1 Introduction

In this note, we work with two fixed real parameters α and β satisfying \(\alpha \geq \beta\geq-1/2\). We use the following notations:

$$ \varrho^{ \alpha,\beta}(x) = (1 - x)^{\alpha}(1 + x)^{\beta}, \quad x \in(- 1, 1), $$
(1)

and, for \(1 \leq p < \infty\),

$$L^{p}_{(\alpha,\beta)} = \biggl\{ f:[-1,1]\to\mathbb{R}: \Vert f \Vert _{p} = \biggl( \int_{-1}^{1}\bigl| f(x)\bigr|^{p} \varrho^{ \alpha,\beta}(x)\,dx \biggr)^{1/p} < \infty \biggr\} . $$

Moreover, for each \(n\in\mathbb{N}_{0}\), \(\mathbb{P}_{n}\) is the family of all algebraic polynomials of degree not greater than n,

$$ w^{\alpha,\beta}_{n} = \frac{(2n + \alpha +\beta + 1)\Gamma(n + \alpha+\beta+ 1)\Gamma(n + \alpha + 1)}{\Gamma(n +\beta + 1) \Gamma(n + 1)(\Gamma(\alpha+ 1))^{2}} $$
(2)

(Γ stands for the gamma function) and

$$ \lambda_{n}=n(n+\alpha+\beta+1). $$
(3)

Since α and β are fixed, we set X for one of the spaces \(C[-1, 1]\) or \(L^{p}_{(\alpha, \beta)}\).

For \(n \in\mathbb{N}\), the Jacobi polynomial \(R^{(\alpha,\beta)}_{n}\) is the unique polynomial of degree n which satisfies

$$R^{(\alpha,\beta)}_{n} (1) = 1 \quad\text{and}\quad \int_{-1}^{1} Q_{n- 1}(x)R^{(\alpha,\beta)}_{n} (x) \varrho^{ \alpha,\beta }(x)\,dx= 0 $$

for all \(Q_{n-1} \in\mathbb{P}_{n- 1}\). We also take \(R^{(\alpha,\beta)}_{0} (x) = 1\).

For \(f \in X \), the Fourier–Jacobi coefficients are defined by

$$\bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle = \int_{-1}^{1} f(x)R^{(\alpha,\beta)}_{n} (x) \varrho^{ \alpha,\beta}(x)\,dx,\quad n \in\mathbb{N}_{0}, $$

and the associated expansion is

$$ f(x)\sim\sum_{n=0}^{\infty}\bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle w^{(\alpha,\beta)}_{n} R^{(\alpha,\beta)}_{n} (x). $$
(4)

It is known that each \(f \in L^{1}_{ (\alpha,\beta)}\) is completely determined a.e. by its Fourier–Jacobi coefficients.

Definition 1.1

For fixed \(\gamma> 0\) and \(t > 0\), the generalized Jacobi–Weierstrass kernel is defined by

$$ W_{t,\gamma} (x) = \sum_{n=0}^{\infty}e^{{ - t\lambda_{n}^{\gamma}}} w^{(\alpha,\beta)}_{n} R^{(\alpha,\beta)}_{n} (x),\quad x \in[- 1, 1]. $$
(5)

For \(f \in X \), the generalized Jacobi–Weierstrass (or Abel–Cartwright) operator is defined by

$$ C_{t,\gamma} (f, x) = \int_{-1}^{1} \tau_{y}(f, x)W_{t,\gamma} (y)\varrho^{ \alpha,\beta}(y)\,dy,\quad x \in[- 1, 1], $$
(6)

where \(\tau_{y}(f, x)\) is the translation given in Theorem 2.1 below.

Of course the kernel \(W_{t, \gamma}\) and the operator \(C_{t, \gamma}\) also depend on α and β but, for simplicity, we omit these indexes. The (classical) Jacobi–Weierstrass operators correspond to \(\gamma= 1\).

The generalized Jacobi–Weierstrass operators have been studied in different papers, but only for parameters satisfying \(0 < \gamma\leq 1\). This restriction was considered because in such a case the kernels \(W_{t, \gamma}\) are positive and the family \(\{C_{t, \gamma}\}\) can be considered as formed by positive operators (see [2, 3], [7], pp. 96–97) and/or as a semigroup of contractions (see [2], pp. 49–52, and [18]). For \(\gamma > 1\), one cannot expect the positivity of \(W_{t, \gamma}\). For instance, it is known that the analogous generalized Weierstrass kernels for trigonometric expansion are not positive when \(\gamma > 1\) (see [6], p. 263).

In this paper we will prove that the operators \(C_{t, \gamma}\) can be used as a realization of some K-functionals which usually appear in some approximation problems related to Jacobi expansions.

For fixed real \(\gamma> 0\), let \(\Phi^{\gamma}(X )\) denote the family of all \(f \in X \) for which there exists \(\Psi^{\gamma}(f) \in X \) satisfying

$$\Psi^{\gamma}(f) (x) \sim\sum_{n=0}^{\infty}\lambda_{n}^{\gamma}\bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle w^{(\alpha,\beta)}_{n} R^{(\alpha,\beta)}_{n} (x). $$

The associated K-functional is defined by

$$ K_{\gamma}(f, t)=K_{\gamma}(f, t)_{\alpha,\beta} = \inf_{ g \in \Psi^{\gamma}(X )} \bigl\{ \Vert f - g \Vert _{X} + t \bigl\Vert \Psi^{\gamma}(g) \bigr\Vert _{X } \bigr\} $$
(7)

for \(f \in X \) and \(t > 0\). For different realizations of these K-functionals, see [8], Theorem 7.1, and [10], Lemma 2.3. We will not use the characterization of these K-functionals in terms of moduli of smoothness. We will show that, for any \(\gamma > 0\),

$$\sup_{ 0< s\leq t} \bigl\Vert (I - C_{s,\gamma}) (f) \bigr\Vert _{X} \approx K_{\gamma}(f, t). $$

The notation \(A(f, t) \approx B(f, t)\) means that there exists a positive constant C such that \(C^{-1}A(f, t) \leq B(f, t) \leq CA(f, t)\) with C independent of f and t.

Following [19], for \(\gamma > 0\), define

$$ (I - C_{t, 1})^{\gamma}= \sum _{j=0}^{\infty}(- 1)^{j} \binom{\gamma}{j} C_{jt, 1}, $$
(8)

where

$$\binom{\gamma}{0} = 1 \quad\text{and} \quad\binom{\gamma}{j} = \prod _{k=1}^{j} \frac{\gamma- k + 1}{ k} \quad\text{for } j \in \mathbb{N}. $$

For these operators, we will show the relations

$$K_{\gamma}\bigl(f, t^{\gamma}\bigr) \approx\sup _{ 0< s\leq t} \bigl\Vert (I - C_{s,1})^{\gamma}(f) \bigr\Vert _{X } \approx\sup_{ 0< s\leq t^{\gamma}} \bigl\Vert (I - C_{s,\gamma}) (f) \bigr\Vert _{X }. $$

It is known that, if \(Q_{n}\) is a trigonometric polynomial of degree not greater than n and \(r \in\mathbb{N}\), then

$$\bigl\Vert Q^{(r)}_{n} \bigr\Vert _{p} \leq \biggl(\frac{ n}{ 2 \sin(nh)} \biggr)^{r} \bigl\Vert (1 - T_{h})^{r}(Q_{n}) \bigr\Vert _{p},\quad h \in(0, \pi/n), $$

where \(\Vert\cdot\Vert_{p}\) denotes the \(L^{p}\)-norm of 2π-periodic functions and \(T_{h}\) is the translation operator. That is, \(T_{h}Q(x) = Q(x + h)\). These inequalities are due to Nikolskii [11] and Stechkin [13]. For similar inequalities for algebraic polynomials, see [4] and the references given there. Here we will verify an analogous inequality by considering the operators \(\Psi^{r}\) and the linear combination of the Jacobi–Weierstrass operators \(C_{t,1}\).

In Sect. 2 we collect some definitions and results which will be needed later. The main results are given in Sect. 3, where the result concerning simultaneous approximation is also included. Finally, in Sect. 4 we present a Nikolskii–Stechkin type inequality.

2 Auxiliary results

We need a convolution structure due to Askey and Wainger (see [1]).

Theorem 2.1

For each \(h \in[-1, 1)\), there exists a function \(\tau_{h}: X \to X \) with the following properties:

  1. (i)

    For each \(f \in X \), one has

    $$\Vert \tau_{h} f \Vert _{X } \leq \Vert f \Vert _{X}, \qquad \lim_{ h\to1-} \bigl\Vert \tau_{h}(f) - f \bigr\Vert _{X } = 0 $$

    and

    $$\bigl\langle \tau_{h}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle = R^{(\alpha,\beta)}_{n}(h) \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle ,\quad n \in\mathbb{N}_{0}. $$
  2. (ii)

    For \(f\in X\) and \(g \in L^{1}_{\alpha,\beta}\), the integral

    $$(f * g) (x):= \int_{-1}^{1} \tau_{y}(f, x) g(y) \varrho^{ \alpha,\beta}(y)\,dy $$

    exists a.e. in \([-1.1]\),

    $$f * g = g * f,\qquad f * g \in X,\qquad \Vert f * g \Vert _{p}\leq \Vert g \Vert _{1} \Vert f \Vert _{X} $$

    and

    $$ \bigl\langle f * g,R^{(\alpha,\beta)}_{n} \bigr\rangle = \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle \bigl\langle g,R^{(\alpha,\beta)}_{n} \bigr\rangle ,\quad n \in\mathbb{N}_{0}. $$
    (9)

For \(j > \alpha + 1/2\) and \(f \in X \), let

$$S^{j}_{m} (f) = \sum_{k=0}^{m} \frac{A^{j}_{ m-k}}{ A^{j}_{m}} \bigl\langle f,R^{\alpha,\beta)}_{k} \bigr\rangle w^{\alpha,\beta}_{k} R^{(\alpha,\beta)}_{k} (x),\quad A^{j}_{m} = \binom{m+j}{m}, $$

be the mth Cesàro means of order j. It is known that there exists a constant C such that

$$ \bigl\Vert S^{j}_{m} \bigr\Vert \leq C, $$
(10)

and, for each \(f \in X \), one has ([2], Corollary 3.3.3, or [7], Theorem A)

$$ \lim_{ m\to\infty} \bigl\Vert f - S^{j}_{m} (f) \bigr\Vert _{X } = 0. $$
(11)

We need some classical results related to Banach spaces.

Definition 2.2

Let Y be a real Banach space and \(B(Y )\) be the Banach algebra of all bounded linear operators \(B: Y \to Y\). A uniformly bounded family of operators \(\{T(t): t \geq0\}\) in \(B(Y )\) is called an equi-bounded semigroup of class \((C_{0})\) if

$$ T(s)T(t) = T(s + t)\quad \text{for }s, t \geq 0,\qquad T(0) = I, $$
(12)

and \(\lim_{ t\to0+} \Vert f - T(t)f\Vert _{Y} = 0\) for each \(f \in Y \).

Let Y, \(B(Y )\) and \(\{T(t): t > 0\}\) be an equi-bounded semigroup as in Definition 2.2. Let \(D(Q)\) be the family of all \(g \in Y\), for which there exists \(Q(g) \in Y\) such that

$$ Q(g) = \lim_{ t\to0+} \frac{1}{t} \bigl[T(t) - I \bigr]g $$
(13)

(the limit is considered in the norm of Y). The operator \(Q: D(Q) \to Y\) is called the infinitesimal generator of the semigroup \(\{T(t): t \geq0\}\). It is known that Q is a closed linear operator and \(D(Q) \) is dense in Y. For properties of semigroups of operators, see [5].

For \(r \in\mathbb{N}\), set

$$D \bigl(Q^{r+1} \bigr) = \bigl\{ f \in Y: f \in D \bigl(Q^{r} \bigr) \text{ and } Q^{r}(f) \in D(Q) \bigr\} $$

and, for \(f \in D(Q^{r+1})\),

$$ Q^{r+1}(f) = Q \bigl(Q^{r}(f) \bigr). $$
(14)

A family of operators \(S = \{S_{t},: t > 0\}\), \(S_{t} \in B(Y )\) for each \(t > 0\) is called a (commutative) strong approximation process for Y if, for all \(f \in Y\) and \(s, t > 0\),

$$S_{s} \bigl(S_{t}(f) \bigr) = S_{t} \bigl(S_{s}(f) \bigr),\qquad \bigl\Vert S_{t}(f) \bigr\Vert _{Y} \leq \Lambda \Vert f \Vert _{Y} \quad\text{and} \quad\lim _{ t\to0+} \bigl\Vert f -S_{t}(f) \bigr\Vert _{Y} = 0, $$

where Λ is a constant. In such a case, we set

$$\theta_{S}(f, t) =\sup_{ 0< s\leq t} \bigl\Vert f - S_{s}(f) \bigr\Vert _{Y}. $$

Let \(\phi: [0,1)\to \mathbb{R}^{+}\) be a positive increasing function, \(\phi(t)\to0\) as \(t \to0\), and \(Y_{0}\) be a subspace of Y. We say that S is saturated with order ϕ and with trivial subspace \(Y_{0}\) if every \(f \in Y\) satisfying

$$\lim_{ t\to0+}\frac{\theta_{S} (f, t)}{\phi(t)} = 0 $$

belongs to \(Y_{0}\) and there exists \(f \in Y\setminus Y_{0}\) satisfying \(\theta_{S}(f, t) \leq C(f)\phi(t)\). The following assertion is known (for instance, see [2], Theorem 2.4.2).

Theorem 2.3

Assume that Y is a Banach space, \(D(B)\) is a dense subspace of Y, and \(B: D(B) \to Y\) is a closed linear operator. Let \(S=\{ S_{t}: t > 0\}\) be a strong approximation process in Y satisfying \(S_{t}(f) \in D(B)\) for any \(f \in Y\) and each \(t > 0\). If there exists a constant \(\gamma_{0}\) such that, for all \(g \in D(B)\),

$$ \lim_{ t\to0+}\biggl\Vert \frac{S_{t}(g) - g}{ t^{\gamma_{0}}} - B(g)\biggr\Vert _{Y} = 0, $$
(15)

then the strong approximation process S is saturated with order \(t^{\gamma_{0}}\) and the trivial space is the kernel of B.

3 The operators \(C_{t, \gamma}\) as a semigroup

In fact, it is known that, for \(x \in(-1,1)\), \(| R^{(\alpha,\beta)}_{n} (x) |< 1\), [14], pp. 163–164, and there exists a constant C such that, for each \(n\in\mathbb{N}_{0}\),

$$ w^{(\alpha,\beta)}_{n} \leq Cn^{2\alpha+1}. $$
(16)

These relations can be used to prove that the series in (5) converges absolutely and uniformly in \([-1, 1]\). Thus \(W_{t, \gamma} \in L^{1}_{(\alpha,\beta)}\) and, for each \(f \in L^{1}_{ (\alpha,\beta)}\), the series \(C_{t, \gamma} (f)\) converges absolutely and uniformly in \([-1, 1]\). Moreover,

$$C_{t, \gamma} (f, x) = (W_{t, \gamma}* f) (x)=\sum _{n=0}^{\infty}e^{{ -t\lambda_{n}^{\gamma}}} \bigl\langle f,R_{n}^{(\alpha,\beta)} \bigr\rangle w_{n}^{(\alpha,\beta)} R_{n}^{(\alpha,\beta)}(x). $$

For these assertions, see [2], p. 30.

Our first result seems to be known. For convenience of the reader, we include a proof.

Theorem 3.1

For each \(\gamma > 0\), the family of operators \(\{ C_{t, \gamma}: t > 0\}\) is an equi-bounded semigroup of operators in X.

Proof. It follows from Theorem 3.9 of [15] that the family of operators \(\{ C_{t, \gamma}: t > 0\}\) is uniformly bounded.

Condition (12) is derived from the properties of the convolution. In fact, it follows from (9) that, for each \(f \in X\) and \(k \in\mathbb{N}_{0}\),

$$\begin{aligned} \bigl\langle C_{s+t} (f),R^{(\alpha,\beta)}_{k} \bigr\rangle &= e^{{ -(s + t)\lambda _{n}^{\gamma}}} \bigl\langle f,R^{(\alpha,\beta)}_{k} \bigr\rangle = e^{{ -s\lambda_{n}^{\gamma}}} \bigl\langle C_{t, \gamma} (f),R^{(\alpha,\beta)}_{k} \bigr\rangle \\ &= \bigl\langle C_{s,\gamma} \bigl(C_{t, \gamma} (f) \bigr),R^{(\alpha,\beta)}_{k} \bigr\rangle \end{aligned}$$

and this implies \(C_{s+t}(f) = (C_{s,\gamma} \circ C_{t, \gamma}) (f)\).

Finally, for each \(k \in\mathbb{N}_{0}\),

$$ C_{t, \gamma} \bigl(R^{(\alpha,\beta)}_{k} \bigr) (x) = e^{{ -t\lambda_{n}^{\gamma}}} R^{(\alpha,\beta)}_{k} (x). $$
(17)

Hence

$$\lim_{t\to0+} \bigl\Vert R^{(\alpha,\beta)}_{k} - C_{t, \gamma} \bigl(R^{(\alpha,\beta)}_{k} \bigr) \bigr\Vert _{X} = 0. $$

Since the operators \(C_{t, \gamma}\) are linear and uniformly bounded and the polynomials are dense in X, the last equation holds for every \(f \in X\).

Taking into account Theorem 3.1, we denote by \(A_{\gamma}\) the infinitesimal generator of \(C_{t, \gamma}\) and by \(D(A_{\gamma})=D(A_{\gamma}(\alpha, \beta))\) the domain of \(A_{\gamma}\). In the next result we give a description of the infinitesimal generator.

Theorem 3.2

If \(\gamma, t > 0\) and \(A_{\gamma}: D(A_{\gamma}) \to X\) is the infinitesimal generator of \(C_{t, \gamma}\), then

$$D(A_{\gamma}) = \Psi^{\gamma}(X)\quad \textit{and} \quad{-} A_{\gamma}(f) = \Psi^{\gamma}(f) $$

for each \(f \in\Psi^{\gamma}(X)\).

Moreover, for each \(r \in\mathbb{N}\) and \(f \in D(A^{r}_{\gamma})\),

$$ D \bigl(A^{r}_{\gamma}\bigr) = \Psi^{r\gamma} (X)\quad \textit{and}\quad (-1)^{r}A^{r}_{\gamma}(f) = \Psi^{r\gamma} (f), $$
(18)

where \(A^{r}_{\gamma}\) is defined as in (14).

Proof

Since \(A_{\gamma}\) is the infinitesimal generator of the semi-group (see (13)), \(A_{\gamma}: D(A_{\gamma}) \to X\) is a closed operator.

If \(f \in D(A_{\gamma})\), then

$$ \bigl\langle A_{\gamma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle = \lim_{ t\to0+}\frac{ 1}{t} \bigl( e^{{ -t\lambda_{n}^{\gamma}}} - 1 \bigr) \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle = -\lambda_{n}^{\gamma}\bigl\langle f, R^{(\alpha,\beta)}_{n} \bigr\rangle . $$
(19)

Thus \(f\in\Psi^{\gamma}(X)\) and

$$\Psi^{\gamma}(f) = -A_{\gamma}(f). $$

In particular, for each polynomial P, one has \(P \in D(A_{\gamma})\) and \(\Psi^{\gamma}(P) = -A_{\gamma}(P)\).

On the other hand, fix an integer \(j > \alpha+ 1/2\). For \(f \in\Psi ^{\gamma}(X)\), let \(S^{j}_{m} (f)\) and \(S^{j}_{m}( \Psi^{\gamma}(f))\) be the mth Cesàro means of order j of f and \(\Psi^{\gamma}(f)\), respectively. We know that (see (11))

$$S^{j}_{m} (f) \to f, \quad m \to \infty $$

and

$$-A_{\gamma}\bigl(S^{j}_{m} (f) \bigr) = \Psi^{\gamma}\bigl(S^{j}_{m} (f) \bigr) = S^{j}_{m} \bigl(\Psi^{\gamma}(f) \bigr) \to \Psi^{\gamma}(f). $$

Since \(-A_{\gamma}\) is a closed operator, \(f \in D(A_{\gamma})\) and \(-A_{\gamma}(f) = \Psi^{\gamma}(f)\).

Equations (18) can be proved by recurrence. For instance, (19) can be written as

$$\bigl\langle A^{2}_{\gamma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle = \bigl\langle A_{\gamma}\bigl(A_{\gamma}(f) \bigr),R^{(\alpha,\beta)}_{n} \bigr\rangle =-\lambda_{n}^{\gamma}\bigl\langle A_{\gamma}(f), R^{(\alpha,\beta)}_{n} \bigr\rangle = \lambda_{n}^{2\gamma} \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle . $$

 □

Theorem 3.3

(i) If for \(\gamma, t > 0\), and \(f \in X\)

$$\theta_{\gamma}(f, t)= \theta_{\gamma}(f, t)_{\alpha,\beta} = \sup _{ 0< s\leq t} \bigl\Vert (I - C_{s,\gamma} ) (f) \bigr\Vert , $$

and \(K_{\gamma}(f,t)\) is defined by (7), then

$$\theta_{\gamma}(f, t) \approx K_{\gamma}(f, t). $$

(ii) The strong approximation process \(\{ C_{t, \gamma}; t > 0\}\) is saturated with order t and the trivial class consists of the constant functions.

Proof

(i) From Theorem 3.2 we know that \(-\Psi^{\gamma}\) is the infinitesimal generator of \(\{C_{t, \gamma}\}\) and \(D(A_{\gamma}) =\Psi^{\gamma}(X)\). Thus, the result is a simple consequence of [17], Theorem 1.1, or [5], p. 192.

(ii) We will derive the result from Theorem 2.3, with \(B =\Psi^{\gamma}\) and \(D(B) = D(A_{\gamma})\). We should verify that \(C_{t, \gamma} (f) \in D(A_{\gamma})\) for any \(f \in X \) and each \(t > 0\).

For any \(f\in X\), the Fourier–Jacobi coefficients of f are bounded by \(\Vert f\Vert_{L^{1}_{(\alpha,\beta)}}\). Taking into account (16), for every \(x\in[-1,1]\),

$$\begin{aligned} &\Biggl\vert \sum_{n=1}^{\infty}\lambda_{n}^{\gamma}\exp\{ -t\lambda_{n}\} \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle w^{(\alpha,\beta)}_{n} R^{(\alpha,\beta)}_{n} (x) \Biggr\vert \\ &\quad\leq \Vert f \Vert _{L^{1}_{(\alpha,\beta)}} \sum_{n=1}^{\infty}\lambda_{n}^{\gamma}\exp \bigl\{ -t\lambda_{n}^{\gamma}\bigr\} w^{(\alpha,\beta)}_{n} \\ &\quad\leq C \Vert f \Vert _{L^{1}_{(\alpha,\beta)}} \sum_{n=1}^{\infty}\lambda_{n}^{\gamma}\exp \bigl\{ -t\lambda_{n}^{\gamma}\bigr\} n^{2\alpha+1} < \infty. \end{aligned}$$

Since the series converges absolutely and uniformly, it defines a function \(g_{t}\in X\) satisfying

$$\bigl\langle g_{t},R^{(\alpha,\beta)}_{n} \bigr\rangle = \lambda_{n}^{\gamma}\exp \bigl\{ -t\lambda_{n}^{\gamma}\bigr\} \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle = \lambda_{n}^{\gamma}\bigl\langle C_{t,\gamma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle ,\quad n\in\mathbb{N}. $$

By definition of the operator \(\Psi^{\gamma}\), \(C_{t, \gamma} (f) \in \Psi^{\gamma}(X )\) (Theorem 3.2) and

$$\Psi^{\gamma}\bigl(C_{t, \gamma} (f) \bigr) = g_{t}. $$

We have proved that \(C_{t, \gamma} (X) \in D(A_{\gamma})\).

If \(g\in\Psi^{\gamma}(X) =D(A_{\gamma})\), by definition of the infinitesimal generator,

$$\lim_{ t\to0+}\biggl\Vert \frac{C_{t,\gamma}(g) - g}{ t} - A_{\gamma}(g)\biggr\Vert _{Y} = 0. $$

If \(f \in\Psi^{\gamma}(X )\) and \(A_{\gamma}(f)=-\Psi^{\gamma}(f) = 0\), then \(\langle f,R^{(\alpha,\beta)}_{n} \rangle = 0\) for all \(n \in \mathbb{N}\). Therefore f is a constant.

From part (i), if \(g \in\Psi^{\gamma}(X )\), then

$$\theta_{\gamma}(g, t) \leq C K _{\gamma}(g, t) \leq C t \bigl\Vert \Psi^{\gamma}(g) \bigr\Vert _{X }. $$

Hence, the family

$$\bigl\{ f \in X: \exists C(f) \text{ such that } \theta_{\gamma}(f, t) \leq C(f) t \bigr\} $$

contains nonconstant functions.

Now, from Theorem 2.3, we know that the strong approximation process \(\{ C_{t, \gamma}: t > 0\}\) is saturated with order t. □

Remark 3.4

Some characterizations of the saturation class of the strong approximation process \(\{C_{t, \gamma}: t > 0\}\) can be given as in [2], Theorems 5.1.1 and 7.4.1, where the case \(\gamma=1\) was considered. When \(\gamma > 0\) is not an integer, fractional derivatives should be considered. This task would lead us far from our main topic.

Remark 3.5

A relation similar to (i) in Theorem 3.3 is asserted in [16], p. 2885, for the discrete case and Gauss–Weierstrass type means

$$\widetilde{W}_{\Omega(n),\gamma} (f,x) =\sum_{n=0}^{\infty}e^{{ -(\Omega(k)/\Omega(n))^{\gamma}}} \bigl\langle f,R_{n}^{(\alpha,\beta )} \bigr\rangle w^{(\alpha,\beta)}_{n} R^{(\alpha,\beta)}_{n} (x), $$

with Ω varying in a specified class of functions. The proof suggested there is different from the one given here (it does not use the semi-group structure). The main argument in [16] is that some abstract Riesz means are equivalent (as approximation processes) to some Gauss–Weierstrass type means. This kind of equivalence can also be derived by using Corollary 5.4 of [9]. Anyway, the arguments of [16] and the proof given here are related because both use [15], Theorem 3.9, to obtain a uniformly bounded family of multipliers. Apart from this, other topics considered here are not connected with [16].

The arguments used in the proof of Theorem 3.2 can be used to derive similar relations concerning the fractional powers of the Jacobi–Weierstrass operators \(\{C_{t, 1}\}\).

Recall that \(A_{1}: D(A_{1}) \to X\) is the infinitesimal generator of \(\{ C_{t, 1}, t > 0\}\). For \(\gamma> 0\), let \(D((-A_{1})^{\gamma},X )\) be the family of all \(f \in X \), for which there exists an element \((-A_{1})^{\gamma}(f) \in X \) satisfying

$$ \lim_{ t\to0+}\biggl\Vert (-A_{1})^{\gamma}(f) -\frac{1}{t^{\gamma}} (I - C_{t, 1})^{\gamma}(f)\biggr\Vert _{X } = 0, $$
(20)

where \((I - C_{t, 1})^{\gamma}(f)\) is defined by (8). This induces a map

$$\bigl(- A^{1} \bigr)^{\gamma}: D \bigl( \bigl(-A^{1} \bigr)^{\gamma},X \bigr) \to X $$

which is called the fractional power of order γ of \(- A_{1}\).

Proposition 3.6

If \(\gamma> 0\) and \((- A_{1})^{\gamma}\) is the fractional power of order γ of \(- A_{1}\), then

$$D \bigl((- A_{1})^{\gamma},X \bigr) = \Psi^{\gamma}(X ) $$

and, for each \(f \in\Psi^{\gamma}(X )\),

$$ \Psi^{\gamma}(f) = \lim_{ t\to0+} \frac{1}{t^{\gamma}} (I - C_{t, 1})^{\gamma}(f) = \lim _{ t\to0+} \frac{1}{t} \bigl( f-C_{t, \gamma} (f) \bigr). $$
(21)

Proof

If γ is a positive integer or \(| a |< 1\), the Taylor expansion gives

$$(1 - a)^{\gamma}= \sum_{j=0}^{\infty}(-1)^{j} \binom{\gamma}{j} a^{j}. $$

Notice that

$$\begin{aligned} \bigl\langle (I - C_{t, 1})^{\gamma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle &= \sum_{k=0}^{\infty}(-1)^{k} \binom{\gamma}{k} \bigl\langle C_{kt, 1}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle \\ &= \sum_{k=0}^{\infty}(-1)^{k} \binom{\gamma}{k} \bigl\langle W_{kt},R^{(\alpha,\beta)}_{n} \bigr\rangle \bigl\langle f,R^{(\alpha,\beta )}_{n} \bigr\rangle \\ &= \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle \sum _{k=0}^{\infty}(-1)^{k} \binom {\gamma}{k} \exp ( - kt\lambda_{n}) ) \\ &= \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle \bigl( 1-\exp(-t\lambda_{n}) \bigr)^{\gamma}. \end{aligned}$$
(22)

Therefore, if \(f \in D((- A_{1})^{\gamma},X )\), then

$$\bigl\langle (-A_{1})^{\gamma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle = (\lambda_{n})^{\gamma}\bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle . $$

Hence \(f \in\Psi^{\gamma}(X )\) and \((- A_{1})^{\gamma}(f) =\Psi^{\gamma}(f)\).

It is clear that, for each polynomial P, one has \(P \in D((- A_{1})^{\gamma},X)\) and

$$(- A_{1})^{\gamma}(P) = \Psi^{\gamma}(P). $$

On the other hand, fix an integer \(j > \alpha + 1/2\). For \(f \in\Psi ^{\gamma}(X )\), let \(S^{j}_{m} (f)\) and \(S^{j}_{m} (\Psi^{\gamma}(f))\) be the mth Cesàro means of order j of f and \(\Psi^{\gamma}(f)\), respectively. From (11), as in the proof of Theorem 3.2, one has \(\lim_{ m\to\infty} \Vert S^{j}_{m}(f) - f\Vert_{X } = 0\) and

$$\begin{aligned} \lim_{ m\to\infty} \bigl\Vert \bigl(- A^{1} \bigr)^{\gamma}\bigl(S^{j}_{m} (f) \bigr) - \Psi^{\gamma}(f) \bigr\Vert _{X }& = \lim_{ m\to\infty} \bigl\Vert \Psi^{\gamma}\bigl(S^{j}_{m} (f) \bigr) - \Psi^{\gamma}(f) \bigr\Vert _{X } \\ &= \lim_{ m\to\infty} \bigl\Vert S^{j}_{m} \bigl( \Psi^{\gamma}(f) \bigr) - \Psi^{\gamma}(f) \bigr\Vert _{X} = 0. \end{aligned}$$

It was proved in [19], Theorem 4, that \(D((- A_{1})^{\gamma},X)\) is dense in X and \((- A_{1})^{\gamma}\) is a closed operator. Hence \(f \in D((- A_{1})^{\gamma},X)\) and \((- A_{1})^{\gamma}(f) = \Psi^{\gamma}(f)\).

The last equality in (21) was proved in Theorem 3.2, because \(\Psi^{\gamma}\) is the infinitesimal generator of \(\{ C_{t, \gamma}, t > 0\}\). □

Theorem 3.7

For fixed \(\gamma > 0\), one has

$$K_{\gamma}\bigl(f, t^{\gamma}\bigr) \approx \sup _{ 0< s\leq t} \bigl\Vert (I - C_{s,1})^{\gamma}(f) \bigr\Vert _{X } \approx \theta_{\gamma}\bigl(f, t^{\gamma}\bigr) $$

for each \(f \in X \) and \(t > 0\).

Proof

From Theorems 3.1 and 3.2 we know that the family \(\{C_{t, 1}, t \geq0\}\) is a semi-group of operators of class \((C_{0})\) with the infinitesimal generator \(A_{1} =-\Psi^{1}\). From Theorem 1.1 of [17], we know that, for all \(f \in X \) and \(t > 0\),

$$\inf_{ g \in D((- A_{1})^{\gamma},X )} \bigl( \Vert f - g \Vert _{X } + t^{\gamma}\bigl\Vert (- A_{1})^{\gamma}(g) \bigr\Vert _{X } \bigr) \approx\sup_{ 0< s\leq t} \bigl\Vert (I - C_{s,1})^{\gamma}(f) \bigr\Vert _{X }, $$

where \((- A_{1})^{\gamma}\) is given as in (20). But it was verified in Proposition 3.6 that \(\Psi^{\gamma}(X ) = D((- A_{1})^{\gamma},X )\) and \((- A_{1})^{\gamma}(g) = \Psi^{\gamma}(g)\) for each \(g \in\Psi^{\gamma}(X )\).

The equivalence with \(\theta_{\gamma}(f, t^{\gamma})\) follows from Theorem 3.3. □

Remark 3.8

When γ is an integer, Theorem 3.7 is similar to the Main Theorem in [18], p. 390, but the authors assumed that the operators are positive (plus other conditions).

Remark 3.9

The results of Theorem 3.7 allow us to obtain equivalent relations between fractional powers \((I - C_{s,1})^{\gamma}\) and some Riesz means as in Theorem 5.1 of [9].

Some result concerning simultaneous approximation can be derived from the ones given above.

Theorem 3.10

If \(\gamma,\sigma\), and t are positive real numbers and \(f \in \Psi^{\sigma}(X )\), then

$$\begin{aligned} &C_{t, \gamma} (f), (I - C_{t, 1})^{\gamma}(f) \in \Psi^{\sigma}(X ), \\ &\bigl\Vert \Psi^{\sigma}(f) - \Psi^{\sigma}\bigl(C_{t, \gamma} (f) \bigr) \bigr\Vert _{X } \leq C \theta_{\gamma}\bigl( \Psi^{\sigma}(f), t \bigr) \end{aligned}$$

and

$$\bigl\Vert \Psi^{\sigma}\bigl((I - C_{t, 1})^{\gamma}(f) \bigr) \bigr\Vert _{X } \leq C \theta_{\gamma}\bigl( \Psi^{\sigma}(f), t^{\gamma}\bigr), $$

where the constant C is independent of f and t.

Proof

If \(f \in\Psi^{\sigma}(X )\) and \(n \in\mathbb{N}_{0}\), from (17) we obtain

$$\begin{aligned} \bigl\langle C_{t, \gamma} \bigl( \Psi^{\sigma}(f) \bigr),R^{(\alpha,\beta)}_{n} \bigr\rangle &= \exp \bigl(- t \lambda_{n}^{\gamma}\bigr) \bigl\langle \Psi^{\sigma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle \\ &= \lambda_{n}^{\sigma}\exp \bigl(- t\lambda_{n}^{\gamma}\bigr) \bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle = \lambda_{n}^{\sigma}\bigl\langle C_{t,\gamma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle \end{aligned}$$

and from (22) one has

$$\begin{aligned} \bigl\langle (I - C_{t, 1})^{\gamma}\bigl( \Psi^{\sigma}(f) \bigr),R^{(\alpha,\beta)}_{n} \bigr\rangle &= \bigl( 1- \exp(- t \lambda_{n}) \bigr)^{\gamma}\bigl\langle \Psi^{\sigma}(f);R^{(\alpha,\beta)}_{n} \bigr\rangle \\ &= \lambda_{n}^{\sigma}\bigl( 1 - \exp(- t\lambda_{n}) \bigr)^{\gamma}\bigl\langle f,R^{(\alpha,\beta)}_{n} \bigr\rangle = \lambda_{n}^{\sigma}\bigl\langle (I - C_{t, 1})^{\gamma}(f),R^{(\alpha,\beta)}_{n} \bigr\rangle . \end{aligned}$$

Therefore \(C_{t, \gamma} (f), (I - C_{t, 1})^{\gamma}(f) \in \Psi^{\sigma}(X)\),

$$\Psi^{\sigma}\bigl(C_{t, \gamma} (f) \bigr) = C_{t,\gamma} \bigl( \Psi^{\sigma}(f) \bigr) \quad\text{and} \quad\Psi^{\sigma}\bigl((I - C_{t, 1})^{\gamma}(f) \bigr) = (I - C_{t, 1})^{\gamma}\bigl( \Psi^{\sigma}(f) \bigr). $$

Now, from Theorem 3.3 one has

$$\bigl\Vert \Psi^{\sigma}(f) - \Psi^{\sigma}(C_{t, \gamma}) \bigr\Vert _{X} = \bigl\Vert (I - C_{t, \gamma} ) \bigl( \Psi^{\sigma}(f) \bigr) \bigr\Vert _{X } \leq C \theta_{\gamma}\bigl( \Psi^{\sigma}(f), t \bigr), $$

and using Theorem 3.7 we obtain

$$\bigl\Vert \Psi^{\sigma}\bigl((I - C_{t, 1})^{\gamma}(f) \bigr) \bigr\Vert _{X } = \bigl\Vert (I - C_{t, 1})^{\gamma}\bigl( \Psi^{\sigma}(f) \bigr) \bigr\Vert _{X } \leq C \theta_{\gamma} \bigl( \Psi^{\sigma}(f), t^{\gamma}\bigr). $$

 □

4 A Nikolskii–Stechkin type inequality

Theorem 4.1

For each \(r \in\mathbb{N}\), there exists a constant C, depending upon r, such that, for every \(\lambda\geq1\) and for each polynomial \(P \in\mathbb{P}_{\xi(\lambda)}\),

$$\bigl\Vert \Psi^{r}(P) \bigr\Vert _{X} \leq C \lambda^{r} \sup_{ 0< h\leq1/\lambda} \bigl\Vert (I - C_{h,1})^{r}(P) \bigr\Vert _{X }, $$

where

$$\xi(\lambda) = \max \bigl\{ k \in\mathbb{N}_{0}: k(k + \alpha+\beta + 1) < \lambda \bigr\} . $$

Proof

In this proof the infinitesimal generator of \(\{C_{t, 1}: t > 0\}\) is denoted by A.

From the proof of Lemma 1 in [12] we know that, given \(r \in\mathbb{N}\), there exists a constant \(C_{1} = C(r)\) such that, for each \(f \in X \) and \(t > 0\), there is \(g_{t} \in D(A^{r+1})\) satisfying

$$\begin{aligned} & \Vert f - g_{t} \Vert _{X } \leq\sup _{ 0< h\leq t} \bigl\Vert (I - C_{h,1})^{r}f \bigr\Vert _{X }, \end{aligned}$$
(23)
$$\begin{aligned} & \bigl\Vert A^{r+1}(g_{t}) \bigr\Vert _{X } \leq C_{1} \frac{1}{t^{r+1}} \sup_{0< h\leq t} \bigl\Vert (I - C_{h,1})^{r}f \bigr\Vert _{X } \end{aligned}$$
(24)

and

$$ \bigl\Vert (- A)^{r}(g_{t}) \bigr\Vert _{X } \leq C_{1} \frac{1}{ t^{r}} \sup_{ 0< h\leq t} \bigl\Vert (I - C_{h,1})^{r}f \bigr\Vert _{X }. $$
(25)

As in [9], for \(\lambda> 0\) and \(f \in X \), consider the best approximation

$$E_{\lambda}(f) = \inf \bigl\{ \Vert f - P \Vert _{X }: P \in \mathbb{P}_{\xi(\lambda)} \bigr\} . $$

It was proved there (Theorem 6.1) that there exists a constant \(C_{2} = C(r, \alpha, \beta)\) such that, for \(\lambda> 0\) and \(f \in X \),

$$ E_{\lambda}(f) \leq C_{2} K_{r+1} \bigl(f, \lambda^{- r- 1} \bigr), $$
(26)

and (Theorem 3.2) for each \(Q \in\mathbb{P}_{\xi(\lambda)}\),

$$ \bigl\Vert \Psi^{r}(Q) \bigr\Vert _{ X } \leq C_{2} \lambda^{r} \Vert Q \Vert _{ X }. $$
(27)

Now, fix \(\lambda > 0\) and \(P \in\mathbb{P}_{\xi(\lambda)}\). Let \(g_{t} \in D(A^{r+1}) = \Psi^{r+1}(X)\) (see (18)) be given as (23)–(25) with \(t = 1/\lambda\) and \(f = P\).

For \(\varepsilon > 0\) and \(k \in\mathbb{N}_{0}\), choose

$$ q(g_{t}, k) \in\mathbb{P}_{\xi(2^{k}\lambda)} $$
(28)

such that

$$ \bigl\Vert g_{t} - q(g_{t}, k)) \bigr\Vert _{X} \leq (1 + \varepsilon)E_{2^{k}\lambda} (g_{t}). $$
(29)

From (26), (18), and (24) we know that

$$\begin{aligned} \bigl\Vert g_{t} - q(g_{t}, k))\bigr\Vert _{X } &\leq C_{2}(1 + \varepsilon)K_{r+1} \bigl(g_{t}, \bigl(2^{k}\lambda \bigr)^{- r- 1} \bigr) \\ &\leq\frac{C_{2}(1 + \varepsilon)}{ (2^{k}\lambda)^{r+1}} \bigl\Vert \Psi ^{r+1}(g_{t}) \bigr\Vert _{X } \\ &= \frac{C_{2}(1 + \varepsilon)}{ (2^{k}\lambda)^{r+1}} \bigl\Vert A^{r+1}(g_{t}) \bigr\Vert _{X} \\ &\leq\frac{C_{1}C_{2}(1 + \varepsilon)}{ (2^{k}\lambda)^{r+1}}\frac{1}{t^{r+1}} \sup_{ 0< h\leq t} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X } \\ &= \frac{C_{1}C_{2}(1 + \varepsilon)}{(2^{k})^{r+1}}\sup_{ 0< h\leq1/\lambda} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X }. \end{aligned}$$

On the other hand, from the identity

$$q(g_{t}, 0) - g_{t} = \sum_{k=0}^{\infty}\bigl(q(g_{t}, k) - q(g_{t}, k + 1) \bigr), $$

(28), (27), (29), and (26), one has

$$\begin{aligned} \bigl\Vert \Psi^{r} \bigl(q(g_{t}, 0) - g_{t} \bigr) \bigr\Vert _{X } &\leq\sum_{k=0}^{\infty}\bigl\Vert \Psi^{r} \bigl(q(g_{t}, k) - q(g_{t}, k +1) \bigr) \bigr\Vert _{X } \\ &\leq C_{2}\sum_{k=0}^{\infty}\bigl(2^{k+1}\lambda \bigr)^{r} \bigl\Vert q(g_{t}, k) - q(g_{t}, k + 1) \bigr\Vert _{X } \\ &\leq C_{2} \sum_{k=0}^{\infty}\bigl(2^{k+1}\lambda \bigr)^{r} \bigl( \bigl\Vert q(g_{t}, k) - g_{t} \bigr\Vert _{X } + \bigl\Vert g_{t} - q(g_{t}, k + 1) \bigr\Vert _{X } \bigr) \\ &\leq2C_{1}C_{2}^{2} (1 + \varepsilon) \sup _{ 0< h\leq1\lambda} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X } \sum_{k=0}^{\infty}\bigl(2^{k+1}\lambda \bigr)^{r} \frac{1}{(2^{k})^{r+1}} \\ &= 2^{r+1}C_{1}C_{2}^{2} (1 + \varepsilon) \lambda^{r} \sup_{ 0< h\leq1\lambda} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X } \sum _{k=0}^{\infty}\frac{1}{2^{k}} \\ &= C_{3}(1 + \varepsilon)\lambda^{r} \sup _{ 0< h\leq1/\lambda} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X }. \end{aligned}$$

We also need the inequality (see (18) and (25))

$$\begin{aligned} \bigl\Vert \Psi^{r}(g_{t}) \bigr\Vert _{X } &= \bigl\Vert A^{r}(g_{t}) \bigr\Vert _{X } \leq C_{1} \frac{1}{ t^{r}}\sup_{ 0< h\leq t} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X } \\ &= C_{1}\lambda^{r} \sup_{ 0< h\leq1/\lambda} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X }. \end{aligned}$$

From the inequalities given above, for \(P \in\mathbb{P}_{\xi(\lambda )}\), we obtain

$$\begin{aligned} \bigl\Vert \Psi^{r}(P) \bigr\Vert _{X } &\leq \bigl\Vert \Psi^{r} \bigl(P - q(g_{t}, 0) \bigr) \bigr\Vert _{X} + \bigl\Vert \Psi^{r} \bigl(q(g_{t}, 0) \bigr) \bigr\Vert _{X } \\ &\leq C_{2}\lambda^{r} \bigl\Vert P - q(g_{t}, 0) \bigr\Vert _{X } + \bigl\Vert \Psi ^{r}(g_{t}) \bigr\Vert _{ X } + \bigl\Vert \Psi^{r} \bigl(g_{t} - q(g_{t}, 0) \bigr) \bigr\Vert _{X } \\ &\leq C_{1}\lambda^{r} ( \Vert P - g_{t} \Vert _{X } + \bigl\Vert g_{t} - q(g_{t}, 0) \bigr\Vert _{X_{\alpha,\beta}} + C_{4}\lambda^{r} \sup _{ 0< h\leq1/\lambda} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X } \\ &\leq C_{5} \lambda^{r} \sup_{ 0< h\leq1/\lambda} \bigl\Vert (I - C_{h,1})^{r}P \bigr\Vert _{X }. \end{aligned}$$

 □

Remark 4.2

The problem of obtaining a Nikolskii–Stechkin inequality for fractional derivatives is open.