1 Introduction and preliminaries

In 1912, Bernstein [1] introduced a sequence of operators \(B_{n}:C[0,1]\rightarrow C[0,1]\) defined by

$$ B_{n}(f,x)=\sum_{k=0}^{n} \begin{pmatrix} n \cr k \end{pmatrix} x^{k}(1-x)^{n-k}f \biggl(\frac{k}{n} \biggr),\quad x\in[0,1], $$
(1.1)

for \(n\in\mathbb{N}\) and \(f\in C[0,1]\). Szász in 1950 (see [4]) and Mirakjan in 1941 (see [5]) generalized the Bernstein polynomial to an infinite interval as

$$ S_{n}(f,x)=e^{-nx}\sum _{k=0}^{\infty}\frac{(nx)^{k}}{k!}f \biggl( \frac{k}{n} \biggr),\quad f\in C[0,\infty). $$
(1.2)

In [6], the Kantorovich type of the Szász-Mirakjan operators was defined as

$$ K_{n}(f,x)=ne^{-nx}\sum _{k=0}^{\infty}\frac{(nx)^{k}}{k!} \int_{\frac{k}{n}}^{\frac{k+1}{n}}f(t)\,dt, $$
(1.3)

where f is a continuous nondecreasing function on \([0,\infty)\).

In the field of approximation theory, q-calculus plays an important role. In 1987, the first q-analog of the well-known Bernstein polynomial was introduced by Lupas [7]. Later on several researchers have proposed the q-analog of various operators and investigated their approximation properties [8, 9, 11, 12]. We recall some definitions of q-calculus (see [13]). Let \(k\in \mathbb{N}_{0}\) and \(q\in(0,1)\) then the q-integer \([k]_{q}\) is defined as

$$ [k]_{q} = \textstyle\begin{cases} \frac{1-q^{k}}{1-q} & \mbox{if }q\neq1, \\ k & \mbox{if }q=1.\end{cases} $$

The q-factorial \([k]_{q}!\) is defined as

$$ [k]_{q}! = \textstyle\begin{cases} [k]_{q}[k-1]_{q}\cdots[1]_{q} & \mbox{if }k\in\mathbb{N}, \\ 1 & \mbox{if }k=0,\end{cases} $$

and for \(k\in\mathbb{N}\), q-binomial coefficient \(\bigl[{\scriptsize\begin{matrix}{}k \cr r\end{matrix}} \bigr] _{q}\) is defined by

$$ \left [ \textstyle\begin{array}{c} k \\ r\end{array}\displaystyle \right ] _{q}=\frac{[k]_{q}!}{[r]_{q}![k-r]_{q}!},\quad 1 \leq r\leq k, $$

with \(\bigl[{\scriptsize\begin{matrix}{}k \cr 0\end{matrix}} \bigr] _{q}=1\) and \(\bigl[{\scriptsize\begin{matrix}{}k \cr r\end{matrix}} \bigr] _{q}=0\) for \(r>k\).

There are two q-analogs of the exponential function \(e^{x}\) (see [14]):

$$ e_{q}(x)=\sum_{k=0}^{\infty} \frac{x^{k}}{[k]_{q}!}=\frac{1}{(1-(1-q)x)_{q}^{\infty}},\quad \vert x\vert < \frac{1}{1-q}, \vert q\vert < 1, $$

and

$$ E_{q}(x)=\sum_{k=0}^{\infty}q^{\frac{k(k-1)}{2}} \frac{x^{k}}{[k]_{q}!}= \bigl(1+(1-q)x \bigr)_{q}^{\infty}, \quad \vert q\vert < 1, $$

where

$$ (1-x)_{q}^{\infty}=\prod_{j=0}^{\infty} \bigl(1-q^{j}x \bigr). $$

Now assume that \(0< a< b\), \(0< q<1\), and f is a real valued function. The q-Jackson integrals of f over the interval \([0,b]\) and over a general interval \([a,b]\) are defined by (see [15])

$$ \int_{0}^{b}f(t)\,d_{q}t=(1-q)b\sum _{j=0}^{\infty}f \bigl(bq^{j} \bigr)q^{j} $$

and

$$ \int_{a}^{b}f(t)\,d_{q}t= \int_{0}^{b}f(t)\,d_{q}t- \int_{0}^{a}f(t)\,d_{q}t, $$

respectively, provided the series converges. It is obvious that q-calculus reduces to the ordinary version when \(q=1\). A generalization of the exponential function is given by Sucu. Sucu [16] defined a Dunkl analog of the Szász operator given by

$$ S_{n}(f;x)=\frac{1}{e_{\mu}(nx)}\sum _{k=0}^{\infty}\frac{(nx)^{k}}{\gamma_{\mu}(k)}f \biggl( \frac{k+2\mu\theta_{k}}{n} \biggr), $$
(1.4)

where \(\mu\geq0\), \(n\in\mathbb{N}\), \(x\geq0\), \(f\in C[0,\infty)\), and

$$ e_{\mu}(x)=\sum_{k=0}^{\infty} \frac{x^{k}}{\gamma_{\mu}(k)}. $$

Here

$$ \gamma_{\mu}(2k)=\frac{2^{2k}k!\Gamma(k+\mu+\frac{1}{2})}{\Gamma(\mu+\frac{1}{2})} $$

and

$$ \gamma_{\mu}(2k+1)=\frac{2^{2k+1}k!\Gamma(k+\mu+\frac{3}{2})}{\Gamma(\mu+\frac{1}{2})}. $$

The recursion relation for \(\gamma_{\mu}\) is given by

$$ \gamma_{\mu}(k+1)=(k+1+2\mu\theta_{k+1})\gamma_{\mu}(k), \quad k \in\mathbb{N}_{0}, $$

where

$$ \theta _{k}=\textstyle\begin{cases} 0 & \text{if }k\in 2\mathbb{N}, \\ 1 & \text{if }k\in 2\mathbb{N}+1.\end{cases}$$

Cheikh et al. [17] stated the Dunkl analog of classical q-Hermite polynomials and gave definitions of the q-Dunkl analog of exponential functions, an explicit formula, and recursion relations for \(\mu>-\frac{1}{2}\) and \(0< q<1\), respectively:

$$ e_{\mu,q}(x)=\sum_{k=0}^{\infty} \frac{x^{k}}{\gamma_{\mu,q}(k)},\quad x\in [ 0,\infty ), $$
(1.5)

and

$$ E_{\mu,q}(x)=\sum_{k=0}^{\infty}q^{\frac{k(k-1)}{2}} \frac{x^{k}}{\gamma_{\mu,q}(k)},\quad x\in [ 0,\infty ). $$
(1.6)

An explicit formula for \(\gamma_{\mu,q}(k)\) is given by

$$ \gamma_{\mu,q}(k)=\frac{(q^{2\mu+1},q^{2})_{[\frac{k+1}{2}]}(q^{2},q^{2})_{[\frac{k}{2}]}}{(1-q)^{k}},\quad k\in \mathbb{N}_{0}, $$
(1.7)

where

$$ (x,q)_{k}= \textstyle\begin{cases} \prod_{n=0}^{k-1} (1-q^{n}x ) &\mbox{if } k=1,2,\ldots, \\ 1 & \mbox{if } k=0. \end{cases} $$

Some of the special cases of \(\gamma_{\mu,q}(k)\) are defined as

$$\begin{aligned}& \gamma_{\mu,q}(0)=1, \qquad \gamma_{\mu,q}(1)=\frac{1-q^{2\mu+1}}{1-q}, \\& \gamma_{\mu,q}(2)= \biggl(\frac{1-q^{2\mu+1}}{1-q} \biggr) \biggl( \frac{1-q^{2}}{1-q} \biggr), \\& \gamma_{\mu,q}(3)= \biggl(\frac{1-q^{2\mu+1}}{1-q} \biggr) \biggl( \frac{1-q^{2}}{1-q} \biggr) \biggl(\frac{1-q^{2\mu+3}}{1-q} \biggr), \\& \gamma_{\mu,q}(4)= \biggl(\frac{1-q^{2\mu+1}}{1-q} \biggr) \biggl( \frac{1-q^{2}}{1-q} \biggr) \biggl(\frac{1-q^{2\mu+3}}{1-q} \biggr) \biggl( \frac{1-q^{4}}{1-q} \biggr). \end{aligned}$$

The recursion relation for \(\gamma_{\mu,q}\) is given by

$$ \gamma_{\mu,q}(k+1)=[k+1+2\mu\theta_{k+1}]_{q} \gamma_{\mu,q}(k),\quad k\in \mathbb{N}_{0}, $$
(1.8)

where

$$ \theta _{k}=\textstyle\begin{cases} 0 & \text{if }k\in 2\mathbb{N}, \\ 1 & \text{if }k\in 2\mathbb{N}+1.\end{cases}$$

It has been observed that a sequence of linear positive operator preserve constant as well as linear functions, i.e., \(L_{n}(e_{i},x)=e_{i}(x)\) for \(e_{i}(x)=x^{i}(i=0,1)\). These conditions hold good for Bernstein polynomials, Szász-Mirakjan operators, Baskakov operators, Phillips operators, and so on. For each of the above operators \(L_{n}(e_{2},x)\neq e_{2}(x)\). In order to preserve \(e_{0}\) and \(e_{2}\), King [18] gave the modification of the well-known Bernstein polynomials as

$$ V_{n}(f,x)=\sum_{k=0}^{n} \binom{n}{k} \bigl(r_{n}^{*}(x) \bigr)^{k} \bigl(1-r_{n}^{*}(x) \bigr)^{n-k}f \biggl(\frac{k}{n} \biggr), $$
(1.9)

with \(0\leq r_{n}^{*}(x)\leq1\), \(n=1,2,\ldots,0\leq x\leq1\), where \(r_{n}^{*}(x)\) is given by

$$ r_{n}^{*}(x)= \textstyle\begin{cases} -\frac{1}{2(n-1)}+\sqrt{(\frac{n}{n-1})x^{2}+\frac{1}{4(n-1)^{2}}} & \mbox{if }n=2,3,\ldots, \\ x^{2} & \mbox{if }n=1. \end{cases} $$
(1.10)

Obviously, \(\lim_{n\rightarrow\infty}r_{n}^{*}(x)=x\). Also,

$$ V_{n}(e_{0},x)=1,\qquad V_{n}(e_{1},x)=r_{n}^{*}(x), \qquad V_{n}(e_{2},x)=x^{2}. $$

The Kantorovich variant of Szász operators preserves only a constant function. Approximation results on modified Szász-Mirakjan-Kantorovich operators preserving \(e_{0}\) and \(e_{1}\) have been investigated in [19] and q-Szász-Mirakjan-Kantorovich-type operators preserving test functions \(e_{1}\) and \(e_{2}\) have been studied in [20]. Previous studies demonstrated that providing a better error estimation for positive linear operators plays an important role in approximation theory, which allows us to approximate much faster to the function being approximated. In [19, 21, 22], various better approximation properties of the Szász-Mirakjan-Kantrovich operators, Szász-Mirakjan operators, and Szász-Mirakjan-Beta operators were investigated.

The purpose of this paper is to construct and investigate the Dunkl analog of q-Szász-Mirakjan-Kantorovich operators which preserves the test functions \(e_{1}\) and \(e_{2}\). Also, we have show that our modified operators have a better error estimation than the classical ones.

2 Operators and estimation of moments

Içöz gave a Dunkl generalization of Szász-Mirakjan-Kantorovich operators in [23] and one gave a Dunkl generalization of Szász operators via q-calculus in [24]. For \(\mu >\frac{1}{2}\), \(0< q<1\), and \(f\in C[0,\infty )\), we define a q-Dunkl analog of Szász-Mirakjan operators as

$$ D_{n,q}(f;x)=\frac{1}{e_{\mu,q}([n]_{q}\frac{x}{q})}\sum _{k=0}^{\infty }\frac{([n]_{q}x)^{k}}{\gamma _{\mu,q}(k)q^{k}} f \biggl( \frac{q[k+2\mu\theta _{k}]_{q}}{[n]_{q}} \biggr),\quad x\in[0,\infty ). $$
(2.1)

Lemma 2.1

Let \(D_{n,q}(\cdot;\cdot)\) be the operator given by (2.1). Then we have the following identities and inequalities:

  1. (1)

    \(D_{n,q}(e_{0};x)=1\),

  2. (2)

    \(D_{n,q}(e_{1};x)=x\),

  3. (3)

    \(x^{2}+[1-2 \mu]_{q} q^{2\mu+1}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} \frac{x}{[n]_{q}} \leq D_{n,q}(e_{2};x) \leq x^{2}+[1+2 \mu]_{q}\frac{x}{[n]_{q}}\).

Now in this paper, we define a q-Dunkl analog of the Szász-Mirakjan-Kantorovich operators as follows:

$$ K_{n,q}(f;x)=\frac{[n]_{q}}{e_{\mu,q}([n]_{q}\frac{x}{q})}\sum _{k=0}^{\infty}\frac{([n]_{q}x)^{k}}{\gamma_{\mu,q}(k)q^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}}_{\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}}}f(t)\,d_{q}t, $$
(2.2)

where \(\mu>\frac{1}{2}\), \(n\in\mathbb{N}\), \(0< q<1\), \(0\leq x<\frac{1}{1-q^{n}}\), and f is a continuous nondecreasing function on the interval \([0,\infty)\). It is seen that the operators \(K_{n,q}\) are linear and positive. In the case of \(q=1\), the operators \(K_{n,q}\) turn to the Szász-Mirakjan-Kantorovich operators [6].

Lemma 2.2

For \(\mu>\frac{1}{2}\), \(n\in\mathbb{N}\), \(0< q<1\), \(0\leq x<\frac{1}{1-q^{n}}\), and \(m\in\mathbb{N}_{0}\), we have a recurrence relation given by

$$ K_{n,q} \bigl(e_{m}(t);x \bigr)= \frac{1}{[m+1]_{q}}\sum_{j=0}^{m}\sum _{i=0}^{j}\binom{j}{i}\frac{1}{[n]_{q}^{j-i}}D_{n,q} \bigl(e_{m+i-j}(t);x \bigr). $$
(2.3)

Proof

$$ \int^{\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}}_{\frac{q[k+2\mu \theta_{k}]_{q}}{[n]_{q}}}t^{m}\,d_{q}t= \frac{1}{[m+1]_{q}} \biggl\{ \biggl(\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr)^{m+1}- \biggl(\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr)^{m+1} \biggr\} . $$

Using the expansion \(a^{m+1}-b^{m+1}=(a-b)(a^{m}+a^{m-1}b+\cdots+ab^{m-1}+b^{m})\), we have

$$\begin{aligned} \int^{\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}}_{\frac{q[k+2\mu \theta_{k}]_{q}}{[n]_{q}}}t^{m}\,d_{q}t =& \frac{1}{[m+1]_{q}} \biggl(\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}-\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr) \\ &{}\times \sum_{j=0}^{m} \biggl( \frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr)^{j} \biggl(\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr)^{m-j}. \end{aligned}$$

Using \([k+1+2\mu\theta_{k}]_{q}=1+q[k+2\mu\theta_{k}]_{q}\),

$$\begin{aligned} \int^{\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}}_{\frac{q[k+2\mu \theta_{k}]_{q}}{[n]_{q}}}t^{m}\,d_{q}t =& \frac{1}{[m+1]_{q}[n]_{q}} \sum_{j=0}^{m} \biggl( \frac{1+q[k+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr)^{j} \biggl(\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr)^{m-j} \\ =&\frac{1}{[m+1]_{q}[n]_{q}}\sum_{j=0}^{m}\sum _{i=0}^{j}\binom{j}{i} \biggl(\frac{q^{i}[k+2\mu\theta_{k}]_{q}^{i}}{[n]_{q}^{j}} \biggr) \biggl(\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}} \biggr)^{m-j} \\ =&\frac{1}{[m+1]_{q}[n]_{q}}\sum_{j=0}^{m}\sum _{i=0}^{j}\binom{j}{i} \frac{q^{m+i-j}[k+2\mu\theta_{k}]_{q}^{m+i-j}}{[n]_{q}^{m}}. \end{aligned}$$

From (2.2), we have

$$\begin{aligned} &{K_{n,q} \bigl(e_{m}(t);x \bigr)} \\ &{\quad =\frac{[n]_{q}}{e_{\mu,q}([n]_{q}\frac{x}{q})}\sum_{k=0}^{\infty} \frac{([n]_{q}x)^{k}}{\gamma_{\mu,q}(k)q^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}}_{\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}}}t^{m}\,d_{q}t} \\ &{\quad =\frac{[n]_{q}}{e_{\mu,q}([n]_{q}\frac{x}{q})}\sum_{k=0}^{\infty} \frac{([n]_{q}x)^{k}}{\gamma_{\mu,q}(k)q^{k}} \frac{1}{[m+1]_{q}[n]_{q}}\sum _{j=0}^{m} \sum_{i=0}^{j} \binom{j}{i} \frac{q^{m+i-j}[k+2\mu \theta_{k}]_{q}^{m+i-j}}{[n]_{q}^{m}}} \\ &{\quad =\frac{1}{[m+1]_{q}e_{\mu,q}([n]_{q}\frac{x}{q})}\sum_{j=0}^{m}\sum_{i=0}^{j}\binom{j}{i} \frac{1}{[n]_{q}^{j-i}} \sum_{k=0}^{\infty} \frac{([n]_{q}x)^{k}}{\gamma_{\mu,q}(k)q^{k}}\frac{q^{m+i-j}[k+2\mu \theta_{k}]_{q}^{m+i-j}}{[n]_{q}^{m+i-j}}} \\ &{\quad =\frac{1}{[m+1]_{q}}\sum_{j=0}^{m} \sum_{i=0}^{j}\binom{j}{i} \frac{1}{[n]_{q}^{j-i}}D_{n,q} \bigl(e_{m+i-j}(t);x \bigr).} \end{aligned}$$

 □

Lemma 2.3

Let \(e_{i}(x)=x^{i}\) (\(i=0,1,2\)) and \(K_{n,q}(\cdot;\cdot)\) be the operator defined by (2.2). Then we have the following identities and inequalities:

  1. (1)

    \(K_{n,q}(e_{0};x)=1\),

  2. (2)

    \(K_{n,q}(e_{1};x)=\frac{1}{[2]_{q}[n]_{q}}+\frac{2x}{[2]_{q}}\),

  3. (3)

    \(\frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{3x}{[3]_{q}[n]_{q}} (1+q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} ) +\frac{3x^{2}}{[3]_{q}}\leq K_{n,q}(e_{2};x)\leq\frac{1}{[3]_{q}[n]_{q}^{2}}+ \frac{3x}{[3]_{q}[n]_{q}} (1+[1+2\mu]_{q} )+\frac{3x^{2}}{[3]_{q}}\),

  4. (4)

    \(K_{n,q}((e_{1}-e_{0}x);x)=\frac{1}{[2]_{q}[n]_{q}}+ (\frac{2}{[2]_{q}}-1 )x\),

  5. (5)

    \(\frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{x}{[3]_{q}[n]_{q}} \{3q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})}+ (3-\frac{2[3]_{q}}{[2]_{q}} )\} + (\frac{3}{[3]_{q}}-\frac{4}{[2]_{q}}+1 )x^{2}\leq K_{n,q}((e_{1}-e_{0}x)^{2};x)\leq \frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{x}{[3]_{q}[n]_{q}}\{3[1+2\mu]_{q} + (3-\frac{2[3]_{q}}{[2]_{q}} )\}+ (\frac{3}{[3]_{q}}-\frac{4}{[2]_{q}}+1 )x^{2}\).

Proof

Here the proof is based on Lemma 2.2, and we can calculate only \(K_{n,q}(e_{2};x)\) and \(K_{n,q}((e_{1}-e_{0}x)^{2};x)\):

Put \(m=2\) in (2.3), we have

$$\begin{aligned} K_{n,q} \bigl(e_{2}(t);x \bigr) =&\frac{1}{[3]_{q}}\sum _{j=0}^{2}\sum_{i=0}^{j} \binom{j}{i}\frac{1}{[n]_{q}^{j-i}}D_{n,q} \bigl(e_{2+i-j}(t);x \bigr) \\ =&\frac{1}{[3]_{q}} \biggl(3D_{n,q}(e_{2};x)+ \frac{3}{[n]_{q}}D_{n,q}(e_{1};x)+\frac{1}{[n]_{q}^{2}}D_{n,q}(e_{0};x) \biggr). \end{aligned}$$

Using Lemma (2.1), we have

$$\begin{aligned} K_{n,q} \bigl(e_{2}(t);x \bigr) \geq&\frac{1}{[3]_{q}} \biggl(3 \biggl(x^{2}+[1-2 \mu]_{q} q^{2\mu+1}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} \frac{x}{[n]_{q}} \biggr)+\frac{3x}{[n]_{q}}+ \frac{1}{[n]_{q}^{2}} \biggr) \\ \geq&\frac{3x^{2}}{[3]_{q}}+\frac{3x}{[3]_{q}[n]_{q}} \biggl(1+[1-2\mu]_{q} q^{2\mu+1}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} \biggr)+\frac{1}{[3]_{q}[n]_{q}^{2}}, \end{aligned}$$

on the other hand, we have

$$\begin{aligned} K_{n,q} \bigl(e_{2}(t);x \bigr) \leq&\frac{1}{[3]_{q}} \biggl(3 \biggl(x^{2}+[1+2 \mu]_{q} \frac{x}{[n]_{q}} \biggr)+ \frac{3x}{[n]_{q}}+\frac{1}{[n]_{q}^{2}} \biggr) \\ \leq&\frac{3x^{2}}{[3]_{q}}+\frac{3x}{[3]_{q}[n]_{q}} \bigl(1+[1+2\mu]_{q} \bigr)+\frac{1}{[3]_{q}[n]_{q}^{2}}. \end{aligned}$$

Now, we have to prove (5). By linearity of \(K_{n,q}\) and from (1), (2), (3), we have

$$\begin{aligned} K_{n,q} \bigl((e_{1}-e_{0}x)^{2};x \bigr) =&K_{n,q}(e_{2};x)-2xK_{n,q}(e_{1};x)+x^{2}K_{n,q}(e_{0};x) \\ \geq&\frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{3x}{[3]_{q}[n]_{q}} \biggl(1+q^{2\mu+1}[1-2 \mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} \biggr) +\frac{3x^{2}}{[3]_{q}} \\ &{}-2x \biggl(\frac{1}{[2]_{q}[n]_{q}}+\frac{2x}{[2]_{q}} \biggr)+x^{2} \\ \geq& \frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{x}{[3]_{q}[n]_{q}} \biggl\{ 3q^{2\mu+1}[1-2 \mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})}+ \biggl(3-\frac{2[3]_{q}}{[2]_{q}} \biggr) \biggr\} \\ &{}+ \biggl(\frac{3}{[3]_{q}}-\frac{4}{[2]_{q}}+1 \biggr)x^{2}. \end{aligned}$$

Similarly, on the other hand

$$\begin{aligned} K_{n,q} \bigl((e_{1}-e_{0}x)^{2};x \bigr) \leq&\frac{1}{[3]_{q}[n]_{q}^{2}}+ \frac{3x}{[3]_{q}[n]_{q}} \bigl(1+[1+2\mu]_{q} \bigr)+\frac{3x^{2}}{[3]_{q}} \\ &{}-2x \biggl(\frac{1}{[2]_{q}[n]_{q}}+ \frac{2x}{[2]_{q}} \biggr)+x^{2} \\ \leq&\frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{x}{[3]_{q}[n]_{q}} \biggl\{ 3[1+2 \mu]_{q} + \biggl(3-\frac{2[3]_{q}}{[2]_{q}} \biggr) \biggr\} \\ &{}+ \biggl(\frac{3}{[3]_{q}}-\frac{4}{[2]_{q}}+1 \biggr)x^{2}. \end{aligned}$$

 □

Now, we want to transform the operators defined at (2.2) in order to preserve the linear function \(e_{1}\). Let \({r_{n,q}(x)}\) be the following sequence of real valued continuous function defined on \([0,\infty)\) with \(0\leq r_{n,q}(x)<\infty\):

$$ r_{n,q}(x)=\frac{[2]_{q}x}{2}-\frac{1}{2[n]_{q}},\qquad \frac{1}{[2]_{q}[n]_{q}}\leq x< \frac{1}{1-q^{n}},\quad n\in\mathbb{N}. $$
(2.4)

Then we consider the following linear positive operators:

$$\begin{aligned} \bar{K}_{n,q}(f;x) =&K_{n,q} \bigl(f;r_{n,q}(x) \bigr) \\ =& \frac{[n]_{q}}{e_{\mu,q}([n]_{q}\frac{r_{n,q}(x)}{q})}\sum_{k=0}^{\infty} \frac{([n]_{q}r_{n,q}(x))^{k}}{\gamma_{\mu,q}(k)q^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}}_{\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}}}f(t)\,d_{q}t, \end{aligned}$$
(2.5)

where f be a continuous and nondecreasing function on the interval \([0,\infty)\).

Lemma 2.4

Let \(\bar{K}_{n,q}(f;x)\) be the operator defined by (2.5). Then, for each \(\frac{1}{[2]_{q}[n]_{q}}\leq x<\frac{1}{1-q^{n}}\), we have

  1. (1)

    \(\bar{K}_{n,q}(e_{0};x)=1\),

  2. (2)

    \(\bar{K}_{n,q}(e_{1};x)=x\),

  3. (3)

    \(\frac{1}{[3]_{q}[n]_{q}^{2}} (\frac{1}{4}-\frac{3}{2}q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q}\frac{r_{n,q}(x)}{q})} ) +\frac{3[2]_{q}x}{2[3]_{q}[n]_{q}}q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q}\frac{r_{n,q}(x)}{q})} +\frac{3[2]_{q}^{2}}{4[3]_{q}}x^{2}\leq\bar{K}_{n,q}(e_{2};x)\leq \frac{1}{[3]_{q}[n]_{q}^{2}} (\frac{1}{4}-\frac{3}{2}[1+2\mu]_{q} )+\frac{3[2]_{q}}{2[3]_{q}[n]_{q}}[1+2\mu]_{q}x+ \frac{3[2]_{q}^{2}}{4[3]_{q}}x^{2}\),

  4. (4)

    \(\bar{K}_{n,q}((e_{1}-e_{0}x);x)=0\),

  5. (5)

    \(\frac{1}{[3]_{q}[n]_{q}^{2}} (\frac{1}{4}-\frac{3}{2}q^{2\mu+1}[1-2\mu]_{q} \frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q}\frac{r_{n,q}(x)}{q})} ) +\frac{3[2]_{q}x}{2[3]_{q}[n]_{q}}q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q}\frac{r_{n,q}(x)}{q})} + (\frac{3[2]_{q}^{2}}{4[3]_{q}}-1 )x^{2}\leq\bar{K}_{n,q}((e_{1}-e_{0}x)^{2};x)\leq \frac{1}{[3]_{q}[n]_{q}^{2}} (\frac{1}{4}-\frac{3}{2}[1+2\mu]_{q} )+\frac{3[2]_{q}}{2[3]_{q}[n]_{q}}[1+2\mu]_{q}x+ (\frac{3[2]_{q}^{2}}{4[3]_{q}}-1 )x^{2}\).

Proof

Using Lemma 2.3 and (2.4), we have

$$\begin{aligned} &{\bar{K}_{n,q}(e_{0};x)=1,} \\ &{\bar{K}_{n,q}(e_{1};x)=\frac{1}{[2]_{q}[n]_{q}}+ \frac{2r_{n,q}(x)}{[2]_{q}}} \\ &{\phantom{\bar{K}_{n,q}(e_{1};x)} =\frac{1}{[2]_{q}[n]_{q}}+\frac{2}{[2]_{q}} \biggl( \frac{[2]_{q}x}{2}- \frac{1}{2[n]_{q}} \biggr)} \\ &{\phantom{\bar{K}_{n,q}(e_{1};x)} =x.} \end{aligned}$$

Also,

$$\begin{aligned} &{\frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{3r_{n,q}(x)}{[3]_{q}[n]_{q}} \biggl(1+q^{2\mu+1}[1-2 \mu]_{q}\frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q} \frac{r_{n,q}(x)}{q})} \biggr)+\frac{3(r_{n,q}(x))^{2}}{[3]_{q}}} \\ &{\quad \leq \bar{K}_{n,q}(e_{2};x) \leq \frac{1}{[3]_{q}[n]_{q}^{2}}+ \frac{3r_{n,q}(x)}{[3]_{q}[n]_{q}} \bigl(1+[1+2 \mu]_{q} \bigr)+\frac{3(r_{n,q}(x))^{2}}{[3]_{q}}.} \end{aligned}$$

Now,

$$\begin{aligned} &{\bar{K}_{n,q}(e_{2};x)} \\ &{\quad \geq \frac{1}{[3]_{q}[n]_{q}^{2}}+ \frac{3r_{n,q}(x)}{[3]_{q}[n]_{q}} \biggl(1+q^{2\mu+1}[1-2\mu]_{q} \frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q} \frac{r_{n,q}(x)}{q})} \biggr)+\frac{3(r_{n,q}(x))^{2}}{[3]_{q}}} \\ &{\quad \geq \frac{1}{[3]_{q}[n]_{q}^{2}}+\frac{3}{[3]_{q}[n]_{q}} \biggl( \frac{[2]_{q}x}{2}- \frac{1}{2[n]_{q}} \biggr) \biggl(1+q^{2\mu+1}[1-2 \mu]_{q} \frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q} \frac{r_{n,q}(x)}{q})} \biggr)} \\ &{\qquad {}+\frac{3}{[3]_{q}} \biggl(\frac{[2]_{q}x}{2}- \frac{1}{2[n]_{q}} \biggr)^{2}} \\ &{\quad \geq \frac{1}{[3]_{q}[n]_{q}^{2}} \biggl(\frac{1}{4}- \frac{3}{2}q^{2\mu+1}[1-2 \mu]_{q} \frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q} \frac{r_{n,q}(x)}{q})} \biggr)+\frac{3[2]_{q}x}{2[3]_{q}[n]_{q}}q^{2\mu+1}[1-2 \mu]_{q}} \\ &{\qquad{} \times\frac{e_{\mu,q}([n]_{q}r_{n,q}(x))}{e_{\mu,q}([n]_{q}\frac{r_{n,q}(x)}{q})}+ \biggl(\frac{3[2]_{q}^{2}}{4[3]_{q}}-1 \biggr)x^{2}.} \end{aligned}$$

Similarly, on the other hand

$$ \bar{K}_{n,q}(e_{2};x)\leq \frac{1}{[3]_{q}[n]_{q}^{2}} \biggl( \frac{1}{4}-\frac{3}{2}[1+2\mu]_{q} \biggr)+ \frac{3[2]_{q}}{2[3]_{q}[n]_{q}}[1+2\mu]_{q}x+ \frac{3[2]_{q}^{2}}{4[3]_{q}}x^{2}. $$

By using the linearity of \(\bar{K}_{n,q}\) and (1), (2), (3) of Lemma 2.4 we obtain (4) and (5). □

Let \({u_{n,q}(x)}\) be the following sequence of real valued continuous function defined on \([0,\infty)\) with \(0\leq u_{n,q}(x)<\infty\):

$$\begin{aligned} u_{n,q}(x) =& \frac{- (1+q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} )}{2[n]_{q}} \\ &{} +\sqrt{ \frac{ (1+q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} )^{2}}{4[n]_{q}^{2}}+ \frac{[3]_{q}}{3}x^{2}- \frac{1}{3[n]_{q}^{2}}}, \end{aligned}$$
(2.6)

where \(\frac{1}{\sqrt{[3]_{q}}[n]_{q}}\leq x<\frac{1}{1-q^{n}}\), \(n\in \mathbb{N}\). Then we consider the following linear positive operators:

$$\begin{aligned} K^{\ast}_{n,q}(f;x) =& K_{n,q} \bigl(f;u_{n,q}(x) \bigr) \\ =&\frac{[n]_{q}}{e_{\mu,q}([n]_{q}\frac{u_{n,q}(x)}{q})}\sum_{k=0}^{\infty} \frac{([n]_{q}u_{n,q}(x))^{k}}{\gamma_{\mu,q}(k)q^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q}}{[n]_{q}}}_{\frac{q[k+2\mu\theta_{k}]_{q}}{[n]_{q}}}f(t)\,d_{q}t, \end{aligned}$$
(2.7)

where f is a continuous and nondecreasing function on the interval \([0,\infty)\).

Lemma 2.5

Let \(K^{\ast}_{n,q}(f;x)\) be the operator defined by (2.7). Then, for each \(\frac{1}{\sqrt{[3]_{q}}[n]_{q}}\leq x<\frac{1}{1-q^{n}}\), we have

  1. (1)

    \(K^{\ast}_{n,q}(e_{0};x)=1\),

  2. (2)

    \(K^{\ast}_{n,q}(e_{1};x)=\frac{2}{[2]_{q}} (\frac{-q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})}}{2[n]_{q}}+\sqrt{\frac{1}{4[n]_{q}^{2}} (1+q^{2\mu+1}[1-2\mu]_{q}\frac{e_{\mu,q}([n]_{q}x)}{e_{\mu,q}([n]_{q}\frac{x}{q})} )^{2} +\frac{[3]_{q}}{3}x^{2}-\frac{1}{3[n]_{q}^{2}}} )\),

  3. (3)

    \(K^{\ast}_{n,q}(e_{2};x)=x^{2}\).

Proof

Using Lemma 2.3, (2.6) and following similar steps to Lemma 2.4, we have the proof of Lemma 2.5. □

3 Korovkin’s and weighted Korovkin’s type approximation properties

In order to obtain the convergence results for our constructed operators, we take \(q=q_{n}\) where \((q_{n})\) be a sequence in the interval \((0,1)\) so that

$$ \lim_{n\rightarrow\infty}q_{n}=1\quad\mbox{and}\quad \lim_{n\rightarrow\infty}\frac{1}{[n]_{q_{n}}}=0. $$
(3.1)

We obtain the Korovkin’s type approximation properties for our constructed operators \(\bar{K}_{n,q}(\cdot;\cdot)\), \(K^{\ast}_{n,q}(\cdot;\cdot)\) defined by (2.5) and (2.7), respectively.

Theorem 3.1

Let \((q_{n})\) be a sequence satisfying (3.1) and \(\bar{K}_{n,q_{n}}(\cdot;\cdot)\) be the operator given by (2.5). Then, for each nondecreasing \(f\in C_{\gamma}[0,\infty)\), we have

$$ \lim_{n\rightarrow\infty}\bar{K}_{n,q_{n}}(f;x)=f(x) $$

uniformly with respect to \(x\in[{{{\frac{1}{[2]_{q_{n}}[n]_{q_{n}}},a}}}]\) provided \(\gamma\geq2\) and \(a>\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\).

Proof

The proof is based on the well known Korovkin’s theorem regarding the convergence of a sequence of linear and positive operators; so, it is enough to prove the conditions

$$ \lim_{n\rightarrow\infty}\bar{K}_{n,q_{n}}(e_{i};x)=e_{i}(x) \quad\mbox{for }i=0,1,2. $$

From Lemma 2.4 and (3.1), the result follows. □

Theorem 3.2

Let \((q_{n})\) be a sequence satisfying (3.1) and \(K^{\ast}_{n,q_{n}}(\cdot;\cdot)\) be the operator given by (2.7). Then, for each nondecreasing \(f\in C_{\gamma}[0,\infty)\), we have

$$ \lim_{n\rightarrow\infty}K^{\ast}_{n,q_{n}}(f;x)=f(x) $$

uniformly with respect to \(x\in[{{{\frac{1}{\sqrt{[3]_{q_{n}}}[n]_{q_{n}}},b}}}] \) provided \(\gamma\ge2\) and \(b>\frac{1}{\sqrt{[3]_{q_{n}}}[n]_{q_{n}}}\).

The weighted space of the functions which are defined on the positive semi axis \(\mathbb{R}^{+}=[0,\infty)\) is addressed as follows:

Let \(P_{\rho}(\mathbb{R}^{+})\) be the set of all functions f satisfying the condition \(\vert f(x)\vert \leq M_{f}\rho(x)\), where \(x\in\mathbb{R}^{+}\) and \(M_{f}\) is a constant depending on f. Introduce

$$\begin{aligned} &{Q_{\rho} \bigl(\mathbb{R}^{+} \bigr)=P_{\rho} \bigl( \mathbb{R}^{+} \bigr)\cap C[0,\infty),} \\ &{Q_{\rho}^{k} \bigl(\mathbb{R}^{+} \bigr)= \biggl\{ f:f \in Q_{\rho} \bigl(\mathbb{R}^{+} \bigr) \mbox{ and } \lim_{x\rightarrow\infty}\frac{f(x)}{\rho(x)}=k\mbox{ (constant)} \biggr\} ,} \end{aligned}$$

where \(\rho(x)=1+x^{2}\) is a weight function. These spaces are endowed with the norm

$$ \Vert f\Vert _{\rho}=\sup_{x\in[0,\infty)} \frac{\vert f(x)\vert }{\rho(x)}. $$

Theorem 3.3

Let \((q_{n})\) be a sequence satisfying (3.1) and \(K_{n,q_{n}}(\cdot;\cdot)\) be the operator defined by (2.2). Then, for each function \(f\in Q_{\rho}^{k}(\mathbb{R}^{+})\), we have

$$ \lim_{n\rightarrow\infty} \bigl\Vert K_{n,q_{n}}(f;x)-f \bigr\Vert _{\rho}=0. $$

Proof

Using the Korovkin-type theorem on weighted approximations in [25], we see that it is sufficient to verify the following three conditions:

$$ \lim_{n\rightarrow\infty} \bigl\Vert K_{n,q_{n}} \bigl(e_{i}(t);x \bigr)-e_{i}(x) \bigr\Vert _{\rho}=0,\quad i=0,1,2. $$
(3.2)

Since \(K_{n,q_{n}}(e_{0}(t);x)=1\), (3.2) holds for \(i=0\).

Using Lemma 2.3, we have

$$\begin{aligned} \bigl\Vert K_{n,q_{n}} \bigl(e_{1}(t);x \bigr)-e_{1}(x) \bigr\Vert _{\rho} =&\sup_{x\in[0,\infty)}\frac{\vert K_{n,q_{n}}(e_{1}(t);x)-e_{1}(x)\vert }{1+x^{2}} \\ =&\sup_{x\in[0,\infty)}\frac{| \frac{1}{[2]_{q_{n}}[n]_{q_{n}}}+\frac{2x}{[2]_{q_{n}}}-x| }{1+x^{2}} \\ \leq&\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\sup_{x\in[0,\infty)}\frac{1}{1+x^{2}}+ \biggl(\frac{2}{[2]_{q_{n}}}-1 \biggr)\sup_{x\in[0,\infty)} \frac{x}{1+x^{2}} \\ \leq&\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}+ \biggl(\frac{2}{[2]_{q_{n}}}-1 \biggr), \end{aligned}$$

which implies that (3.2) holds for \(i=1\) as \(n\rightarrow\infty\). Similarly, we can write

$$\begin{aligned} \bigl\Vert K_{n,q_{n}} \bigl(e_{2}(t);x \bigr)-e_{2}(x) \bigr\Vert _{\rho} =&\sup_{x\in[0,\infty)}\frac{\vert K_{n,q_{n}}(e_{2}(t);x)-e_{2}(x)\vert }{1+x^{2}} \\ \leq&\sup_{x\in[0,\infty)}\frac{| \frac{1}{[3]_{q_{n}}[n]_{q_{n}}^{2}}+ \frac{3x}{[3]_{q_{n}}[n]_{q_{n}}} (1+[1+2\mu]_{q_{n}} )+\frac{3x^{2}}{[3]_{q_{n}}}| -x^{2}}{1+x^{2}} \\ \leq&\frac{1}{[3]_{q_{n}}[n]_{q_{n}}^{2}}\sup_{x\in[0,\infty)}\frac{1}{1+x^{2}}+ \frac{3(1+[1+2\mu]_{q_{n}})}{[3]_{q_{n}}[n]_{q_{n}}}\sup_{x\in[0,\infty)}\frac{x}{1+x^{2}} \\ &{}+ \biggl(\frac{3}{[3]_{q_{n}}}-1 \biggr)\sup_{x\in[0,\infty)} \frac{x^{2}}{1+x^{2}} \\ \leq&\frac{1}{[3]_{q_{n}}[n]_{q_{n}}^{2}}+\frac{3(1+[1+2\mu]_{q_{n}})}{[3]_{q_{n}}[n]_{q_{n}}} + \biggl(\frac{3}{[3]_{q_{n}}}-1 \biggr), \end{aligned}$$

which implies that

$$ \lim_{n\rightarrow\infty} \bigl\Vert K_{n,q_{n}} \bigl(e_{2}(t);x \bigr)-e_{2}(x) \bigr\Vert _{\rho}=0. $$

 □

Theorem 3.4

Let \(\bar{K}_{n,q_{n}}(f;x)\) be the operator defined by (2.5). Then, for each function \(f\in Q_{\rho}^{k}(\mathbb{R}^{+})\), we have

$$ \lim_{n\rightarrow\infty} \bigl\Vert \bar{K}_{n,q_{n}}(f;x)-f \bigr\Vert _{\rho}=0. $$

Proof

In order to prove this theorem it is sufficient to verify (3.2). Since

$$ \bar{K}_{n,q_{n}} \bigl(e_{0}(t);x \bigr)=1,\qquad \bar{K}_{n,q_{n}} \bigl(e_{1}(t);x \bigr)=x, $$

we can easily see that (3.2) holds for \(i=0,1\). By using Lemma 2.4, we have

$$\begin{aligned} &{ \bigl\Vert \bar{K}_{n,q_{n}} \bigl(e_{2}(t);x \bigr)-e_{2}(x) \bigr\Vert _{\rho}} \\ &{\quad=\sup_{x\in[0,\infty)}\frac{\vert \bar{K}_{n,q_{n}}(e_{2}(t);x)-e_{2}(x)\vert }{1+x^{2}}} \\ &{\quad\leq\sup_{x\in[0,\infty)}\frac{\frac{1}{[3]_{q_{n}}[n]_{q_{n}}^{2}} (\frac{1}{4}-\frac{3}{2}[1+2\mu]_{q_{n}} )+ \frac{3[2]_{q_{n}}}{2[3]_{q_{n}}[n]_{q_{n}}}[1+2\mu]_{q_{n}}x+\frac{3[2]_{q_{n}}^{2}}{4[3]_{q_{n}}}x^{2}-x^{2}}{1+x^{2}}} \\ &{\quad\leq\frac{1}{[3]_{q_{n}}[n]_{q_{n}}^{2}} \biggl(\frac{1}{4}- \frac{3}{2}[1+2 \mu]_{q_{n}} \biggr)\sup _{x\in[0,\infty)}\frac{1}{1+x^{2}}} \\ &{\qquad{} {}+\frac{3[2]_{q_{n}}}{2[3]_{q_{n}}[n]_{q_{n}}}[1+2\mu]_{q_{n}}\sup _{x\in[0,\infty)} \frac{x}{1+x^{2}} + \biggl(\frac{3[2]_{q_{n}}^{2}}{4[3]_{q_{n}}}-1 \biggr)\sup_{x\in[0,\infty)}\frac{x^{2}}{1+x^{2}}} \\ &{\quad\leq\frac{1}{[3]_{q_{n}}[n]_{q_{n}}^{2}} \biggl(\frac{1}{4}- \frac{3}{2}[1+2 \mu]_{q_{n}} \biggr) + \frac{3[2]_{q_{n}}}{2[3]_{q_{n}}[n]_{q_{n}}}[1+2\mu]_{q_{n}} + \biggl( \frac{3[2]_{q_{n}}^{2}}{4[3]_{q_{n}}}-1 \biggr),} \end{aligned}$$

which implies that

$$ \lim_{n\rightarrow\infty} \bigl\Vert \bar{K}_{n,q_{n}} \bigl(e_{2}(t);x \bigr)-e_{2}(x) \bigr\Vert _{\rho}=0. $$

 □

4 Rate of convergence

In this section we compute rate of convergence of the constructed operators in terms of the modulus of continuity and the class of Lipschitz functions:

Let \(f\in C_{B}[0,\infty)\), the space of all bounded and continuous functions on \([0,\infty)\). Then, for any \(\delta>0\), \(x\geq0\) the modulus of continuity is denoted by \(\omega(f,\delta)\) and is defined as

$$ \omega(f,\delta)=\sup_{\vert t-x\vert \leq\delta,\quad t\in[0,\infty)} \bigl\vert f(t)-f(x) \bigr\vert . $$
(4.1)

Also,

$$ \bigl\vert f(t)-f(x) \bigr\vert \leq\omega(f,\delta) \biggl(1+ \frac{\vert t-x\vert }{\delta} \biggr). $$
(4.2)

If \(f(x)\) is uniformly continuous on \([0,\infty)\) then it is necessary and sufficient that

$$ \lim_{\delta\rightarrow0}\omega(f,\delta)=0. $$

In order to obtain the convergence result we use the following lemma.

Lemma 4.1

[20]

Let \(0< q<1\) and \(a\in[0,bq]\), \(b>0\). The inequality

$$ \int_{a}^{b}\vert t-x\vert \,d_{q}t \leq \biggl( \int_{a}^{b}(t-x)^{2}\,d_{q}t \biggr)^{\frac{1}{2}} \biggl( \int_{a}^{b}\,d_{q}t \biggr)^{\frac{1}{2}} $$

is satisfied.

Theorem 4.2

Let \((q_{n})\) be a sequence satisfying (3.1). For the operator \(K_{n,q_{n}}\) given by (2.2), for all nondecreasing \(f\in C_{B}[0,\infty)\), \(0\leq x<\frac{1}{1-q_{n}^{n}}\) and \(n\in\mathbb{N}\), we have

$$\begin{aligned} &{ \bigl\vert K_{n,q_{n}}(f;x)-f(x) \bigr\vert } \\ &{\!\quad \leq \biggl\{ 1} \\ &{\!\qquad{}+ \sqrt{\frac{1}{[3]_{q_{n}}}+[n]_{q_{n}} \biggl( \frac{3}{[3]_{q_{n}}} \bigl(1+[1+2\mu]_{q_{n}} \bigr) - \frac{2}{[2]_{q_{n}}} \biggr)x+[n]_{q_{n}}^{2} \biggl( \frac{3}{[3]_{q_{n}}}- \frac{4}{[2]_{q_{n}}}+1 \biggr)x^{2}} \biggr\} } \\ &{\!\qquad{}\times\omega \biggl(f,\frac{1}{\sqrt{[n]_{q_{n}}}} \biggr),} \end{aligned}$$

where \(\omega(f,\cdot)\) is the modulus of continuity of the function \(f\in C_{B}[0,\infty)\) defined in (4.1).

Proof

Let \(\mu>\frac{1}{2}\), \(n\in\mathbb{N}\), nondecreasing \(f\in C_{B}[0,\infty) \), \(\delta>0\), and \(0\leq x<\frac{1}{1-q_{n}^{n}}\). Applying linearity and monotonicity of \(K_{n,q_{n}}\) and using (4.2), we get

$$\begin{aligned} &{ \bigl\vert K_{n,q_{n}}(f;x)-f(x) \bigr\vert } \\ &{\quad = \bigl\vert K_{n,q_{n}} \bigl(f(t)-f(x);x \bigr) \bigr\vert \leq K_{n,q_{n}} \bigl( \bigl\vert f(t)-f(x) \bigr\vert ;x \bigr) \leq \omega(f,\delta) \biggl(1+\frac{K_{n,q_{n}}(\vert t-x\vert ;x)}{\delta} \biggr)} \\ &{\quad \leq\omega(f,\delta) \Biggl(1+\frac{1}{\delta} \frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})}\sum_{k=0}^{\infty} \frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu \theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}\vert t-x\vert \, d_{q_{n}}t \Biggr).} \end{aligned}$$

Using Lemma 4.1, with \(a=\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}\) and \(b=\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}\), we have

$$\begin{aligned} \bigl\vert K_{n,q_{n}}(f;x)-f(x) \bigr\vert \leq&\omega(f,\delta) \Biggl\{ 1+\frac{1}{\delta}\frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})} \sum _{k=0}^{\infty}\frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}} \\ &{}\times \biggl( \int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}(t-x)^{2}\,d_{q_{n}}t \biggr)^{\frac{1}{2}} \biggl( \int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}\,d_{q_{n}}t \biggr)^{\frac{1}{2}} \Biggr\} . \end{aligned}$$

Using the Hölder inequality for sums, we get

$$\begin{aligned} &{ \bigl\vert K_{n,q_{n}}(f;x)-f(x) \bigr\vert } \\ &{\quad \leq \omega(f,\delta) \Biggl\{ 1+\frac{1}{\delta} \Biggl(\frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})} \sum_{k=0}^{\infty} \frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}(t-x)^{2}\,d_{q_{n}}t \Biggr)^{\frac{1}{2}}} \\ &{\qquad {}\times \Biggl(\frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})} \sum _{k=0}^{\infty} \frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}\,d_{q_{n}}t \Biggr)^{\frac{1}{2}} \Biggr\} } \\ &{\quad \leq \omega(f,\delta) \biggl\{ 1+\frac{1}{\delta} \bigl(K_{n,q_{n}} \bigl((t-x)^{2};x \bigr) \bigr)^{\frac{1}{2}} \times \bigl(K_{n,q_{n}}(1;x) \bigr)^{\frac{1}{2}} \biggr\} } \\ &{\quad \leq \omega(f,\delta) \biggl\{ 1+\frac{1}{\delta} \bigl(K_{n,q_{n}} \bigl((t-x)^{2};x \bigr) \bigr)^{\frac{1}{2}} \biggr\} .} \end{aligned}$$

Choosing \(\delta=\delta_{n}=\frac{1}{\sqrt{[n]_{q_{n}}}}\) and using (5) of Lemma 2.3, we have the result. □

Theorem 4.3

Let \((q_{n})\) be a sequence satisfying (3.1). For the operator \(\bar{K}_{n,q_{n}}\) given by (2.5), for all nondecreasing \(f\in C_{B}[0,\infty)\), \(\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\leq x<\frac{1}{1-q_{n}^{n}}\) and \(n\in\mathbb{N}\), we have

$$\begin{aligned} &{ \bigl\vert \bar{K}_{n,q_{n}}(f;x)-f(x) \bigr\vert } \\ &{\quad \leq \biggl\{ 1} \\ &{\qquad{}+\sqrt{\frac{1}{[3]_{q_{n}}} \biggl(\frac{1}{4}- \frac{3}{2}[1+2\mu]_{q_{n}} \biggr)+\frac{3[2]_{q_{n}}[n]_{q_{n}}}{2[3]_{q_{n}}}[1+2 \mu]_{q_{n}}x+[n]_{q_{n}}^{2} \biggl( \frac{3[2]_{q_{n}}^{2}}{4[3]_{q_{n}}}-1 \biggr)x^{2}} \biggr\} } \\ &{\qquad{} \times\omega \biggl(f,\frac{1}{\sqrt{[n]_{q_{n}}}} \biggr),} \end{aligned}$$

where \(\omega(f,\cdot)\) is the modulus of continuity of the function \(f\in C_{B}[0,\infty)\) defined in (4.1).

Proof

Using (5) of Lemma 2.4 and following similar steps to Theorem 4.2, we have the proof of Theorem 4.3. □

Now we claim that the error estimation in Theorem 4.3 is better than that of Theorem 4.2 provided \(f\in C_{B}[0,\infty)\) and \(\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\leq x<\frac{1}{1-q_{n}^{n}}\).

For \(\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\leq x<\frac{1}{1-q_{n}^{n}}\), \(\mu>\frac{1}{2}\), and \(n\in\mathbb{N}\) it is guaranteed that

$$\begin{aligned} &{\frac{1}{[3]_{q_{n}}} \biggl(\frac{1}{4}- \frac{3}{2}[1+2 \mu]_{q_{n}} \biggr)+\frac{3[2]_{q_{n}}[n]_{q_{n}}}{2[3]_{q_{n}}}[1+2\mu]_{q_{n}}x+[n]_{q_{n}}^{2} \biggl( \frac{3[2]_{q_{n}}^{2}}{4[3]_{q_{n}}}-1 \biggr)x^{2}} \\ &{\quad \leq \frac{1}{[3]_{q_{n}}}+[n]_{q_{n}} \biggl( \frac{3}{[3]_{q_{n}}} \bigl(1+[1+2\mu]_{q_{n}} \bigr) - \frac{2}{[2]_{q_{n}}} \biggr)x} \\ &{\qquad{}+[n]_{q_{n}}^{2} \biggl(\frac{3}{[3]_{q_{n}}}-\frac{4}{[2]_{q_{n}}}+1 \biggr)x^{2}.} \end{aligned}$$
(4.3)

If we put \(\mu=0\) in (4.3) then we have

$$\begin{aligned} &{\frac{1}{[3]_{q_{n}}} \biggl(\frac{1}{4}-\frac{3}{2} \biggr)+\frac{3[2]_{q_{n}}[n]_{q_{n}}}{2[3]_{q_{n}}}x+[n]_{q_{n}}^{2} \biggl( \frac{3[2]_{q_{n}}^{2}}{4[3]_{q_{n}}}-1 \biggr)x^{2}} \\ &{\quad \leq \frac{1}{[3]_{q_{n}}}+[n]_{q_{n}} \biggl( \frac{6}{[3]_{q_{n}}} -\frac{2}{[2]_{q_{n}}} \biggr)x+[n]_{q_{n}}^{2} \biggl(\frac{3}{[3]_{q_{n}}}-\frac{4}{[2]_{q_{n}}}+1 \biggr)x^{2}.} \end{aligned}$$

Again, if we put \(q_{n}=1\) then clearly

$$ nx-\frac{5}{12}\leq nx+\frac{1}{3}. $$

Now, we can also compute the rate of convergence of the our constructed operators in terms of the element of the usual Lipschitz class \(\operatorname{Lip}_{M}(\nu)\):

Let \(f\in C_{B}[0,\infty)\), \(M>0\), and \(0<\nu\leq1\). The class of \(\operatorname{Lip}_{M}(\nu)\) is defined as

$$ \operatorname{Lip}_{M}(\nu)= \bigl\{ f: \bigl\vert f( \zeta_{1})-f(\zeta_{2}) \bigr\vert \leq M\vert \zeta_{1}-\zeta_{2}\vert ^{\nu}, \zeta_{1},\zeta_{2}\in[0,\infty) \bigr\} . $$
(4.4)

Theorem 4.4

Let \((q_{n})\) be a sequence satisfying (3.1) and \(K_{n,q_{n}}\) be the operator defined in (2.2). Then, for each \(f\in \operatorname{Lip}_{M}(\nu)\) (\(M>0\), \(0<\nu\leq1\)) satisfying (4.4), we have

$$ \bigl\vert K_{n,q_{n}}(f;x)-f(x) \bigr\vert \leq M \bigl( \delta_{n}(x) \bigr)^{\frac{\nu}{2}}, $$
(4.5)

where \(\delta_{n}(x)=K_{n,q_{n}}((e_{1}-e_{0}x)^{2};x)\).

Proof

We prove this theorem by using (4.4) and Hölder’s inequality:

$$\begin{aligned} \bigl\vert K_{n,q_{n}}(f;x)-f(x) \bigr\vert =& \bigl\vert K_{n,q_{n}} \bigl(f(t)-f(x);x \bigr) \bigr\vert \leq K_{n,q_{n}} \bigl( \bigl\vert f(t)-f(x) \bigr\vert ;x \bigr) \\ \leq& M K_{n,q_{n}} \bigl(\vert t-x\vert ^{\nu};x \bigr). \end{aligned}$$

Therefore,

$$\begin{aligned} \bigl\vert K_{n,q_{n}}(f;x)-f(x) \bigr\vert \leq&M \frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})} \sum_{k=0}^{\infty}\frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}} \int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}} \vert t-x\vert ^{\nu} \,d_{q_{n}}t \\ \leq&M \frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})}\sum_{k=0}^{\infty} \biggl(\frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}} \biggr)^{\frac{\nu}{2}} \biggl(\frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}} \biggr)^{\frac{2-\nu}{2}} \\ &{}\times \int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}\vert t-x\vert ^{\nu} \,d_{q_{n}}t \\ \leq&M \Biggl(\frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})}\sum_{k=0}^{\infty} \frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}}\int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}} \,d_{q_{n}}t \Biggr)^{\frac{2-\nu }{2}} \\ &{}\times \Biggl(\frac{[n]_{q_{n}}}{e_{\mu,q_{n}}([n]_{q_{n}}\frac{x}{q_{n}})}\sum _{k=0}^{\infty} \frac{([n]_{q_{n}}x)^{k}}{\gamma_{\mu,q_{n}}(k)q_{n}^{k}}\int^{\frac{[k+1+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}}_{\frac{q_{n}[k+2\mu\theta_{k}]_{q_{n}}}{[n]_{q_{n}}}} \vert t-x\vert ^{2} \,d_{q_{n}}t \Biggr)^{\frac{\nu}{2}} \\ \leq&M \bigl(K_{n,q_{n}}(1;x) \bigr)^{\frac{2-\nu}{2}} \bigl( K_{n,q}(e_{1}-e_{0}x)^{2};x \bigr) )^{\frac{\nu}{2}} \\ \leq&M \bigl( K_{n,q_{n}} \bigl((e_{1}-e_{0}x)^{2};x \bigr) \bigr)^{\frac{\nu}{2}}. \end{aligned}$$

Choosing \(\delta_{n}(x)=K_{n,q_{n}}((e_{1}-e_{0}x)^{2};x)\), the proof is completed. □

Theorem 4.5

Let \((q_{n})\) be a sequence satisfying (3.1) and \(\bar{K}_{n,q_{n}}\) be the operator defined in (2.5). Then, for each \(f\in \operatorname{Lip}_{M}(\nu)\) (\(M>0\), \(0<\nu\leq1\)) satisfying (4.4), we have

$$ \bigl\vert \bar{K}_{n,q_{n}}(f;x)-f(x) \bigr\vert \leq M \bigl(\bar{\delta}_{n}(x) \bigr)^{\frac{\nu}{2}}, $$
(4.6)

where \(\bar{\delta}_{n}(x)=\bar{K}_{n,q_{n}}((e_{1}-e_{0}x)^{2};x)\).

Proof

Taking into account (5) of Lemma 2.4 and following similar steps to Theorem 4.4, we have the proof of Theorem 4.5. So we omit the details of the proof. □

From (4.3), it follows that the above claim also holds for Theorem 4.5, i.e., the rate of convergence of the operators \(\bar{K}_{n,q_{n}}\) by means of an element of the Lipschitz class functions is better than the ordinary error estimation given by (4.5), where \(x\geq\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\).

5 A Voronovskaja-type theorem

Now, we prove the Voronovskaja-type result for our modified Dunkl analog of q-Szász-Mirakjan-Kantorovich operators \(\bar{K}_{n,q_{n}}\).

Lemma 5.1

Let \((q_{n})\) be a sequence satisfying (3.1) and \(\bar{K}_{n,q_{n}}\) be the operator defined in (2.5). Therefore for every \(x\geq\frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\) there holds,

$$\begin{aligned} &{\lim_{ n\rightarrow\infty}[n]_{q_{n}} \bar{K}_{n,q_{n}}(e_{1}-e_{0}x;x)=0,} \end{aligned}$$
(5.1)
$$\begin{aligned} &{(1-2\mu)x\leq\lim _{ n\rightarrow\infty}[n]_{q_{n}} \bar{K}_{n,q_{n}} \bigl((e_{1}-e_{0}x)^{2};x \bigr)\leq(1+2\mu)x.} \end{aligned}$$
(5.2)

Theorem 5.2

Let \(q=q_{n}\) satisfies (3.1) and \(\bar{K}_{n,q_{n}}\) be the operator defined in (2.5). For any \(f\in C_{\gamma}[0,\infty)\) such that \(f^{\prime},f^{\prime\prime}\in C_{\gamma}[0,\infty)\), \(\gamma\geq2 \), we have

$$ \lim_{ n\rightarrow\infty}[n]_{q_{n}} \bigl(\bar{K}_{n,q_{n}}(f;x)-f(x) \bigr)\leq\frac{1}{2}(1+2\mu)xf^{\prime\prime}(x), $$

uniformly with respect to \(x\in[{{{\frac{1}{[2]_{q_{n}}[n]_{q_{n}}},a}}}]\) \((a>\frac{1}{[2]_{q_{n}}[n]_{q_{n}}})\).

Proof

Let \(f,f^{\prime },f^{\prime \prime }\in C_{\gamma }[0,\infty )\) and \(x\geq \frac{1}{[2]_{q_{n}}[n]_{q_{n}}}\). By the Taylor formula, we write

$$ f(t)=f(x)+(t-x)f^{\prime }(x)+\frac{1}{2}(t-x)^{2}f^{\prime \prime }(x)+(t-x)^{2}r(t;x), $$
(5.3)

where \(r(t;x)\) is the Peano form of the remainder \(r(\cdot;x)\in C_{\gamma }[0,\infty )\) and \(\lim_{t\rightarrow x}r(t;x)=0\). Applying \(\bar{K}_{n,q_{n}}\) to (5.3), we obtain

$$\begin{aligned} \bar{K}_{n,q_{n}} \bigl(f(t);x \bigr)-f(x) =&f^{\prime }(x) \bar{K}_{n,q_{n}} \bigl((e_{1}-e_{0}x);x \bigr)+\frac{1}{2}f^{\prime \prime }(x)\bar{K}_{n,q_{n}} \bigl((e_{1}-e_{0}x)^{2};x \bigr) \\ &{}+\bar{K}_{n,q_{n}} \bigl(r(t;x) (e_{1}-e_{0}x)^{2};x \bigr). \end{aligned}$$

By the Cauchy-Schwartz inequality, we have

$$ \bar{K}_{n,q_{n}} \bigl(r(\cdot;x) (\cdot-x)^{2};x \bigr)\leq \sqrt{ \bar{K}_{n,q_{n}} \bigl(r^{2}(\cdot;x);x \bigr)} \sqrt{ \bar{K}_{n,q_{n}} \bigl((\cdot-x)^{4};x \bigr)}. $$
(5.4)

Let \(\eta (\cdot;x):=r^{2}(\cdot;x)\). In this case observe that \(\eta (x;x)=0\) and \(\eta (\cdot;x)\in C_{\gamma }[0,\infty )\). Then it follows from Theorem 3.1 that

$$ \lim_{n\rightarrow \infty }\bar{K}_{n,q_{n}} \bigl(r^{2}( \cdot;x);x \bigr)=\lim_{n\rightarrow \infty }\bar{K}_{n,q_{n}} \bigl(\eta ( \cdot;x);x \bigr)=\eta (x;x)=0 $$
(5.5)

uniformly with respect to \(x\in {}[ {{\frac{1}{[2]_{q_{n}}[n]_{q_{n}}},a}}]\). Now consider (5.4), (5.5) and using Lemma 5.1, we have

$$\begin{aligned} &{\lim_{n\rightarrow \infty }[n]_{q_{n}}\bar{K}_{n,q_{n}} \bigl(f(t);x \bigr)-f(x)} \\ &{\quad = f^{\prime }(x)\lim_{n\rightarrow \infty }[n]_{q_{n}} \bar{K}_{n,q_{n}} \bigl((e_{1}-e_{0}x);x \bigr) + \frac{1}{2}f^{\prime \prime }(x)\lim_{n\rightarrow \infty }[n]_{q_{n}}\bar{K}_{n,q_{n}} \bigl((e_{1}-e_{0}x)^{2};x \bigr)} \\ &{\quad \leq \frac{1}{2}(1+2\mu )xf^{\prime \prime }(x).} \end{aligned}$$

This completes the proof. □

Remark

The further properties of the operators such as convergence properties via summability methods (see, for example, [2, 3, 10]) can be studied.

Conclusion

In this paper we have constructed and investigated a Dunkl analog of the q-Szász-Mirakjan-Kantorovich operators which preserves the test functions \(e_{1}\) and \(e_{2}\). We have showed that our modified operators have a better error estimation than the classical ones. We have also obtained some approximation results with the help of the well-known Korovkin theorem and the weighted Korovkin theorem for these operators. Furthermore, we studied convergence properties in terms of the modulus of continuity and the class of Lipschitz functions.