Skip to main content

Theory and Modern Applications

Approximation by a power series summability method of Kantorovich type Szász operators including Sheffer polynomials

Abstract

The main purpose of this paper is to use a power series summability method to study some approximation properties of Kantorovich type Szász–Mirakyan operators including Sheffer polynomials. We also establish Voronovskaya type result.

1 Introduction and background

Let \(\mathcal{K}_{m}=\{i\leq m:i\in \mathcal{K}\subseteq \mathbb{N}\}\). Then the natural density of \(\mathcal{K}\) is defined by \(\sigma (\mathcal{K})=\lim_{m}\frac{1}{m}{|\mathcal{K}_{m}|}\) provided the limit exists, where \(|\mathcal{K}_{m}|\) denotes the cardinality of \(\mathcal{K}_{m}\). A sequence \(\eta =(\eta _{i})\) is “statistically convergent” (see [9]) to \(\mathfrak{s}\) if for every \(\epsilon >0\)

$$ \lim_{m}\frac{1}{m} \bigl\vert \bigl\{ i\leq m: \vert \eta _{i}-\mathfrak{s} \vert \geq \epsilon \bigr\} \bigr\vert =0 $$

and we write \(st-\lim_{m}\eta {_{m}}=\mathfrak{s}\).

Let \(\mathfrak{T}=(\mathfrak{d}_{ij})\) be an infinite matrix. It is said to be regular if it transforms a convergent sequence into a convergent one with the same limit.

Let \(\mathfrak{T}=(\mathfrak{d}_{ij})\) be regular matrix. A sequence \(\zeta =(\zeta _{j})\) is said to be \(\mathfrak{T}\)-statistically convergent (see [10]) to the number \(\mathfrak{s}\) if, for any \(\epsilon >0\), \(\lim_{i}{\sum_{j:|\eta _{j}-\mathfrak{s}|\geq \epsilon }} \mathfrak{d}_{ij}=0\), and denote \(st_{\mathfrak{T}}-\lim \eta =\mathfrak{s}\). If

$$ \mathfrak{d}_{ij}= \textstyle\begin{cases} \frac{1}{j}, & i\leq j, \\ 0; & i>j.\end{cases} $$
(1.1)

Then it reduces to statistical convergence.

For a sequence of positive real numbers \((\mathfrak{p}_{j})\), denote the corresponding power series \(\mathfrak{p}(\mathfrak{y})=\sum_{j=1}^{\infty }\mathfrak{p}{_{j}}\mathfrak{y}^{j-1}\) which has radius of convergence \(R>0\). A sequence \(\eta =(\eta _{j})\) is convergent in the sense of power series method (see [12, 21]) if \(\lim_{\mathfrak{y}\rightarrow R^{-}}{ \frac{1}{\mathfrak{p}(\mathfrak{y})}\sum_{j=1}^{\infty }}\eta _{j}{{ \mathfrak{p}}}_{j}\mathfrak{y}{{^{j-1}}}=\mathcal{L}\) for all \(\mathfrak{y}\in (0,R)\). Moreover, the power series method is regular if and only if \(\lim_{\mathfrak{y}\rightarrow R^{-}}{ \frac{{{\mathfrak{p}}}_{j}\mathfrak{y}{{^{j-1}}}}{\mathfrak{p}(\mathfrak{y})}}=0\) holds for each \(j\in \{1,2,\ldots \}\) (see [2]). The power series method is more effective than the ordinary convergence (see [22, 23]). For more summability methods, see [35, 7, 13, 1519].

We study a Korovkin type theorem for the Kantorovich type generalization of Szász operators involving Sheffer polynomials via power series method. We determine the rate of convergence for these operators. Furthermore, we give a Voronovskaya type theorem for \(\mathfrak{T}\)-statistical convergence.

The multiple Sheffer polynomials \(\{S_{k_{1},k_{2}}(x)\}_{k_{1},k_{2}=0}^{\infty }\) are defined as follows. The generating function is

$$ A(t_{1},t_{2})e^{xH(t_{1},t_{2})}=\sum _{k_{1}=0}^{\infty }\sum_{k_{2}=0}^{ \infty }S_{k_{1},k_{2}}(x) \frac{t_{1}^{k_{1}}t_{2}^{k_{2}}}{k_{1}!k_{2}!}, $$
(1.2)

where \(A(t_{1},t_{2})\) and \(H(t_{1},t_{2})\) have series expansions of the form

$$ A(t_{1},t_{2})=\sum_{k_{1}=0}^{\infty } \sum_{k_{2}=0}^{\infty }a_{k_{1},k_{2}} \frac{t_{1}^{k_{1}}t_{2}^{k_{2}}}{k_{1}!k_{2}!} $$
(1.3)

and

$$ H(t_{1},t_{2})=\sum_{k_{1}=0}^{\infty } \sum_{k_{2}=0}^{\infty }h_{k_{1},k_{2}} \frac{t_{1}^{k_{1}}t_{2}^{k_{2}}}{k_{1}!k_{2}!}, $$
(1.4)

respectively, with the conditions

$$ A(0,0)=a_{0,0}\neq 0 \quad\text{and}\quad H(0,0)=h_{0,0}\neq 0. $$

In [1], one defined the positive linear operators involving multiple Sheffer polynomials for \(x\in [ 0,\infty )\) as follows:

$$ G_{n}(f,x)=\frac{e^{-\frac{nx}{2}H(1,1)}}{A(1,1)}\sum_{k_{1}=0}^{ \infty } \sum_{k_{2}=0}^{\infty } \frac{S_{k_{1},k_{2}} ( \frac{nx}{2} ) }{k_{1}!k_{2}!}f \biggl( \frac{k_{1}+k_{2}}{n} \biggr), $$
(1.5)

provided that the right-hand side of the above series converge, under conditions that:

  1. (1)

    \(S_{k_{1},k_{2}}(x)\geq 0,k_{1},k_{2}\in \mathbb{N}\),

  2. (2)

    \(A(1,1)\neq 0, H_{t_{1}}(1,1)=1,H_{t_{2}}(1,1)=1 \),

  3. (3)

    Series (1.2), (1.3) and (1.4) are convergent for \(|t_{1}|< R\), \(|t_{2}|< R\) and \((R_{1},R_{2})>1\).

In [6] one defined the Kantorovich variant of Szász operators induced by multiple Sheffer polynomials as follows:

$$ K_{n}^{(S)}(f,x)=\frac{ne^{-\frac{nx}{2}H(1,1)}}{A(1,1)}\sum _{k_{1}=0}^{\infty }\sum_{k_{2}=0}^{\infty } \frac{S_{k_{1},k_{2}} ( \frac{nx}{2} ) }{k_{1}!k_{2}!} \int _{\frac{k_{1}+k_{2}}{n}}^{ \frac{k_{1}+k_{2}+1}{n}}{f(t)}\,dt, \quad x\in [ 0,\infty ), $$
(1.6)

provided that the right-hand side of the above relation exists.

Example 1.1 of [6], gives us the following expressions for moments of the Kantorovich variant of Szász operators induced by multiple Sheffer polynomials:

$$\begin{aligned} &K_{n}^{(S)}(1,x) =1, \\ &K_{n}^{(S)}(t,x) =\frac{1}{2}\cdot \frac{2aa_{0,1}+2aa_{1,0}+aa_{0,0}}{aa_{0,0}n}+x, \\ &K_{n}^{(S)} \bigl(t^{2},x \bigr) = \frac{1}{3}\cdot \frac{3aa_{0,2}+3aa_{0,1}+3aa_{0,2}+6aa_{1,1}+3aa_{1,0}+aa_{0,0}}{aa_{0,0}n^{2}} \\ &\phantom{K_{n}^{(S)} (t^{2},x ) =}{} +\frac{x}{2}\cdot \frac{aa_{0,0}hh_{0,2}+2aa_{0,0}hh_{1,1}+aa_{0,0}hh_{2,0}+2aa_{0,0}+4aa_{0,1}+4aa_{1,0}}{aa_{0,0}n}+x^{2}. \end{aligned}$$

The central moments of the Kantorovich variant of Szász operators induced by multiple Sheffer polynomials are [6]

$$\begin{aligned} &K_{n}^{(S)}(t-x,x)=(2\tilde{a}_{0,1}+2 \tilde{a}_{1,0}+\tilde{a}_{0,0}) \frac{1}{2\tilde{a}_{0,0}n}, \\ &K_{n}^{(S)} \bigl((t-x)^{2},x \bigr)=(3 \tilde{a}_{0,2}+6\tilde{a}_{1,1}+3 \tilde{a}_{1,0}+3\tilde{a}_{2,0}+3\tilde{a}_{0,1}+\tilde{a}_{0,0}) \frac{1}{3\tilde{a}_{0,0}n^{2}} \\ &\phantom{K_{n}^{(S)} ((t-x)^{2},x )=}{}+(\tilde{h}_{0,2}+2\tilde{h}_{1,1}+ \tilde{h}_{2,0}) \frac{x}{2n}, \\ &K_{n}^{(S)} \bigl((t-x)^{3},x \bigr)\\ &\quad=(6 \tilde{a}_{0,2}+4\tilde{a}_{3,0}+6 \tilde{a}_{2,0}+4\tilde{a}_{1,0} +4\tilde{a}_{0,3}+12\tilde{a}_{2,1}+ \tilde{a}_{0,0}+4 \tilde{a}_{0,1}+12 \tilde{a}_{1,2} \\ &\qquad{}+12\tilde{a}_{1,1})\frac{1}{4\tilde{a}_{0,0}n^{3}}+(3 \tilde{a}_{0,0} \tilde{h}_{0,2}+2 \tilde{a}_{0,0}\tilde{h}_{0,3} +6\tilde{a}_{0,0} \tilde{h}_{1,1}+6\tilde{a}_{0,0}\tilde{h}_{1,2}+3 \tilde{a}_{0,0} \tilde{h}_{2,0} \\ &\qquad{}+6\tilde{a}_{0,0}\tilde{h}_{2,1}+2 \tilde{a}_{0,0} \tilde{h}_{3,0} +6 \tilde{a}_{0,1} \tilde{h}_{0,2}+12\tilde{a}_{0,1}\tilde{h}_{1,1} +6 \tilde{a}_{0,1}\tilde{h}_{2,0}+6\tilde{a}_{1,0} \tilde{h}_{0,2} \\ &\qquad{}+12\tilde{a}_{1,0}\tilde{h}_{1,1}+6 \tilde{a}_{1,0} \tilde{h}_{2,0}) \frac{x}{4\tilde{a}_{0,0}n^{2}}. \end{aligned}$$

Similarly, there exist constants \(C_{di}\) (dependent only on \(\tilde{a}_{i,j}\) and \(\tilde{h}_{i,j}\)) such that

$$\begin{aligned} &K_{n}^{(S)} \bigl((t-x)^{4},x \bigr)= \frac{C_{44}}{4\tilde{a}_{0,0}n^{4}} + \frac{xC_{43}}{2\tilde{a}_{0,0}n^{3}}+\frac{3x^{2}C_{42}}{4n^{2}}, \\ &K_{n}^{(S)} \bigl((t-x)^{5},x \bigr)= \frac{C_{55}}{6\tilde{a}_{0,0}n^{5}} + \frac{xC_{54}}{12\tilde{a}_{0,0}n^{4}}+\frac{5x^{2}C_{53}}{8\tilde{a}_{0,0}n^{3}}, \\ &K_{n}^{(S)} \bigl((t-x)^{6},x \bigr)= \frac{C_{66}}{7\tilde{a}_{0,0}n^{6}} + \frac{xC_{65}}{2\tilde{a}_{0,0}n^{5}}+\frac{5x^{2}C_{64}}{4\tilde{a}_{0,0}n^{4}}. + \frac{15x^{2}C_{63}}{8n^{3}}. \end{aligned}$$

As a consequence of the above relations, we obtain

$$\begin{aligned} &\lim_{n\to \infty } nK_{n}^{(S)}(t-x,x)= \frac{2\tilde{a}_{0,1}+2\tilde{a}_{1,0}+\tilde{a}_{0,0}}{2\tilde{a}_{0,0}}, \\ &\lim_{n\to \infty } nK_{n}^{(S)} \bigl((t-x)^{2},x \bigr)= \frac{(\tilde{h}_{0,2}+2\tilde{h}_{1,1}+\tilde{h}_{2,0})x}{2}, \\ &\lim_{n\to \infty } n^{2}K_{n}^{(S)} \bigl((t-x)^{3},x \bigr)=E_{3}x, \qquad\lim _{n \to \infty } n^{2}K_{n}^{(S)} \bigl((t-x)^{4},x \bigr)=E_{4}x^{2}, \\ &\lim_{n\to \infty } n^{3}K_{n}^{(S)} \bigl((t-x)^{5},x \bigr)=E_{5}x^{3}, \qquad\lim _{n \to \infty } n^{3}K_{n}^{(S)} \bigl((t-x)^{6},x \bigr)=E_{6}x^{3}, \end{aligned}$$

where \(E_{3},E_{4},E_{5},E_{6}\) are constant dependent on the derivatives of \(A(t_{1},t_{2})\) and \(H(t_{1},t_{2})\) up to order three at the point \((t_{1},t_{2})=(1,1)\).

2 Korovkin type results

The statistical form of Korovkin’s theorem was studied in [11] and the A-statistical version was considered in [8] (see also [13, 17] for other summability methods).

Let \(B[0,\infty )\) (\(C[0,\infty )\)) be “the space of all bounded (continuous) functions” on the interval \([0,\infty )\).

Theorem 2.1

Let \(\mathfrak{T}=(\mathfrak{d}_{ij})\) be regular matrix and \(K_{n}^{(S)}(f,x)\) be as in (1.6) on \([0,M]\), for any finite M. If

$$ st_{\mathfrak{T}}-\lim_{n}{ \bigl\Vert K_{n}^{(S)}(f,x)e_{i}-e_{i} \bigr\Vert }=0 \quad (i=1,2), $$

then

$$ st_{\mathfrak{T}}-\lim_{n}{ \bigl\Vert K_{n}^{(S)}(f,x)\mathfrak{h}- \mathfrak{h} \bigr\Vert }=0, $$

\(\mathfrak{h}\in C([0,M])\), where \(\Vert \mathfrak{h} \Vert =\sup_{t\in [ 0,M]}{ |\mathfrak{h}(t)|}\).

Proof

From Example 1.1 of [6], we have \(st_{\mathfrak{T}}-\lim_{n}{ \Vert K_{n}^{(S)}e_{0}-e_{0} \Vert }=0\). Now

$$ { \bigl\Vert K_{n}^{(S)}e_{1}-e_{1} \bigr\Vert }\leq \biggl\Vert \frac{1}{2}\cdot \frac{2aa_{0,1}+2aa_{1,0}+aa_{0,0}}{aa_{0,0}n} \biggr\Vert . $$

Also \(\lim_{n\rightarrow \infty } \Vert K_{n}^{(S)}e_{1}-e_{1} \Vert =0\). Moreover,

$$\begin{aligned} &{ \bigl\Vert K_{n}^{(S)}e_{2}-e_{2} \bigr\Vert }\\ &\quad= \biggl\Vert \frac{1}{3}\cdot \frac{3aa_{0,2}+3aa_{0,1}+3aa_{0,2}+6aa_{1,1}+3aa_{1,0}+aa_{0,0}}{aa_{0,0}n^{2}} \\ &\qquad{}+\frac{x}{2}\cdot \frac{aa_{0,0}hh_{0,2}+2aa_{0,0}hh_{1,1}+aa_{0,0}hh_{2,0}+2aa_{0,0}+4aa_{0,1}+4aa_{1,0}}{aa_{0,0}n} \biggr\Vert \rightarrow 0, \end{aligned}$$

as \(n\rightarrow \infty \). Now the proof follows directly from the statistical version of the Korovkin theorem [11]. □

Example 2.2

([14])

Under the conditions given in Theorem 2.1, set

$$ N_{n}(h,x)=(1+u_{n})K_{n}^{(S)}(h,x), $$

where

$$ u_{n}=\textstyle\begin{cases} 1; & m^{2}-m\leq n\leq m^{2}-1, \\ \frac{1}{m^{4}}; & n=m^{2};m\in \mathbb{N}\setminus \{1\}, \\ 0; & \text{otherwise},\end{cases} $$

then

$$\begin{aligned} &N_{n}(e_{0},x) =(1+u_{n}), \\ &N_{n}(e_{1},x) =(1+u_{n}) \biggl( \frac{1}{2}\cdot \frac{2aa_{0,1}+2aa_{1,0}+aa_{0,0}}{aa_{0,0}n}+x \biggr), \\ &N_{n}(e_{2},x)=(1+u_{n}) \biggl( \frac{1}{3}\cdot \frac{3aa_{0,2}+3aa_{0,1}+3aa_{0,2}+6aa_{1,1}+3aa_{1,0}+aa_{0,0}}{aa_{0,0}n^{2}} \\ &\phantom{N_{n}(e_{2},x)=}{}+ \frac{x}{2}\cdot \frac{aa_{0,0}hh_{0,2}+2aa_{0,0}hh_{1,1}+aa_{0,0}hh_{2,0}+2aa_{0,0}+4aa_{0,1}+4aa_{1,0}}{aa_{0,0}n}+x^{2} \biggr). \end{aligned}$$

If the matrix \(\mathfrak{T}\) is as in (1.1), then, by Theorem 2.1 we obtain \(st_{\mathfrak{T}}-\lim_{n}{ \Vert N_{n}h-h \Vert }=0\), but the operators \(N_{n}(h,x)\), do not satisfy the conditions of the theorem in [11].

In the following result we use a power series method as in [20, 24]; the Abel summability method was used.

Theorem 2.3

Let \((K_{n}^{(S)})\) be a sequence of positive linear operators from \(C[0,M]\) into \(B[0,M]\) \((0< M<\infty )\) such that

$$ \lim_{t\rightarrow R^{-}}{\frac{1}{\mathfrak{p}(t)} \Biggl\Vert \sum _{n=0}^{ \infty }{ \bigl(K_{n}^{(S)}e_{i}-e_{i} \bigr){\mathfrak{p}}_{n}t^{n}} \Biggr\Vert }=0, \quad i=0,1,2. $$
(2.1)

Then, for \(\mathfrak{h}\in C[0,M]\),

$$ \lim_{t\rightarrow R^{-}}{\frac{1}{\mathfrak{p}(t)} \Biggl\Vert \sum _{n=0}^{ \infty }{ \bigl(K_{n}^{(S)} \mathfrak{h}-\mathfrak{h} \bigr)\mathfrak{p}_{n}t^{n}} \Biggr\Vert }=0. $$
(2.2)

Proof

Clearly, from (2.2) follows (2.1). Now we show the converse that (2.1) implies (2.2). Let \(\mathfrak{h}\in C[0,M]\). Then there exists a constant \(K>0\) such that \(|\mathfrak{h}(u)|\leq K\) for all \(u\in [ 0,M]\). Therefore

$$ \bigl\vert \mathfrak{h}(u)-\mathfrak{h}(x) \bigr\vert \leq 2K, \quad u\in [ 0,M]. $$
(2.3)

For every given \(\epsilon >0\), there exists a \(\delta >0\) such that

$$ \bigl\vert \mathfrak{h}(u)-\mathfrak{h}(x) \bigr\vert \leq \epsilon $$
(2.4)

whenever \(|u-x|<\delta \) for all \(u\in [ 0,M]\). Define \(\psi \equiv \psi (u,x)=(u-x)^{2}\). If \(|u-x|\geq \delta \), then

$$ \bigl\vert \mathfrak{h}(u)-\mathfrak{h}(x) \bigr\vert \leq \frac{2K}{\delta ^{2}}\psi (u,x). $$
(2.5)

From (2.3)–(2.5), we have \(|\mathfrak{h}(u)-\mathfrak{h}(x)|<\epsilon +\frac{2K}{\delta ^{2}}\psi (u,x)\), namely,

$$ -\epsilon -\frac{2K}{\delta ^{2}}\psi (u,x)< \mathfrak{h}(t)- \mathfrak{h}(x)< \frac{2K}{\delta ^{2}}\psi (u,x)+\epsilon. $$

By applying the operator \(K_{n}^{(S)}(1,x)\), \(K_{n}^{(S)}(1,x)\) is a monotone and linear operator, we obtain

$$ K_{n}^{(S)}(1,x) \biggl( -\epsilon -\frac{2K}{\delta ^{2}}\psi \biggr) < K_{n}^{(S)}(1,x) \bigl( \mathfrak{h}(u)- \mathfrak{h}(x) \bigr) < K_{n}^{(S)}(1,x) \biggl( \frac{2K}{\delta ^{2}}\psi +\epsilon \biggr), $$

which implies

$$\begin{aligned} -\epsilon K_{n}^{(S)}(1,x)-\frac{2K}{\delta ^{2}}K_{n}^{(S)} \bigl(\psi (u),x \bigr)&< K_{n}^{(S)}( \mathfrak{h},x)- \mathfrak{h}(x)K_{n}^{(S)}(1,x) \\ &< \frac{2K}{\delta ^{2}}K_{n}^{(S)} \bigl(\psi (u),x \bigr)+\epsilon K_{n}^{(S)}(1,x). \end{aligned}$$
(2.6)

On the other hand

$$ K_{n}^{(S)}(\mathfrak{h},x)-\mathfrak{h}(x)=K_{n}^{(S)}( \mathfrak{h},x)-\mathfrak{h}(x)K_{n}^{(S)}(1,x)+ \mathfrak{h}(x) \bigl[K_{n}^{(S)}(1,x)-1 \bigr]. $$
(2.7)

From (2.6) and (2.7) we get

$$ K_{n}^{(S)}(\mathfrak{h},x)-\mathfrak{h}(x)< \frac{2K}{\delta ^{2}}K_{n}^{(S)} \bigl(\psi (u),x \bigr)+\epsilon K_{n}^{(S)}(1,x)+\mathfrak{h}(x) \bigl[K_{n}^{(S)}(1,x)-1 \bigr]. $$
(2.8)

Now we estimate the following expression:

$$\begin{aligned} K_{n}^{(S)} \bigl(\psi (u),x \bigr) &= K_{n}^{(S)} \bigl((x-u)^{2},x \bigr)=K_{n}^{(S)} \bigl( \bigl(x^{2}-2xu+u^{2} \bigr),x \bigr) \\ &= x^{2}K_{n}^{(S)}(1,x)-2xK_{n}^{(S)}(u,x)+K_{n}^{(S)} \bigl(u^{2},x \bigr). \end{aligned}$$

By (2.8), we obtain

$$\begin{aligned} K_{n}^{(S)}(\mathfrak{h},x)-\mathfrak{h}(x) < {}& \frac{2K}{\delta ^{2}} \bigl\{ x^{2} \bigl[K_{n}^{(S)}(1,x)-1 \bigr]-2x \bigl[K_{n}^{(S)}(u,x)-x \bigr] \\ &{}+ \bigl[K_{n}^{(S)} \bigl(u^{2},x \bigr)-x^{2} \bigr] \bigr\} +\epsilon K_{n}^{(S)}(1,x)+f(x) \bigl[K_{n}^{(S)}(1,x)-1 \bigr] \\ ={}&\epsilon +\epsilon \bigl[ K_{n}^{(S)}(1,x)-1 \bigr]+ \mathfrak{h}(x) \bigl[K_{n}^{(S)}(1,x)-1 \bigr] \\ &{}+\frac{2K}{\delta ^{2}} \bigl\{ x^{2} \bigl[K_{n}^{(S)}(1,x)-1 \bigr]-2x \bigl[K_{n}^{(S)}(u,x)-x \bigr]+ \bigl[K_{n}^{(S)} \bigl(u^{2},x \bigr)-x^{2} \bigr] \bigr\} . \end{aligned}$$

Therefore,

$$\begin{aligned} \bigl\vert K_{n}^{(S)}(\mathfrak{h},x)-\mathfrak{h}(x) \bigr\vert \leq {}&\epsilon + \biggl( \epsilon +K+\frac{2KM^{2}}{\delta ^{2}} \biggr) \bigl\vert K_{n}^{(S)}(1,x)-1 \bigr\vert \\ &{}+\frac{4KM}{\delta ^{2}} \bigl\vert K_{n}^{(S)}(u,x)-x \bigr\vert + \frac{2K}{\delta ^{2}} \bigl\vert K_{n}^{(S)} \bigl(u^{2},x \bigr)-x^{2} \bigr\vert . \end{aligned}$$

From the above relations and the linearity of \(K_{n}^{(S)}\), we obtain

$$\begin{aligned} &\frac{1}{\mathfrak{p}(v)} \Biggl\Vert \sum_{n=0}^{\infty }{ \bigl(U_{n,p}( \mathfrak{h};x)-\mathfrak{h}(x) \bigr) \mathfrak{p}_{n}v^{n}} \Biggr\Vert \\ &\quad\leq \epsilon + \biggl( \epsilon +K+\frac{2KM^{2}}{\delta ^{2}} \biggr) \frac{1}{\mathfrak{p}(t)} \Biggl\Vert \sum_{n=0}^{\infty }{ \bigl(K_{n}^{(S)}(1;x)-1 \bigr))\mathfrak{p}_{n}t^{n}} \Biggr\Vert \\ &\qquad{}+\frac{4KM}{\delta ^{2}}\frac{1}{\mathfrak{p}(v)} \Biggl\Vert \sum _{n=0}^{ \infty }{ \bigl(K_{n}^{(S)}(u;x)-x \bigr))\mathfrak{p}_{n}v^{n}} \Biggr\Vert +\frac{2K}{\delta ^{2}}\frac{1}{\mathfrak{p}(v)} \Biggl\Vert \sum _{n=0}^{ \infty }{ \bigl(K_{n}^{(S)} \bigl(u^{2};x \bigr)-x^{2} \bigr))\mathfrak{p}_{n}v^{n}} \Biggr\Vert . \end{aligned}$$

Hence, (2.2) follows from the last relation and (2.1). □

3 Rate of convergence

The modulus of continuity is defined by

$$ \omega (\mathfrak{h},\delta )=\sup_{ \vert h \vert < \delta }{ \bigl\vert \mathfrak{h}(x+h)-\mathfrak{h}(x) \bigr\vert }, \quad \mathfrak{h}(x) \in C[0,M]\cap E. $$

Note that

$$ \bigl\vert \mathfrak{h}(x)-\mathfrak{h}(y) \bigr\vert \leq \omega ( \mathfrak{h},\delta ) \biggl( \frac{ \vert x-y \vert }{\delta }+1 \biggr). $$
(3.1)

Theorem 3.1

Let \(\mathfrak{T}=(\mathfrak{a}_{ij})\) be regular and \(\mathfrak{h}\in C[0,M]\). If \((\alpha _{n})\) is a sequence of positive real numbers such that \(\omega (\mathfrak{h},\delta _{n})=st_{\mathfrak{T}}-0 ( {\alpha _{n}} ) \), then

$$ \bigl\Vert K_{n}^{(S)}\mathfrak{h}-\mathfrak{h} \bigr\Vert =st_{\mathfrak{T}}-0(\alpha _{n}), $$

where

$$\begin{aligned} \delta _{n} ={}& \biggl\{ \frac{1}{3}\cdot \biggl\Vert \frac{3aa_{0,2}+3aa_{0,1}+3aa_{0,2}+6aa_{1,1}+3aa_{1,0}+aa_{0,0}}{aa_{0,0}n^{2}} \biggr\Vert \\ &{} +M \biggl[ \biggl\Vert \frac{1}{2}\cdot \frac{aa_{0,0}hh_{0,2}+2aa_{0,0}hh_{1,1}+aa_{0,0}hh_{2,0}+2aa_{0,0}+4aa_{0,1}+4aa_{1,0}}{aa_{0,0}n} \biggr\Vert \\ &{} + \biggl\Vert \frac{2aa_{0,1}+2aa_{1,0}+aa_{0,0}}{aa_{0,0}n} \biggr\Vert \biggr] \biggr\} ^{2}, \end{aligned}$$

for any positive integer n.

Proof

By (3.1), we see

$$\begin{aligned} &\bigl\vert K_{n}^{(S)}(\mathfrak{h};x)-\mathfrak{h} \bigr\vert \\ &\quad \leq K_{n}^{(S)} \bigl( \bigl\vert \mathfrak{h}(t)-\mathfrak{h}(x) \bigr\vert ;x \bigr)\\ &\quad\leq \frac{ne^{-\frac{nx}{2}H(1,1)}}{A(1,1)}\sum_{k_{1}=0}^{\infty } \sum_{k_{2}=0}^{\infty } \frac{S_{k_{1},k_{2}} ( \frac{nx}{2} ) }{k_{1}!k_{2}!} \int _{ \frac{k_{1}+k_{2}}{n}}^{\frac{k_{1}+k_{2}+1}{n}}{\omega (\mathfrak{h},\delta ) \biggl( 1+ \frac{ \vert t-x \vert }{\delta } \biggr) }\,dt \\ &\quad \leq \omega (\mathfrak{h},\delta ) \Biggl[ 1+\frac{1}{\delta } \frac{ne^{-\frac{nx}{2}H(1,1)}}{A(1,1)}\sum_{k_{1}=0}^{\infty }\sum _{k_{2}=0}^{ \infty }\frac{S_{k_{1},k_{2}} ( \frac{nx}{2} ) }{k_{1}!k_{2}!} \int _{\frac{k_{1}+k_{2}}{n}}^{\frac{k_{1}+k_{2}+1}{n}} \bigl( \vert t-x \vert \bigr)\,dt \Biggr],\quad \text{see [6]} \\ &\quad =\omega (\mathfrak{h},\delta ) \biggl[ 1+\frac{1}{\delta }K_{n}^{(S)} \bigl( \vert t-x \vert ;x \bigr) \biggr]. \end{aligned}$$

By applying the Cauchy–Schwartz inequality, we have

$$ \bigl\vert K_{n}^{(S)}(\mathfrak{h};x)-\mathfrak{h} \bigr\vert \leq \omega (\mathfrak{h}, \delta ) \biggl[ 1+ \frac{1}{\delta } \bigl( K_{n}^{(S)} \bigl( \vert t-x \vert ^{2};x \bigr) \bigr) ^{ \frac{1}{2}} \biggr]. $$

From Example 1.1 of [6], we obtain

$$\begin{aligned} &K_{n}^{(S)} \bigl((u-x)^{2};x \bigr)\\ &\quad=K_{n}^{(S)}(e_{2};x)-2xK_{n}^{(S)}(e_{1};x)+x^{2}K_{n}^{(S)}(e_{0};x) \\ &\quad\leq \frac{1}{3}\cdot \biggl\Vert \frac{3aa_{0,2}+3aa_{0,1}+3aa_{0,2}+6aa_{1,1}+3aa_{1,0}+aa_{0,0}}{aa_{0,0}n^{2}} \biggr\Vert \\ &\qquad{}+ M \biggl[ \biggl\Vert \frac{1}{2}\cdot \frac{aa_{0,0}hh_{0,2}+2aa_{0,0}hh_{1,1}+aa_{0,0}hh_{2,0}+2aa_{0,0}+4aa_{0,1}+4aa_{1,0}}{aa_{0,0}n} \biggr\Vert \\ &\qquad{}+ \biggl\Vert \frac{2aa_{0,1}+2aa_{1,0}+aa_{0,0}}{aa_{0,0}n} \biggr\Vert \biggr]. \end{aligned}$$

By taking

$$\begin{aligned} \delta _{n} ={}& \biggl\{ \frac{1}{3}\cdot \biggl\Vert \frac{3aa_{0,2}+3aa_{0,1}+3aa_{0,2}+6aa_{1,1}+3aa_{1,0}+aa_{0,0}}{aa_{0,0}n^{2}} \biggr\Vert \\ &{}+ M \biggl[ \biggl\Vert \frac{1}{2}\cdot \frac{aa_{0,0}hh_{0,2}+2aa_{0,0}hh_{1,1}+aa_{0,0}hh_{2,0}+2aa_{0,0}+4aa_{0,1}+4aa_{1,0}}{aa_{0,0}n} \biggr\Vert \\ &{} + \biggl\Vert \frac{2aa_{0,1}+2aa_{1,0}+aa_{0,0}}{aa_{0,0}n} \biggr\Vert \biggr] \biggr\} ^{2}, \end{aligned}$$

we get \(\Vert K_{n}^{(S)}\mathfrak{h}-\mathfrak{h} \Vert \leq 2\cdot \omega (\mathfrak{h},\delta _{n})\). Therefore, for every \(\epsilon >0\), we have

$$ \frac{1}{\alpha _{n}}\sum_{ \Vert K_{n}^{(S)}\mathfrak{h}-\mathfrak{h} \Vert \geq \epsilon }\mathfrak{t} {_{nj}}\leq \frac{1}{\alpha _{n}}\sum_{2 \cdot \omega (f,\delta _{n})\geq {\epsilon }} \mathfrak{t} {_{nj}}. $$

From the conditions that are given in the theorem, we have \(\Vert K_{n}^{(S)}\mathfrak{h}-\mathfrak{h} \Vert =st_{\mathfrak{T}}-0(\alpha _{i})\), as claimed. □

Now, we obtain the rate of convergence for our method.

Theorem 3.2

Let \(\mathfrak{h}\in C[0,M]\) and let ϕ be a positive real function defined on \((0,M)\). If \(\omega (\mathfrak{h},\psi (u))=O(\phi (u))\), as \(v\rightarrow R^{-}\), then we have

$$ {\frac{1}{\mathfrak{p}(v)} \Biggl\Vert \sum_{n=0}^{\infty }{ \bigl(K_{n}^{(S)}e_{i}-e_{i} \bigr) \mathfrak{p}_{n}v^{n}} \Biggr\Vert }=O \bigl(\phi (v) \bigr), $$

where the function \(\psi:(0,M)\rightarrow \mathbb{R}\) is defined by the relation

$$ \psi (u)= \Bigl\{ \sup_{\substack{ x\in (0,M) \\ n\in \mathbb{N}}} \bigl\{ K_{n}^{(S)} \bigl((u-x)^{2};x \bigr) \bigr\} \Bigr\} ^{\frac{1}{2}}. $$

Proof

For any \(u\in (0,R)\), \(x\in (0,M)\) and \(\delta >0\), we have

$$\begin{aligned} & \Biggl\vert \sum_{n=0}^{\infty }{ \bigl[K_{n}^{(S)}(\mathfrak{h};x)- \mathfrak{h}(x) \bigr]\mathfrak{p}_{n}v^{n}} \Biggr\vert \\ &\quad\leq \sum_{n=0}^{\infty }{K_{n}^{(S)} \bigl( \bigl\vert \mathfrak{h}(u)-\mathfrak{h}(x) \bigr\vert ;x \bigr)\mathfrak{p}_{n}v^{n}} \\ & \quad\leq \sum_{n=0}^{\infty }{K_{n}^{(S)} \biggl( \omega \biggl( \mathfrak{h},\frac{ \vert u-x \vert }{\delta }\delta \biggr);x \biggr) \mathfrak{p}_{n}v^{n}} \leq \sum _{n=0}^{\infty }{K_{n}^{(S)} \biggl( \biggl( 1+ \biggl[ \biggl\vert \frac{u-x}{\delta } \biggr\vert \biggr] \biggr) \omega (\mathfrak{h},\delta );x \biggr) \mathfrak{p}_{n}v^{n}} \\ &\quad \leq \omega (\mathfrak{h},\delta )\sum_{n=0}^{\infty }{K_{n}^{(S)} \biggl( 1+\frac{(u-x)^{2}}{\delta ^{2}};x \biggr) \mathfrak{p}_{n}v^{n}} \\ &\quad\leq \omega (\mathfrak{h},\delta )\sum_{n=0}^{\infty }{K_{n}^{(S)} \bigl(e_{0}(u);x \bigr) \mathfrak{p}_{n}v^{n}} +\frac{\omega (\mathfrak{h},\delta )}{\delta ^{2}}\sum_{n=0}^{ \infty }{K_{n}^{(S)} \bigl((u-x)^{2};x \bigr)\mathfrak{p}_{n}v^{n}}\\ &\quad=p(v) \omega ( \mathfrak{h},\delta )+\frac{\omega (\mathfrak{h},\delta )}{\delta ^{2}}\sup _{ \substack{ 0\leq x\leq 1 \\ n\in \mathbb{N}}} \bigl\{ K_{n}^{(S)} \bigl((u-x)^{2};x \bigr) \bigr\} \sum_{n=0}^{\infty } \mathfrak{p} {_{n}v^{n}}, \end{aligned}$$

which leads to

$$ \Biggl\vert \sum_{n=0}^{\infty }{ \bigl[K_{n}^{(S)}(f;x)-f(x) \bigr]\mathfrak{p}_{n}v^{n}} \Biggr\vert \leq \mathfrak{p}(v)\omega (f,\delta )+ \frac{\omega (f,\delta )}{\delta ^{2}}\sup _{0\leq x\leq 1} \bigl\{ K_{n}^{(S)} \bigl((u-x)^{2};x \bigr) \bigr\} \mathfrak{p}(v). $$

If we set \(\delta =\psi (u)\), then from the last inequality we have

$$ 0\leq \frac{1}{\mathfrak{p}(v)} \Biggl\Vert \sum_{n=0}^{\infty }{ \bigl(K_{n}^{(S)}\mathfrak{h}-\mathfrak{h} \bigr) \mathfrak{p}_{n}v^{n}} \Biggr\Vert \leq 2 \omega (\mathfrak{h},\delta ), $$

as required. □

4 Voronovskaya type theorems

It is well known that there is a Voronovskaya type theorem for the Kantorovich type generalization of Szász operators involving Sheffer type polynomials and it is stated as follows.

Theorem 4.1

([6])

For \(f\in C_{B}[0,\infty )\),

$$ \lim_{n\to \infty }n \bigl[K_{n}^{(S)} \bigl(f(t),x \bigr)-f(x) \bigr]=f^{{\prime }}(x) \biggl[ \frac{2\tilde{a}_{0,1}+2\tilde{a}_{1,0}+\tilde{a}_{0,0}}{2\tilde{a}_{0,0}} \biggr]+ \frac{f^{{\prime \prime }}(x)}{2} \biggl[ (\tilde{h}_{0,2}+2 \tilde{h}_{1,1}+\tilde{h}_{2,0})\frac{x}{2} \biggr], $$

for every \(x\in [0,M]\) and any finite M.

We extend the Voronovskaya type theorem for the \(\mathfrak{T}\)-statistical method for these operators. Let us consider the following operators.

Example 4.2

Define the operators

$$ NB_{n}(h,x)=(1+u_{n})K_{n}^{(S)}(h,x), $$

where

$$ u_{n}=\textstyle\begin{cases} \frac{1}{m^{3}} & m^{2}-m\leq n\leq m^{2}-1, \\ \frac{1}{m^{4}} & n=m^{2};m\in \mathbb{N}\setminus \{1\}, \\ 0 & \text{otherwise}.\end{cases} $$

Lemma 4.3

Let \(\mathfrak{h}\in C[0,M]\) such that \(\mathfrak{h}^{{\prime }},\mathfrak{h}^{{\prime \prime }}\in C[0,M]\), \(x\in [ 0,M]\). Then we obtain

$$ n^{2}NB_{n}^{(S)} \bigl((y-x)^{4};x \bigr)\sim E_{4}x^{2}(st_{\mathfrak{T}}) \quad\textit{on } [ 0,M]. $$

Proof

It follows directly from Remark 2.6 given in [6]. □

Theorem 4.4

Let \(\mathfrak{h}\in C[0,M]\) such that \(\mathfrak{h}^{{\prime }},\mathfrak{h}^{{\prime \prime }}\in C[0,M]\), \(x\in [ 0,M]\), for any finite M. Then

$$ n \bigl[NB_{n}^{(S)}(\mathfrak{h};x)-\mathfrak{h}(x) \bigr] \sim \mathfrak{h}^{{\prime }}(x) \biggl[ \frac{2\tilde{a}_{0,1}+2\tilde{a}_{1,0}+\tilde{a}_{0,0}}{2\tilde{a}_{0,0}} \biggr] + \frac{\mathfrak{h}^{{\prime \prime }}(x)}{2} \biggl(\frac{x(\tilde{h}_{0,2}+2\tilde{h}_{1,1}+\tilde{h}_{2,0})}{2} \biggr) (st_{T}), $$

on \([ 0,M]\).

Proof

Taylor’s formula gives

$$ \mathfrak{h}(y)=\mathfrak{h}(x)+(y-x)\mathfrak{h}^{{\prime }}(x)+ \frac{1}{2}(y-x)^{2}\mathfrak{h}^{{\prime \prime }}(x)+(y-x)^{2} \psi (y-x), $$
(4.1)

where \(\psi (y-x)\rightarrow 0\), as \(y-x\rightarrow 0\). After applying \(NB_{n}^{(S)}\) on both sides of Eq. (4.1), we obtain

$$\begin{aligned} NB_{n}^{(S)}(\mathfrak{h}) ={}&(1+u_{n}) \mathfrak{h}(x)+(1+u_{n}) \mathfrak{h}^{{\prime }}(x) \biggl( (2\tilde{a}_{0,1}+2\tilde{a}_{1,0}+\tilde{a}_{0,0})\frac{1}{2\tilde{a}_{0,0}n} \biggr) \\ &{} +(1+u_{n})\frac{\mathfrak{h}^{{\prime \prime }}(x)}{2} \biggl((3 \tilde{a}_{0,2}+6 \tilde{a}_{1,1}+3\tilde{a}_{1,0}+3\tilde{a}_{2,0}+3 \tilde{a}_{0,1}+\tilde{a}_{0,0})\frac{1}{3\tilde{a}_{0,0}n^{2}} \\ & {}+(\tilde{h}_{0,2}+2\tilde{h}_{1,1}+\tilde{h}_{2,0}) \frac{x}{2n} \biggr)+(1+x_{n})NB_{n}^{(S)} \bigl(\Phi ^{2}\psi (y-x);x \bigr). \end{aligned}$$

This yields

$$\begin{aligned} nNB_{n}^{(S)}(\mathfrak{h})={}& n(1+u_{n}) \mathfrak{h}(x)+n(1+u_{n}) \mathfrak{h}^{{\prime }}(x) \biggl( (2\tilde{a}_{0,1}+2\tilde{a}_{1,0}+\tilde{a}_{0,0})\frac{1}{2\tilde{a}_{0,0}n} \biggr) \\ &{} +n(1+u_{n})\frac{\mathfrak{h}^{{\prime \prime }}(x)}{2} \biggl((3 \tilde{a}_{0,2}+6 \tilde{a}_{1,1}+3\tilde{a}_{1,0}+3\tilde{a}_{2,0}+3 \tilde{a}_{0,1}+\tilde{a}_{0,0})\frac{1}{3\tilde{a}_{0,0}n^{2}} \\ &{} +(\tilde{h}_{0,2}+2\tilde{h}_{1,1}+\tilde{h}_{2,0}) \frac{x}{2n} \biggr)+n(1+u_{n})NB_{n}^{(S)} \bigl(\Phi ^{2}\psi (y-x);x \bigr). \end{aligned}$$

Therefore,

$$\begin{aligned} & \biggl|n [NB_{n}^{(S)}(\mathfrak{h};x)-\mathfrak{h}(x)- \mathfrak{h}^{{\prime }}(x) \biggl[ (2\tilde{a}_{0,1}+2 \tilde{a}_{1,0}+\tilde{a}_{0,0})\frac{1}{2\tilde{a}_{0,0}} \biggr] \\ &\qquad{} -\frac{\mathfrak{h}^{{\prime \prime }}(x)}{2} \biggl((\tilde{h}_{0,2}+2\tilde{h}_{1,1}+\tilde{h}_{2,0})\frac{x}{2} \biggr)\biggr| \\ &\quad \leq nKu_{n}+nK_{1}u_{n} \biggl\vert \frac{2\tilde{a}_{0,1}+2\tilde{a}_{1,0}+\tilde{a}_{0,0}}{2\tilde{a}_{0,0}n} \biggr\vert \\ &\qquad{} +n\frac{K_{2}}{2} \biggl\vert \frac{3\tilde{a}_{0,2}+6\tilde{a}_{1,1}+3\tilde{a}_{1,0}+3\tilde{a}_{2,0}+3\tilde{a}_{0,1}+\tilde{a}_{0,0}}{3\tilde{a}_{0,0}n^{2}} \biggr\vert \\ &\qquad{} +nu_{n}\frac{K_{2}}{2} \biggl\vert \frac{3\tilde{a}_{0,2}+6\tilde{a}_{1,1}+3\tilde{a}_{1,0}+3\tilde{a}_{2,0}+3\tilde{a}_{0,1}+\tilde{a}_{0,0}}{3\tilde{a}_{0,0}n^{2}}+( \tilde{h}_{0,2}+2\tilde{h}_{1,1}+ \tilde{h}_{2,0}) \frac{x}{2n} \biggr\vert \\ & \qquad{} +n \bigl\vert NB_{n}^{(S)} \bigl((y-x)^{2} \psi (y-x);x \bigr) \bigr\vert +u_{n}n \bigl\vert NB_{n}^{(S)} \bigl((y-x)^{2} \psi (y-x);x \bigr) \bigr\vert , \end{aligned}$$

where \(K=\sup_{x\in [ 0,M]}{|\mathfrak{h}(x)|}\), \(K_{1}=\sup_{x\in [ 0,M]}{|\mathfrak{h}^{{\prime }}(x)|}\) and \(K_{2}=\sup_{x\in [ 0,M]}{|\mathfrak{h}^{{\prime \prime }}(x)|}\).

Now we have to prove that

$$ \lim_{n\rightarrow \infty }{n \bigl\vert NB_{n}^{(S)} \bigl((y-x)^{2}\psi (y-x);x \bigr) \bigr\vert }=0. $$

By applying the Cauchy–Schwartz inequality, we obtain

$$ n \bigl\vert NB_{n}^{(S)} \bigl((y-x)^{2}\psi (y-x);x \bigr) \bigr\vert \leq \bigl[ n^{2}NB_{n}^{(S)} \bigl((y-x)^{4};x \bigr) \bigr] ^{\frac{1}{2}}\cdot \bigl[ NB_{n}^{(S)} \bigl(\psi ^{2};x \bigr) \bigr]^{ \frac{1}{2}}. $$
(4.2)

Also, by setting \(\eta _{x}(y)=(\psi (y-x))^{2}\), we have \(\eta _{x}(x)=0\) and \(\eta _{x}(\cdot )\in C[0,M]\). So

$$ NB_{n}^{(S)}(\eta _{x})\rightarrow 0(st_{\mathfrak{T}}) \quad\text{on } [ 0,M]. $$
(4.3)

Now from the previous relation, (4.2), (4.3), and Lemma 4.3, we obtain

$$ n^{2}NB_{n}^{(S)} \bigl((y-x)^{2}\psi (y-x);x \bigr)\rightarrow 0(st_{ \mathfrak{T}}) \quad\text{on } [ 0,M]. $$
(4.4)

From the construction of \((u_{n})\), it follows that \(nu_{n}\rightarrow 0(st_{\mathfrak{T}})\) on \([0,M]\).

For a given \(\epsilon >0\), we define the sets

$$\begin{aligned} A ={}&\biggl| \{n:|n [NB_{n}^{(S)}(\mathfrak{h};x)- \mathfrak{h}(x)-\mathfrak{h}^{{\prime }}(x) \biggl[ (2 \tilde{a}_{0,1}+2\tilde{a}_{1,0}+ \tilde{a}_{0,0}) \frac{1}{2\tilde{a}_{0,0}} \biggr] \\ & {}-\frac{\mathfrak{h}^{{\prime \prime }}(x)}{2} \biggl((\tilde{h}_{0,2}+2\tilde{h}_{1,1}+\tilde{h}_{2,0})\frac{x}{2} \biggr)\biggr|, \\ A_{1} = {}& \biggl\vert \biggl\{ n: \vert nu_{n} \vert \geq \frac{\epsilon }{3K} \biggr\} \biggr\vert , \\ A_{2} ={}& \biggl\vert \biggl\{ n: \bigl\vert nNB_{n}^{(S)} \bigl((y-x)^{2}\psi (y-x);x \bigr) \bigr\vert \geq \frac{\epsilon }{3} \biggr\} \biggr\vert , \\ A_{3} ={}& \biggl\vert \biggl\{ n: \bigl\vert nu_{n}NB_{n}^{(S)} \bigl((y-x)^{2}\psi (y-x);x \bigr) \bigr\vert \geq \frac{\epsilon }{3} \biggr\} \biggr\vert . \end{aligned}$$

From these relations we obtain \(A\leq A_{1}+A_{2}+A_{3}\). Hence the result follows. □

Theorem 4.5

Let \(\mathfrak{h},\mathfrak{h}^{{\prime }},\mathfrak{h}^{{\prime \prime }}\in C[0,\infty )\). Then

$$\begin{aligned} & \biggl\vert n \bigl( K_{n}^{(S)}(\mathfrak{h},x)- \mathfrak{h}(x) \bigr) -\mathfrak{h}^{{\prime }}(x) \biggl( (2 \tilde{a}_{0,1}+2\tilde{a}_{1,0}+ \tilde{a}_{0,0}) \frac{1}{2\tilde{a}_{0,0}n} \biggr) \\ &\qquad{} -\frac{\mathfrak{h}^{{\prime \prime }}(x)}{2}\cdot \biggl[(3 \tilde{a}_{0,2}+6 \tilde{a}_{1,1}+3\tilde{a}_{1,0}+3\tilde{a}_{2,0}+3 \tilde{a}_{0,1}+\tilde{a}_{0,0})\frac{1}{3\tilde{a}_{0,0}n^{2}} \\ &\qquad{} +(\tilde{h}_{0,2}+2\tilde{h}_{1,1}+ \tilde{h}_{2,0}) \frac{x}{2n} \biggr] \biggr\vert \\ &\quad=0(1)\omega \bigl( \mathfrak{h}^{{\prime \prime }},n^{-\frac{1}{2}} \bigr), \end{aligned}$$

as \(n\rightarrow \infty \), and for every \(x\in [ 0,M]\), for any finite M.

Proof

From Taylor’s theorem, we have

$$ \mathfrak{h}(u)=\mathfrak{h}(x)+\mathfrak{h}^{{\prime }}(x) (u-x)+ \frac{\mathfrak{h}^{{\prime \prime }}(x)}{2}(u-x)^{2}+R(u,x), $$

where \(R(u,x)= \frac{\mathfrak{h}^{{\prime \prime }}(\theta )-\mathfrak{h}^{{\prime \prime }}(x)}{2}(u-x)^{2}\), for \(\theta \in (u,x)\). Now we obtain

$$ \biggl\vert K_{n}^{(S)}(\mathfrak{h},x)-\mathfrak{h}(x)- \mathfrak{h}^{{\prime }}(x)K_{n}^{(S)} \bigl((u-x);x \bigr)- \frac{\mathfrak{h}^{{\prime \prime }}(x)}{2}K_{n}^{(S)} \bigl((u-x)^{2};x \bigr) \biggr\vert \leq K_{n}^{(S)} \bigl( \bigl\vert R(u,x) \bigr\vert ,x \bigr). $$

From this we get

$$\begin{aligned} & \biggl|n \bigl( K_{n}^{(S)}(\mathfrak{h},x)-\mathfrak{h}(x) \bigr) -\mathfrak{h}^{{\prime }}(x) \biggl( \frac{2\tilde{a}_{0,1}+2\tilde{a}_{1,0}+\tilde{a}_{0,0}}{2\tilde{a}_{0,0}} \biggr) \\ &\qquad{}- \frac{\mathfrak{h}^{{\prime \prime }}(x)}{2}\cdot [(3 \tilde{a}_{0,2}+6 \tilde{a}_{1,1}+3\tilde{a}_{1,0}+3\tilde{a}_{2,0}+3 \tilde{a}_{0,1}+\tilde{a}_{0,0})\frac{1}{3\tilde{a}_{0,0}n}\\ &\qquad{}+( \tilde{h}_{0,2}+2 \tilde{h}_{1,1}+ \tilde{h}_{2,0})\frac{x}{2}\biggr| \\ &\quad \leq n\cdot K_{n}^{(S)} \bigl( \bigl\vert R(u,x) \bigr\vert ,x \bigr). \end{aligned}$$

By the properties of the continuity modulus, we have

$$ \biggl\vert \frac{\mathfrak{h}^{{\prime \prime }}(\theta )-\mathfrak{h}^{{\prime \prime }}(x)}{2!} \biggr\vert \leq \frac{1}{2!} \biggl( 1+ \frac{ \vert \theta -x \vert }{\delta } \biggr) \omega \bigl(\mathfrak{h}^{{\prime \prime }}, \delta \bigr). $$

On the other hand

$$ \biggl\vert \frac{\mathfrak{h}^{{\prime \prime }}(\theta )-\mathfrak{h}^{{\prime \prime }}(x)}{2!} \biggr\vert \leq\textstyle\begin{cases} \omega (\mathfrak{h}^{{\prime \prime },\delta }); & \vert u-x \vert \leq \delta, \\ \frac{(t-x)^{4}}{\delta ^{4}}\omega (\mathfrak{h}^{{\prime \prime }}, \delta ); & \vert u-x \vert \geq \delta.\end{cases} $$

For \(0<\delta <1\), we obtain

$$ \biggl\vert \frac{\mathfrak{h}^{{\prime \prime }}(\theta )-\mathfrak{h}^{{\prime \prime }}(x)}{2!} \biggr\vert \leq \omega \bigl( \mathfrak{h}^{{ \prime \prime }},\delta \bigr) \biggl( 1+\frac{(u-x)^{4}}{\delta ^{4}} \biggr), $$

which gives

$$ \bigl\vert R(u,x) \bigr\vert \leq \omega \bigl(\mathfrak{h}^{{\prime \prime }}, \delta \bigr) \biggl( 1+ \frac{(u-x)^{4}}{\delta ^{4}} \biggr) (u-x)^{2}=\omega \bigl(\mathfrak{h}^{{ \prime \prime }},\delta \bigr) \biggl( (u-x)^{2}+ \frac{(u-x)^{6}}{\delta ^{4}} \biggr). $$

By the linearity of \(K_{n}^{(S)}\) and the above relation we obtain

$$ K_{n}^{(S)} \bigl( \bigl\vert R(u,x) \bigr\vert ,x \bigr) \leq \omega \bigl(\mathfrak{h}^{{\prime \prime }}, \delta \bigr) \biggl( K_{n}^{(S)} \bigl((u-x)^{2},x \bigr)+ \frac{1}{\delta ^{4}}K_{n}^{(S)} \bigl((u-x)^{6},x \bigr) \biggr). $$

Taking into consideration Remark 2.6 in [6], for every \(x\in [0,M]\), we have

$$ K_{n}^{(S)} \bigl( \bigl\vert R(u,x) \bigr\vert ,x \bigr) \leq \omega \bigl(\mathfrak{h}^{{\prime \prime }}, \delta \bigr) \biggl( O \biggl( \frac{1}{n} \biggr) +\frac{1}{\delta ^{4}}O \biggl( \frac{1}{n^{3}} \biggr) \biggr) =O \biggl( \frac{1}{n} \biggr) \omega \bigl( \mathfrak{h}^{{\prime \prime }}, \delta \bigr). $$

For \(\delta =n^{-\frac{1}{2}}\), we complete the proof. □

Availability of data and materials

No data were used to support this study.

References

  1. Ali, M., Paris, R.B.: Generalization of Szász operators involving multiple Sheffer polynomials (2020). arXiv:2006.11131v1 [math.CA]

  2. Boos, J.: Classical and Modern Methods in Summability. Oxford University Press, Oxford (2000)

    MATH  Google Scholar 

  3. Braha, N., Mansour, T., Mursaleen, M., Acar, T.: Convergence of λ-Bernstein operators via power series summability method. J. Appl. Math. Comput. https://doi.org/10.1007/s12190-020-01384-x

  4. Braha, N.L.: Some properties of new modified Szász–Mirakyan operators in polynomial weight spaces via power summability method. Bull. Math. Anal. Appl. 10(3), 53–65 (2018)

    MathSciNet  MATH  Google Scholar 

  5. Braha, N.L.: Some properties of Baskakov–Schurer–Szász operators via power summability method. Quaest. Math. 42(10), 1411–1426 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  6. Braha, N.L., Mansour, T., Mursaleen, M., Acar, T.: Convergence of λ-Bernstein operators via power series summability method. J. Appl. Math. Comput. 65(1–2), 125–146 (2021)

    Article  MathSciNet  Google Scholar 

  7. Braha, N.L., Mansour, T., Mursaleen, M.: Some properties of Kantorovich–Stancu type generalization of Szász operators including Brenke type polynomials via power series summability method. J. Funct. Spaces 2020, Article ID 3480607 (2020)

    MATH  Google Scholar 

  8. Duman, O., Khan, M.K., Orhan, C.: A-statistical convergence of approximating operators. Math. Inequal. Appl. 6, 689–699 (2003)

    MathSciNet  MATH  Google Scholar 

  9. Fast, H.: Sur la convergence statistique. Colloq. Math. 2, 241–244 (1951)

    Article  MathSciNet  MATH  Google Scholar 

  10. Fridy, J.A., Miller, H.I.: A matrix characterization of statistical convergence. Analysis 11, 59–66 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  11. Gadjiev, A.D., Orhan, C.: Some approximation theorems via statistical convergence. Rocky Mt. J. Math. 32(1), 129–138 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  12. Kratz, W., Stadtmuller, U.: Tauberian theorems for \(J_{p}\)-summability. J. Math. Anal. Appl. 139, 362–371 (1989)

    Article  MathSciNet  Google Scholar 

  13. Mohiuddine, S.A., Alotaibi, A., Mursaleen, M.: Statistical summability \((C,1)\) and a Korovkin type approximation theorem. J. Inequal. Appl. 2012, 172 (2012)

    Article  MathSciNet  Google Scholar 

  14. Moricz, F., Orhan, C.: Tauberian conditions under which statistical convergence follows from statistical summability by weighted means. Studia Sci. Math. Hung. 41(4), 391–403 (2004)

    MathSciNet  MATH  Google Scholar 

  15. Mursaleen, M., Alotaibi, A.: Statistical summability and approximation by de la Vallée–Poussin mean. Appl. Math. Lett. 24, 320–324 (2011). [Erratum: Appl. Math. Lett. 25, 665 (2012)]

    Article  MathSciNet  MATH  Google Scholar 

  16. Mursaleen, M., Alotaibi, A.: Korovkin type approximation theorem for functions of two variables through statistical A-summability. Adv. Differ. Equ. 2012, 65 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  17. Mursaleen, M., Karakaya, V., Erturk, M., Gursoy, F.: Weighted statistical convergence and its application to Korovkin type approximation theorem. Appl. Math. Comput. 218, 9132–9137 (2012)

    MathSciNet  MATH  Google Scholar 

  18. Mursaleen, M., Kiliçman, A.: Korovkin second theorem via B-statistical A-summability. Abstr. Appl. Anal. 2013, Article ID 598963 (2013). https://doi.org/10.1155/2013/598963

    Article  MathSciNet  MATH  Google Scholar 

  19. Mursaleen, M., Mohiuddine, S.A.: Korovkin type approximation theorem for functions of two variables via statistical summability \((C,1)\). Acta Sci., Technol. 37(2), 237–243 (2015)

    Article  Google Scholar 

  20. Soylemez, D., Unver, M.: Korovkin type theorems for Cheney–Sharma operators via summability methods. Results Math. 72(3), 1601–1612 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  21. Stadtmuller, U., Tali, A.: On certain families of generalized Nörlund methods and power series methods. J. Math. Anal. Appl. 238, 44–66 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  22. Tas, E.: Some results concerning Mastroianni operators by power series method. Commun. Fac. Sci. Univ. Ank. Sér. A1 Math. Stat. 63(1), 187–195 (2016)

    MathSciNet  MATH  Google Scholar 

  23. Tas, E., Yurdakadim, T.: Approximation by positive linear operators in modular spaces by power series method. Positivity 21(4), 1293–1306 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  24. Unver, M.: Abel transforms of positive linear operators. In: ICNAAM 2013. AIP Conference Proceedings, vol. 1558, pp. 1148–1151 (2013)

    Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed equally. All authors read and approved the final manuscript.

Corresponding author

Correspondence to M. Mursaleen.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Loku, V., Braha, N.L., Mansour, T. et al. Approximation by a power series summability method of Kantorovich type Szász operators including Sheffer polynomials. Adv Differ Equ 2021, 165 (2021). https://doi.org/10.1186/s13662-021-03326-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13662-021-03326-8

MSC

Keywords