2}\), \(c_{3}\) are undetermined coefficients. By substituting with (3.39), (3.40) in (3.38), we have the particular solution \(\varphi (t)=-1/2e_{1,\beta }(t)+2\sin_{1,\beta }(t)- \cos_{1,\beta }(t)\). Thus the general solution of (3.38) is

$$ y(t)= c_{1}e_{3,\beta }(t)+c_{2}e_{-1,\beta }(t)-1/2e_{1,\beta }(t)+2 \sin_{1,\beta }(t)-\cos_{1,\beta }(t). $$

Example 3.26

Find the general solution of

$$ D_{\beta }^{2}y-3D_{\beta }y+2y=e_{3,\beta }(t) \sin_{4,\beta }(t). $$
(3.41)

The corresponding homogeneous equation of (3.41) has the solution

$$ y_{h}(t)=c_{1}e_{2,\beta }(t)+c_{2}e_{1,\beta }(t). $$

Let

$$\begin{aligned} \varphi (t)=Ae_{3,\beta }(t)\sin_{4,\beta }(t)+ Be_{3,\beta }(t) \cos_{4,\beta }(t) \end{aligned}$$
(3.42)

be a particular solution of (3.41), where A and B are constants. Then

$$\begin{aligned}& D_{\beta }\varphi (t) =3Ae_{3,\beta }(t)\sin_{4,\beta }(t)+4A e_{3, \beta }\bigl(\beta (t)\bigr)\cos_{4,\beta }(t) \\& \hphantom{D_{\beta }\varphi (t) =}{}-3Be_{3,\beta }(t)\cos_{4,\beta }(t)-4Be_{3,\beta } \bigl(\beta (t)\bigr) \sin_{4,\beta }(t), \end{aligned}$$
(3.43)
$$\begin{aligned}& D_{\beta }^{2}\varphi (t) =9Ae_{3,\beta }(t) \sin_{4,\beta }(t)+12A e _{3,\beta }\bigl(\beta (t)\bigr) \cos_{4,\beta }(t) \\& \hphantom{D_{\beta }^{2}\varphi (t) =}{}+12A e_{3,\beta }\bigl(\beta (t)\bigr)\cos_{4,\beta } \bigl(\beta (t)\bigr)-16A e_{3, \beta }\bigl(\beta (t)\bigr) \sin_{4,\beta }(t) \\& \hphantom{D_{\beta }^{2}\varphi (t) =}{} +9Be_{3,\beta }(t)\cos_{4,\beta }(t)-12Be_{3,\beta } \bigl(\beta (t)\bigr) \sin_{4,\beta }(t) \\& \hphantom{D_{\beta }^{2}\varphi (t) =}{} -12Be_{3,\beta }\bigl(\beta (t)\bigr)\sin_{4,\beta } \bigl(\beta (t)\bigr)-16Be_{3, \beta }\bigl(\beta (t)\bigr) \cos_{4,\beta }(t). \end{aligned}$$
(3.44)

By substituting with (3.42), (3.43) and (3.44) in (3.41), we get \(A=\frac{1}{2}\) and \(B=0\). Then the particular solution is \(\varphi (t)=1/2e_{3,\beta }(t) \sin_{4,\beta }(t)\). Thus the general solution of (3.41) is

$$ y(t)= c_{1}e_{2,\beta }(t)+c_{2}e_{1,\beta }(t)+1/2e_{3,\beta }(t) \sin_{4,\beta }(t). $$

3.4.2 Method of variation of parameters

We use the method of variation of parameters to obtain a particular solution \(\varphi (t)\) of the non-homogeneous linear β-difference equation (3.27), which can be applied in the case of the coefficients \(a_{j}\) (\(0\leq j \leq n \)) being functions or constants. It depends on replacing the constants \(c_{r}\) in relation (3.29) by the functions \(\zeta_{r}(t)\). Hence, we try to find a solution of the form

$$ \varphi (t)=\zeta_{1}(t)y_{1} (t)+\cdots + \zeta_{n}(t)y_{n} (t). $$
(3.45)

Our objective is to determine the functions \(\zeta_{r} (t)\). We have

$$ D_{\beta }^{i-1}\varphi (t)=\sum_{j=1}^{n} \zeta_{j}(t)D_{\beta }^{i-1}y _{j}(t), \quad 1\leq i\leq n, $$
(3.46)

provided that

$$ \sum_{j=1}^{n}D_{\beta } \zeta_{j}(t)D_{\beta }^{i-1}y_{j}\bigl(\beta (t)\bigr)=0,\quad 1\leq i\leq n-1. $$
(3.47)

Putting \(i = n\) in (3.46) and operating on it by \(D_{\beta }\), we obtain

$$ D_{\beta }^{n}\varphi (t)= \sum_{j=1}^{n} \zeta_{j}(t)D_{\beta }^{n}y _{j}(t)+D_{\beta } \zeta_{j}(t)D_{\beta }^{n-1}y_{j}\bigl(\beta (t)\bigr). $$
(3.48)

Since \(\varphi (t)\) satisfies equation (3.27), it follows that

$$ a_{0}(t) D_{\beta }^{n}\varphi (t)+a_{1}(t)D_{\beta }^{n-1} \varphi (t)+ \cdots +a_{n}(t)\varphi (t)=b(t). $$
(3.49)

Substitute by (3.46) and (3.48) in (3.49) and in view of equation (3.2), we obtain

$$ \sum_{j=1}^{n}D_{\beta } \zeta_{j}(t)D_{\beta }^{n-1}y_{j}\bigl(\beta (t)\bigr)=\frac{b(t)}{a _{0}(t)}. $$

Thus, we get the following system:

$$\begin{aligned}& \begin{aligned} &D_{\beta }\zeta_{1}(t)y_{1}\bigl( \beta (t)\bigr)+\cdots + D_{\beta }\zeta_{n}(t)y _{n} \bigl(\beta (t)\bigr)=0, \\ &\vdots \\ &D_{\beta }\zeta_{1}(t)D_{\beta }^{n-2}y_{1} \bigl(\beta (t)\bigr)+\cdots +D_{ \beta }\zeta_{n}(t)D_{\beta }^{n-2}y_{n} \bigl(\beta (t)\bigr)=0, \\ &D_{\beta }\zeta_{1}(t)D_{\beta }^{n-1}y_{1} \bigl(\beta (t)\bigr)+\cdots +D_{ \beta }\zeta_{n}(t)D_{\beta }^{n-1}y_{n} \bigl(\beta (t)\bigr)= \frac{b(t)}{a_{0}(t)}. \end{aligned} \end{aligned}$$
(3.50)

Consequently,

$$ D_{\beta }\zeta _{r}(t) =\frac{W_{r}(\beta (t))}{W_{\beta }(\beta (t))}\times \frac{b(t)}{a _{0}(t)}, \quad t \in I, $$

where \(1\leq r\leq n\) and \(W_{r}(\beta (t))\) is the determinant obtained from \(W_{\beta }(\beta (t))\) by replacing the rth column by \((0,\ldots,0,1)\). It follows that

$$ \zeta_{r}(t)= \int_{s_{0}}^{t}\frac{W_{r}(\beta (\tau))}{W_{\beta }( \beta (\tau))}\times \frac{b(\tau)}{a_{0}(\tau)}\,d_{\beta }\tau,\quad r=1,\ldots,n. $$

Example 3.27

Consider the equation

$$ D_{\beta }^{2}y(t)+z^{2}y(t)=b(t), $$
(3.51)

where \(z\in \mathbb{C}\setminus \{0\}\). It is known that \(\cos_{z, \beta } (t)\) and \(\sin_{z,\beta }(t)\) are the solutions of the corresponding homogeneous equation of (3.51). We can easily show that

$$\begin{aligned} \varphi (t)=\frac{1}{z} \biggl[\sin_{z,\beta }(t) \int_{s_{0}}^{t}b( \tau)\operatorname{Cos}_{z,\beta } \bigl( \beta (\tau)\bigr)\,d_{\beta }\tau -\cos_{z,\beta } (t) \int_{s_{0}}^{t}b(\tau)\operatorname{Sin}_{z,\beta }\bigl(\beta (\tau)\bigr)\,d_{\beta }\tau \biggr]. \end{aligned}$$

It follows that every solution of equation (3.51) has the form

$$\begin{aligned} y(t) =&c_{1}\cos_{z,\beta }(t)+c_{2} \sin_{z,\beta }(t) \\ &{}+\frac{1}{z} \biggl[\sin_{z,\beta }(t) \int_{s_{0}}^{t}b(\tau)\operatorname{Cos}_{z, \beta } \bigl(\beta (\tau)\bigr)\,d_{\beta }\tau -\cos_{z,\beta } (t) \int_{s_{0}} ^{t}b(\tau)\operatorname{Sin}_{z,\beta }\bigl(\beta (\tau)\bigr)\,d_{\beta }\tau \biggr]. \end{aligned}$$

3.4.3 Annihilator method

In this section, we can use annihilator method to obtain the particular solution of non-homogeneous linear β-difference equation (3.27) when the coefficients \(a_{j}\) (\(0\leq j \leq n \)) are constants.

Definition 3.28

We say that \(f:I\rightarrow \mathbb{C}\) can be annihilated provided that we can find an operator of the form

$$ L(D)=\rho_{n}D_{\beta }^{n}+\rho_{n-1}D_{\beta }^{n-1}+ \cdots +\rho _{0}\mathcal{I} $$

such that \(L(D)f(t)=0\), \(t\in I\), where \(\rho_{i}\), \(0 \leq i\leq n\) are constants, not all zero.

Example 3.29

Since \((D_{\beta }-4\mathcal{I})e_{4,\beta }(t)=0\), \(D_{\beta }-4\mathcal{I}\) is an annihilator for \(e_{4,\beta }(t)\).

Table 1 indicates a list of some functions and their annihilators.

Table 1 A list of some functions and their annihilators

Example 3.30

Consider the equation

$$ D_{\beta }^{2}y(t)-4D_{\beta }y(t)+3y(t)=e_{5,\beta }(t). $$
(3.52)

Equation (3.52) can be rewritten in the form

$$ (D_{\beta }-3\mathcal{I}) (D_{\beta }-\mathcal{I})y(t)= e_{5,\beta }(t). $$

Multiplying both sides by the annihilator \((D_{\beta }-5\mathcal{I})\), we get that if \(y(t)\) is a solution of (3.52), then \(y(t)\) satisfies

$$ (D_{\beta }-3\mathcal{I}) (D_{\beta }-\mathcal{I}) (D_{\beta }-5 \mathcal{I})y(t)=0. $$

Hence,

$$ y(t)=c_{1}e_{3,\beta }(t)+c_{2}e_{1,\beta }(t)+c_{3}e_{5,\beta }(t). $$

One can see that \(\varphi (t)=(1/8)e_{5,\beta }(t)\) is a solution of equation (3.52). Therefore, the general solution of equation (3.52) has the following form:

$$ y(t)=c_{1}e_{3,\beta }(t)+c_{2}e_{1,\beta }(t)+(1/8)e_{5,\beta }(t). $$

4 Conclusion

In this paper, the sufficient conditions for the existence and uniqueness of solutions of the β-Cauchy problem were given. Also, a fundamental set of solutions for the homogeneous linear β-difference equations when the coefficients \(a_{j}\) (\(0\leq j \leq n\)) are constants was constructed. Moreover, β-Wronskian and its properties were introduced. Finally, the undetermined coefficients, the variation of parameters, and the annihilator methods for the non-homogeneous case were presented.

References

  1. Annaby, M.H., Hamza, A.E., Aldowah, K.A.: Hahn difference operator and associated Jackson–Nörlund integrals. J. Optim. Theory Appl. 154, 133–153 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  2. Annaby, M.H., Mansour, Z.S.: q-Fractional Calculus and Equations. Springer, Berlin (2012)

    Book  MATH  Google Scholar 

  3. Askey, R., Wilson, J.: Some basic hypergeometric orthogonal polynomials that generalize the Jacobi polynomials. Mem. Am. Math. Soc. 54, 1–55 (1985)

    MathSciNet  MATH  Google Scholar 

  4. Bangerezako, G.: An Introduction to q-Difference Equations. Bujumbura (2008)

  5. Cresson, J., Frederico, G., Torres, D.F.M.: Constants of motion for non-differentiable quantum variational problems. Topol. Methods Nonlinear Anal. 33, 217–231 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  6. Faried, N., Shehata, E.M., El Zafarani, R.M.: On homogeneous second order linear general quantum difference equations. J. Inequal. Appl. 2017, 198 (2017). https://doi.org/10.1186/s13660-017-1471-3

    Article  MathSciNet  MATH  Google Scholar 

  7. Gasper, G., Rahman, M.: Basic Hypergeometric Series. Cambridge University Press, Cambridge (1990)

    MATH  Google Scholar 

  8. Hamza, A.E., Ahmed, S.M.: Theory of linear Hahn difference equations. J. Adv. Math. 4(2), 441–461 (2013)

    Google Scholar 

  9. Hamza, A.E., Sarhan, A.M., Shehata, E.M.: Exponential, trigonometric and hyperbolic functions associated with a general quantum difference operator. Adv. Dyn. Syst. Appl. 12(1), 25–38 (2017)

    MathSciNet  Google Scholar 

  10. Hamza, A.E., Sarhan, A.M., Shehata, E.M., Aldowah, K.A.: A general quantum difference calculus. Adv. Differ. Equ. 2015, 182 (2015). https://doi.org/10.1186/s13660-015-0518-3

    Article  MathSciNet  Google Scholar 

  11. Hamza, A.E., Shehata, E.M.: Existence and uniqueness of solutions of general quantum difference equations. Adv. Dyn. Syst. Appl. 11, 45–58 (2016)

    MathSciNet  Google Scholar 

  12. Kac, V., Cheung, P.: Quantum Calculus. Springer, New York (2002)

    Book  MATH  Google Scholar 

  13. Malinowska, A.B., Torres, D.F.M.: Quantum Variational Calculus. Briefs in Electrical and Computer Engineering—Control, Automation and Robotics. Springer, Berlin (2014)

    Book  Google Scholar 

  14. Nottale, L.: Fractal Space-Time and Microphysics: Towards a Theory of Scale Relativity. World Scientific, Singapore (1993)

    Book  MATH  Google Scholar 

Download references

Acknowledgements

The authors sincerely thank the referees for their valuable suggestions and comments.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed equally and significantly in writing this article. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Enas M. Shehata.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Faried, N., Shehata, E.M. & El Zafarani, R.M. Theory of nth-order linear general quantum difference equations. Adv Differ Equ 2018, 264 (2018). https://doi.org/10.1186/s13662-018-1715-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13662-018-1715-7

MSC

Keywords