Skip to main content

Some weak type inequalities and almost everywhere convergence of Vilenkin–Nörlund means

Abstract

We prove and discuss some new weak type \((1,1 ) \) inequalities of maximal operators of Vilenkin–Nörlund means generated by monotone coefficients. Moreover, we use these results to prove a.e. convergence of such Vilenkin–Nörlund means. As applications, both some well-known and new inequalities are pointed out.

1 Introduction

In 1947 Vilenkin [31] actually introduced a large class of compact groups (now called Vilenkin groups) and the corresponding characters. In particular, Vilenkin investigated the group \(G_{m}\), which is a direct product of the additive groups \(Z_{m_{k}}:=\{0,1,\ldots ,m_{k}-1\}\) of integers modulo \(m_{k}\), where \(m:=(m_{0},m_{1},\ldots )\) are positive integers not less than 2, and introduced the Vilenkin systems \(\{{\psi}_{j}\}_{j=0}^{\infty}\) as follows:

$$ \psi _{n}(x):=\prod_{k=0}^{\infty }r_{k}^{n_{k}} ( x ), \qquad r_{k} ( x ) :=\exp ( 2\pi ix_{k}/m_{k} ),\quad \bigl(i^{2}=-1,x\in G_{m},k\in \mathbb{N} \bigr), $$

where \(\mathbb{N}_{+}\) denotes the set of positive integers and \(\mathbb{N}:=\mathbb{N}_{+}\cup \{0\}\). In this paper we discuss bounded Vilenkin groups only, that is, \(\sup_{n\in \mathbb{N}}m_{n}<\infty \). The Vilenkin system is orthonormal and complete in \(L^{2} ( G_{m} ) \) (see [31]). Specifically, we call this system the Walsh–Paley system when \(m\equiv 2\).

It is well known (see e.g. the books [1] and [27]) that if \(f\in L^{1}(G_{m})\) and the Vilenkin series \(T (x )=\sum_{j=0}^{\infty}c_{j}\psi _{j} (x ) \) converges to f in \(L^{1}\)-norm, then

$$ c_{j}= \int _{G_{m}}f\overline{\psi }_{j}\,d\mu :=\widehat{f} ( j ), \quad j=0,1,2,\ldots, $$

where \(c_{j}\) is called the jth Vilenkin–Fourier coefficient and μ is the Haar measure on the locally compact abelian groups \(G_{m}\), which coincide with the direct product of measures \(\mu _{k} ( \{j\} ) :=1/m_{k}\) (\(j\in Z_{m_{k}}\)).

The classical theory of Hilbert spaces (for details, see e.g. the books [1, 27]) implies that if we consider the partial sums \(S_{n}\), defined by

$$ S_{n}f:=\sum_{k=0}^{n-1} \widehat{f} (k )\psi _{k}, $$

with respect to any orthonormal systems and among them to Vilenkin systems, then the inequality \(\Vert S_{n} f \Vert _{2}\leq \Vert f \Vert _{2} \) holds. It follows that for every \(f\in L^{2}\),

$$ \Vert S_{n}f-f \Vert _{2}\to 0 \quad \text{as } n\to \infty . $$

Since

$$ S_{n}f (x )= \int _{G_{m}}f (t )D_{n} (x-t )\,d\mu (t ) $$

and the Dirichlet kernels

$$ D_{n} :=\sum_{k=0}^{n-1}\psi _{k } \quad ( n\in \mathbb{N}_{+} ) $$

are not uniformly bounded in \(L^{1}(G_{m})\), the boundedness of partial sums does not hold from \(L^{1}(G_{m})\) to \(L^{1}(G_{m})\).

The analogue of Carleson’s theorem for the Walsh system was proved by Billard [4] for \(p=2\) and by Sjölin [29] for \(1 < p<\infty \), while for bounded Vilenkin systems it was proved by Gosselin [13]. In each proof, they show that the maximal operator of the partial sums is bounded on \(L^{p}(G_{m})\), i.e., there exists an absolute constant \(c_{p}\) such that

$$ \bigl\Vert S^{\ast }f \bigr\Vert _{p}\leq C_{p} \Vert f \Vert _{p},\quad \text{when }f\in L^{p}, 1< p< \infty . $$

A recent proof of almost everywhere convergence of subsequences of Walsh–Fourier series was given by Demeter [7] in 2015. Hence, if \(f\in L^{p}(G_{m})\) for \(p>1\), then

$$ S_{n}f\to f \quad \text{a.e. on } G_{m}. $$

Persson, Schipp, Tephnadze, and Weisz [22] (see also [25]) gave a new and shorter proof of almost everywhere convergence of Vilenkin–Fourier series of \(f\in L^{p}(G_{m})\), which was based on the theory of martingales.

The nth Nörlund mean \(L_{n}\) is defined by

$$\begin{aligned} L_{n}f :=&\frac{1}{l_{n}}\sum_{k=0}^{n-1} \frac{S_{k}f}{n-k}, \quad \text{where } l_{n}:=\sum _{k=1}^{n}\frac{1}{k}. \end{aligned}$$

In [9] Gát and Goginava proved some properties of the Nörlund logarithmic means of integrable functions in \(L^{1}\) norm. Moreover, in [10] they proved that weak type \((1,1)\) inequality does not hold for the maximal operator of Nörlund logarithmic means \(L^{\ast}\), defined by

$$\begin{aligned} L^{\ast}f :=&\sup_{n\in \mathbb{N}} \vert L_{n}f \vert , \end{aligned}$$

but there exists an absolute constant \(c_{p}\) such that the inequality

$$ \bigl\Vert L^{\ast }f \bigr\Vert _{p}\leq c_{p} \Vert f \Vert _{p}\quad \text{when }f\in L^{p}, p>1 $$

holds.

If we define the so-called generalized number system based on m in the following way:

$$ M_{0}:=1,\qquad M_{k+1}:=m_{k}M_{k}\quad (k\in \mathbb{N}), $$

then every \(n\in \mathbb{N}\) can be uniquely expressed as \(n=\sum_{j=0}^{\infty }n_{j}M_{j}\), where \(n_{j}\in Z_{m_{j}}\) (\(j\in \mathbb{N}\)) and only a finite number of \(n_{j}\)s differ from zero. Moreover, if we consider the following restricted maximal operator \(\widetilde{L}_{\#}^{\ast}\), defined by

$$\begin{aligned} \widetilde{L}_{\#}^{\ast}f :=&\sup_{n\in \mathbb{N}} \vert L_{M_{n}}f \vert , \end{aligned}$$

then

$$\begin{aligned} y \mu \bigl\{ \widetilde{L}_{\#}^{\ast}f>y \bigr\} \leq c \Vert f \Vert _{1},\quad f\in L^{1}(G_{m}), y>0. \end{aligned}$$

Hence, if \(f\in L^{1}(G_{m})\), then \(L_{M_{n}}f\to f\) a.e. on \(G_{m}\).

If we consider the Fejér means \(\sigma _{n}\) and Fejér kernels \(K_{n}\), defined by

$$ \sigma _{n}f:=\frac{1}{n}\sum_{k=1}^{n}S_{k}f \quad \text{and}\quad K_{n}:= \frac{1}{n}\sum_{k=0}^{n-1}D_{k}, $$

it is obvious that

$$\begin{aligned} \sigma _{n}f (x ) = (f\ast K_{n} ) (x )= \int _{G_{m}}f (t ) K_{n} (x-t )\,d\mu (t ). \end{aligned}$$

Since \(\Vert K_{n} \Vert _{1} \leq c<\infty \), we obtain that the Fejér means are bounded from the space \(L^{p}\) to the space \(L^{p}\) for \(1\leq p\leq \infty \). The a.e. convergence of Fejér means is due to Schipp [26] for Walsh series and Pál, Simon [21] (see also Simon, Weisz [28] and Weisz [28, 3234]) for bounded Vilenkin series proved that the maximal operator of Fejér means \(\sigma ^{\ast }\), defined by

$$ \sigma ^{\ast }f:=\sup_{n\in \mathbb{N}} \vert \sigma _{n}f \vert , $$

is of weak type \((1,1)\), from which the a.e. convergence follows by standard argument (see [14]). Another well-known summability method is the so-called \((C,\alpha )\)-means (denoted by \(\sigma _{n}^{\alpha}\)), which are defined by

$$ \sigma _{n}^{\alpha}f:=\frac{1}{A_{n}^{\alpha}} \sum _{k=1}^{n}A_{n-k}^{\alpha -1}S_{k}f,\qquad A_{0}^{\alpha}:=0, \qquad A_{n}^{\alpha}:= \frac{ (\alpha +1 )\cdots (\alpha +n )}{n!},\quad \alpha \neq -1,-2,\ldots. $$

It is well known that for \(\alpha =1\) this summability method coincides with the Fejér summation and for \(\alpha =0\) we just have the partial sums of the Vilenkin–Fourier series. Moreover, if we consider the maximal operator of the Cesáro means \(\sigma ^{\alpha ,\ast}\), defined by

$$\begin{aligned} \sigma ^{\alpha ,\ast}f :=&\sup_{n\in \mathbb{N}} \bigl\vert \sigma _{n}^{ \alpha }f \bigr\vert \quad \text{for } 0< \alpha \leq 1, \end{aligned}$$

then the following weak type inequality holds (for details, see [23]):

$$\begin{aligned} y \mu \bigl\{ \sigma ^{\alpha ,\ast}f>y \bigr\} \leq c \Vert f \Vert _{1},\quad f\in L^{1}(G_{m}), y>0. \end{aligned}$$

The boundedness of the maximal operator of the Cesáro means does not hold from \(L^{1}(G_{m})\) to the space \(L^{1}(G_{m})\). However,

$$ \bigl\Vert \sigma _{n}^{\alpha}f-f \bigr\Vert _{p}\rightarrow 0, \quad \text{when }n\rightarrow \infty , \bigl(f\in L^{p}(G_{m}), 1\leq p< \infty \bigr). $$

The nth Nörlund mean \(t_{n}\) for the Fourier series of f is defined by

$$ t_{n}f:=\frac{1}{Q_{n}}\sum_{k=1}^{n}q_{n-k}S_{k}f, $$

where \(\{q_{k}:k\in \mathbb{N}\}\) is a sequence of nonnegative numbers and \(Q_{n}:=\sum_{k=0}^{n-1}q_{k} \).

If we assume that \(q_{0}>0\) and \(\lim_{n\rightarrow \infty }Q_{n}=\infty \), then it is well known (see [15]) that the summability method generated by \(\{q_{k}:k\geq 0\}\) is regular if and only if \(\lim_{n\rightarrow \infty }\frac{q_{n-1}}{Q_{n}}=0 \). The representation

$$ t_{n}f (x )= \int_{G_{m}}f (t )F_{n} (x-t )\,d\mu (t ),\quad \text{where } F_{n}:= \frac{1}{Q_{n}}\sum_{k=1}^{n}q_{n-k}D_{k} $$

plays a central role in the sequel. The Nörlund means are generalizations of the Fejér, Cesàro, and Nörlund logarithmic means.

Móricz and Siddiqi [16] investigated the approximation properties of some special Nörlund means of Walsh–Fourier series of \(L^{p}\) functions in norm. Similar problems for the two-dimensional case can be found in papers by Nagy [1720] (see also [5]).

Let us define the maximal operator \(t^{\ast}\) of Nörlund means by

$$\begin{aligned} t^{\ast}f:=\sup_{n\in \mathbb{N}} \vert t_{n}f \vert , \end{aligned}$$

and if \(\{q_{k}:k\in \mathbb{N}\}\) is nonincreasing and satisfying the condition

$$ \frac{1}{Q_{n}}=O \biggl( \frac{1}{n} \biggr)\quad \text{as } n \rightarrow \infty , $$
(1)

then in [23] it was proved that the weak type inequality

$$\begin{aligned} y \mu \bigl\{ t^{\ast}f>y \bigr\} \leq c \Vert f \Vert _{1},\quad f\in L^{1}(G_{m}), y>0 \end{aligned}$$
(2)

holds. When the sequence \(\{q_{k}:k\in \mathbb{N}\}\) is nonincreasing, then the weak type \((1,1)\) inequality (2) holds for every maximal operator of Nörlund means. The boundedness of the maximal operator of the Nörlund means does not hold from \(L^{1}(G_{m})\) to the space \(L^{1}(G_{m})\). However,

$$ \Vert t_{n}f-f \Vert _{p}\rightarrow 0 \quad \text{as }n \rightarrow \infty \ \bigl(f\in L^{p}(G_{m}), 1 \leq p< \infty \bigr). $$

Moreover, if \(\{q_{k}:k\in \mathbb{N}\}\) is nondecreasing and satisfying the condition

$$ \frac{q_{n-1}}{Q_{n}}=O \biggl( \frac{1}{n} \biggr) \quad \text{as } n\rightarrow \infty , $$
(3)

or \(\{q_{k}:k\in \mathbb{N}\}\) is nonincreasing, then for any \(f\in L^{1}(G_{m})\) we have that

$$ \lim_{n\rightarrow \infty }t_{n}f(x)=f(x) $$

for all Vilenkin–Lebesgue points of f.

In this paper we investigate a wider class of Nörlund means and prove that if \(\{q_{k}:k\in \mathbb{N}\}\) is nondecreasing and satisfying the conditions

$$ \frac{1}{Q_{n}}=O \biggl(\frac{1}{n^{\alpha}} \biggr) \quad \text{and}\quad q_{n}-q_{n+1}=O \biggl(\frac{1}{n^{2-\alpha}} \biggr) \quad \text{as } n\rightarrow \infty , $$
(4)

then the weak type inequality (2) holds. In particular, from this result follows almost everywhere convergence of such Nörlund means.

The paper is organized as follows: In Sect. 3 we present and prove the main results. Moreover, in order not to disturb our discussions in this section, some preliminaries are given in Sect. 2. Also some of these results are new and of independent interest.

2 Preliminaries

Lemma 1

(see [1, 12])

Let \(n\in \mathbb{N}\). Then

$$ D_{M_{n}} (x )=\textstyle\begin{cases} M_{n}, & x\in I_{n}, \\ 0, & x\notin I_{n}. \end{cases} $$

Moreover, if \(n\in \mathbb{N}\) and \(1\leq s_{n}\leq m_{n}-1\), then

$$ D_{s_{n}M_{n}}=D_{M_{n}}\sum_{k=0}^{s_{n}-1} \psi _{kM_{n}}=D_{M_{n}} \sum_{k=0}^{s_{n}-1}r_{n}^{k} $$

and

$$ D_{n}=\psi _{n} \Biggl(\sum_{j=0}^{\infty}D_{M_{j}} \sum_{k=m_{j}-n_{j}}^{m_{j}-1}r_{j}^{k} \Biggr) \quad \textit{for } n=\sum_{i=0}^{\infty}n_{i}M_{i}, $$

where \(n=\sum_{i=0}^{\infty}n_{i}M_{i}\). We note that \(\sum_{k=m_{j}-n_{j}}^{m_{j}-1}r_{j}^{k}\equiv 0\) for all \(n_{j}=0\).

Lemma 2

(see [8])

Let \(n>t\), \(t,n\in \mathbb{N}\). Then

$$ K_{M_{n}} (x )=\textstyle\begin{cases} \frac{M_{t}}{1-r_{t} (x ) },& x\in I_{t}\backslash I_{t+1}, x-x_{t}e_{t}\in I_{n}, \\ \frac{M_{n}+1}{2}, & x\in I_{n}, \\ 0, & \textit{otherwise. } \end{cases} $$

Lemma 3

(see [3, 6, 23, 24, 30])

If \(n\geq M_{N}\) and \(\{q_{k}:k\in \mathbb{N}\}\) is a sequence of nondecreasing numbers, then there exists an absolute constant c such that

$$ \Biggl\vert \frac{1}{Q_{n}}\sum_{j=M_{N}}^{n} q_{n-j}D_{j} \Biggr\vert \leq \frac{c}{M_{N}} \Biggl\{ \sum_{j=0}^{ \vert n \vert }M_{j} \vert K_{M_{j}} \vert \Biggr\} . $$

If the sequence \(\{q_{k}:k\in \mathbb{N}\}\) is either nondecreasing and satisfying condition (3) or nonincreasing and satisfying condition (1), then the inequality

$$ \vert F_{n} \vert \leq \frac{c}{n} \Biggl\{ \sum _{j=0}^{ \vert n \vert }M_{j} \vert K_{M_{j}} \vert \Biggr\} $$

holds. On the other hand, if \(\{q_{k}:k\in \mathbb{N}\}\) is a sequence of nonincreasing numbers satisfying (4) for \(0<\alpha <1\), then there exists a constant \(c_{\alpha}\), depending only on α, such that the following inequality holds:

$$ \vert F_{n} \vert \leq \frac{c_{\alpha}}{n^{\alpha}} \Biggl\{ \sum_{j=0}^{ \vert n \vert }M_{j}^{\alpha} \vert K_{M_{j}} \vert \Biggr\} . $$
(5)

Lemma 4

(see [3, 6, 23, 24])

Let \(\{q_{k}:k\in \mathbb{N}\}\) be either a sequence of nondecreasing numbers or nonincreasing numbers satisfying condition (1) or nonincreasing numbers satisfying the conditions in (4). Then, for any \(n, N\in \mathbb{N_{+}}\),

$$\begin{aligned} & \int _{G_{m}} F_{n}(x)\,d\mu (x)=1, \\ &\sup_{n\in \mathbb{N}} \int _{G_{m}} \bigl\vert F_{n}(x) \bigr\vert \,d\mu (x)\leq c< \infty , \\ &\sup_{n\in \mathbb{N}} \int _{G_{m} \backslash I_{N}} \bigl\vert F_{n}(x) \bigr\vert \,d\mu (x)\rightarrow 0\quad \textit{as } n\rightarrow \infty , \end{aligned}$$

where

$$\begin{aligned} I_{0} ( x ) :=G_{m}, \qquad I_{n}(x):=\{y\in G_{m}| y_{0}=x_{0},\ldots,y_{n-1}=x_{n-1} \} \end{aligned}$$

for any \(x\in G_{m}\), \(n\in \mathbb{N}\).

The next lemma is very important to study problems concerning almost everywhere convergence.

Lemma 5

(see [32])

Suppose that the σ-sublinear operator V is bounded from \(L^{p_{1}}\) to \(L^{p_{1}}\) for some \(1< p_{1}\leq \infty \) and

$$ \int _{\overline{I}} \vert Vf \vert \,d\mu \leq C \Vert f \Vert _{1} $$

for \(f\in L^{1}\) and Vilenkin interval I, which satisfy

$$ \operatorname{supp} f\subset I,\qquad \int _{G_{m}}f\,d\mu =0. $$
(6)

Then the operator V is of weak type \(( 1,1 )\), i.e., the following inequality holds:

$$ \sup_{y>0}y\mu \bigl( \{ Vf>y \} \bigr) \leq \Vert f \Vert _{1}. $$

Lemma 6

(see [14])

Let

$$ T,T_{n}:L^{p} ( G_{m} )\rightarrow L^{p} ( G_{m} ) $$

be sublinear operators for some \(1\leq p<\infty \) with T bounded and

$$ T_{n}f\rightarrow Tf \quad \textit{a.e. on } G_{m} \textit{ as } n \rightarrow \infty , $$

for each \(f\in X_{0}\), where \(X_{0}\) is dense in \(L^{p} ( G_{m} )\). Set

$$ T^{\ast }f:=\sup_{n\in \mathbb{N}} \vert T_{n}f \vert , \quad f\in X. $$

If there is a constant \(C>0\), independent of f and n, such that the weak type inequalities

$$ y^{p} \mu \bigl( \bigl\{ \vert Tf \vert >y \bigr\} \bigr) \leq C \Vert f \Vert ^{p}_{X} $$

and

$$ y^{p}\mu \bigl( \bigl\{ T^{\ast }f>y \bigr\} \bigr) \leq C \Vert f \Vert ^{p}_{X} $$

hold for all \(y>0\) and \(f\in L^{p} ( G_{m} )\), then

$$ Tf=\lim_{n\rightarrow \infty }T_{n}f\quad \textit{a.e. on } G_{m} $$

for every \(f\in L^{p} ( G_{m} )\).

Next we prove a new lemma of independent interest, which is very important to prove almost everywhere convergence of Nörlund means generated by nondecreasing sequences \(\{q_{k}:k\in \mathbb{N}\}\).

Lemma 7

Let \(n\in \mathbb{N}\) and \(\{q_{k}:k\in \mathbb{N}\}\) be a sequence of nondecreasing numbers. Then

$$\begin{aligned}& \int _{\overline{I_{N}}}\sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \,d\mu (x)\leq c< \infty , \end{aligned}$$

where c is an absolute constant.

Proof

If we define

$$\begin{aligned} I_{N}^{k,l}:=\textstyle\begin{cases} I_{N}(0,\ldots ,0,x_{k}\neq 0,0,\ldots,0,x_{l}\neq 0,x_{l+1},\ldots ,x_{N-1}, \ldots ), \\ \quad \text{for } k< l< N, \\ I_{N}(0,\ldots ,0,x_{k}\neq 0,x_{k+1}=0,\ldots ,x_{N-1}=0,x_{N}, \ldots ), \\ \quad \text{for } l=N. \end{cases}\displaystyle \end{aligned}$$

then we can decompose \(\overline{I_{N}}:=G_{m} \backslash I_{N}\) as

$$ G_{m} \backslash I_{N}=\bigcup _{s=0}^{N-1}I_{s} \backslash I_{s+1}= \Biggl(\bigcup_{k=0}^{N-2} \bigcup_{l=k+1}^{N-1}I_{N}^{k,l} \Biggr)\cup \Biggl( \bigcup^{N-1}_{k=0}I_{N}^{k,N} \Biggr). $$
(7)

Let \(n>M_{N}\) and

$$ x\in I_{N}^{k,l},\quad k=0,\dots ,N-2, l=k+1,\dots ,N-1. $$

By using Lemma 3, we get that

$$\begin{aligned} \Biggl\vert \frac{1}{Q_{n}}\sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \leq &\frac{c}{M_{N}}\sum_{i=0}^{l}M_{i} \bigl\vert K_{M_{i}} (x ) \bigr\vert \\ \leq &\frac{c}{M_{N}}\sum_{i=0}^{l}M_{i} M_{k} \\ \leq& \frac{cM_{l}M_{k}}{M_{N}} \end{aligned}$$

so that

$$\begin{aligned} \sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \leq &\frac{c}{M_{N}}\sum_{i=0}^{ \vert n \vert }M_{i} \bigl\vert K_{M_{i}} (x ) \bigr\vert \\ \leq & \frac{cM_{l}M_{k}}{M_{N}}. \end{aligned}$$
(8)

Let \(n>M_{N}\) and \(x\in I_{N}^{k,N}\). By using Lemma 1, we can conclude that

$$ \bigl\vert D_{n}(x) \bigr\vert \leq cM_{k} $$

and

$$\begin{aligned} \Biggl\vert \frac{1}{Q_{n}}\sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \leq &\frac{c}{Q_{n}} \sum^{n }_{j=M_{N}}q_{n-j}M_{k} \\ \leq &\frac{cQ_{n-M_{N}}}{Q_{n}}M_{k}\leq cM_{k}, \end{aligned}$$

so that

$$\begin{aligned} \sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \leq cM_{k}. \end{aligned}$$
(9)

Hence, if we apply estimates (8) and (9), then we get that

$$\begin{aligned}& \int _{\overline{I_{N}}}\sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \,d\mu \\& \quad = \sum_{k=0}^{N-2} \sum _{l=k+1}^{N-1} \sum_{x_{j}=0,j\in \{l+1,\ldots,N-1 \}}^{m_{j-1}} \int _{I_{N}^{k,l}}\sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}}\sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \,d\mu \\& \qquad {} + \sum_{k=0}^{N-1} \int _{I_{N}^{k,N}}\sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}}\sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \,d\mu \\& \quad \leq c\sum_{k=0}^{N-2} \sum _{l=k+1}^{N-1}\frac{m_{l+1}\cdots m_{N-1}}{M_{N}} \frac{M_{l}M_{k}}{M_{N}}+c\sum_{k=0}^{N-1} \frac{M_{k}}{M_{N}} \\& \quad \leq c \sum_{k=0}^{N-2} \frac{(N-k)M_{k}}{M_{N}}+c< C< \infty . \end{aligned}$$

The proof is complete. □

We also need the following new lemmas.

Lemma 8

Let \(\{q_{k}:k\in \mathbb{N}\}\) be a sequence of nonincreasing numbers satisfying condition (1). Then there exists an absolute constant c such that

$$\begin{aligned}& \int _{\overline{I_{N}}}\sup_{n>M_{N}} \vert F_{n} \vert \,d \mu \leq c< \infty . \end{aligned}$$

Proof

The proof is analogous to that of Lemma 7. Hence, we leave out the details. □

Lemma 9

Let \(\{q_{k}:k\in \mathbb{N}\}\) be a sequence of nondecreasing numbers satisfying condition (3). Then there exists an absolute constant c such that

$$\begin{aligned}& \int _{\overline{I_{N}}}\sup_{n>M_{N}} \vert F_{n} \vert \,d \mu \leq c< \infty . \end{aligned}$$

Proof

Also in this case the proof is analogous to that of Lemma 7, so we leave out the details. □

Finally, we prove the following new estimate of independent interest.

Lemma 10

Let \(n\in \mathbb{N}\) and \(\{q_{k}:k\in \mathbb{N}\}\) be a sequence of nonincreasing numbers satisfying the conditions in (4). Then there exists an absolute constant c such that

$$\begin{aligned} \int _{\overline{I_{N}}} \sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \,d\mu (x)\leq c< \infty . \end{aligned}$$
(10)

Proof

Let \(n>M_{N}\) and \(x\in I_{N}^{k,l}\), \(k=0,\dots ,N-2\), \(l=k+1,\dots ,N-1\). By combining Lemma 2 and (5) in Lemma 3, we get that

$$\begin{aligned} & \Biggl\vert \frac{1}{Q_{n}}\sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \leq \frac{cM_{l}^{\alpha }M_{k}}{M_{N}^{\alpha}}, \end{aligned}$$

so that

$$\begin{aligned} &\sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n} q_{n-j}D_{j} (x ) \Biggr\vert \leq \frac{cM_{l}^{\alpha }M_{k}}{M_{N}^{\alpha}}. \end{aligned}$$
(11)

Let \(n>M_{N}\) and \(x\in I_{N}^{k,N}\). By using Lemma 1, we can conclude that

$$\begin{aligned} \Biggl\vert \frac{1}{Q_{n}}\sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \leq \frac{c}{Q_{n}} \sum^{n }_{j=M_{N}}q_{n-j}M_{k} \leq cM_{k}, \end{aligned}$$

so that

$$\begin{aligned} \sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n}q_{n-j}D_{j} (x ) \Biggr\vert \leq cM_{k}. \end{aligned}$$
(12)

By combining (7), (11), and (12), we can conclude that

$$\begin{aligned}& \int _{\overline{I_{N}}}\sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{n}q_{n-j}D_{j} \Biggr\vert \,d\mu \\& \quad = \sum_{k=0}^{N-2} \sum _{l=k+1}^{N-1} \sum_{x_{j}=0,j\in \{l+1,\ldots,N-1 \}}^{m_{j-1}} \int _{I_{N}^{k,l}}\sup_{n>M_{N}} \vert F_{n} \vert \,d\mu \\& \qquad {} +\sum_{k=0}^{N-1} \int _{I_{N}^{k,N}}\sup_{n>M_{N}} \vert F_{n} \vert \,d\mu \\& \quad \leq c\sum_{k=0}^{N-2} \sum _{l=k+1}^{N-1}\frac{m_{l+1}\cdots m_{N-1}}{M_{N}} \frac{M_{l}^{\alpha}M_{k}}{M^{\alpha}_{N}}+c \sum_{k=0}^{N-1} \frac{M_{k}}{M_{N}} \\& \quad \leq c\sum_{k=0}^{N-2} \sum _{l=k+1}^{N-1} \frac{M_{l}^{\alpha -1}M_{k}}{M^{\alpha}_{N}}+c \sum _{k=0}^{N-1}\frac{M_{k}}{M_{N}} \\& \quad \leq c\sum_{l=k+1}^{N-1} \frac{M^{\alpha}_{k}}{M^{\alpha}_{N}}+c \\& \quad < C< \infty , \end{aligned}$$

so (10) holds and the proof is complete. □

3 The main results

Our first main result reads as follows.

Theorem 1

Let \(t_{n}\) be the Nörlund means and \(F_{n}\) be the corresponding Nörlund kernels such that

$$ \int_{\overline{I_{N}}}\sup_{n>M_{N}} \Biggl\vert \frac{1}{Q_{n}}\sum_{k=M_{N}+1}^{n}q_{n-k}D_{k} (x ) \Biggr\vert \,d\mu (x )< c< \infty . $$

If the maximal operator \(t^{*}\) of Nörlund means is bounded from \(L^{p_{1}}\) to \(L^{p_{1}}\) for some \(1< p_{1}\leq \infty \), then the operator \(t^{*}\) is of weak type \(( 1,1 ) \), i.e., for all \(f\in L^{1}(G_{m})\), the following weak type inequality holds:

$$ \sup_{y>0}y\mu \bigl\{ t^{*}f>y \bigr\} \leq \Vert f \Vert _{1}. $$

Proof

In view of Lemma 5 we obtain that the proof is complete if we prove that

$$ \int _{\overline{I}} \bigl\vert t^{*}f(x) \bigr\vert \,d\mu (x) \leq c \Vert f \Vert _{1} $$
(13)

for every function f, which satisfies the conditions in (6), where I denotes the support of the function f.

Without loss of generality we may assume that f is a function with support I and \(\mu ( I ) =M_{N}\). We may also assume that \(I=I_{N}\). It is easy to see that

$$ t_{n}f =0\quad \text{when } n\leq M_{N}. $$

Therefore, we can suppose that \(n>M_{N}\). Moreover,

$$ S_{n}f =0 \quad \text{for } n\leq M_{N}, $$

so that

$$ \frac{1}{Q_{n}} \Biggl(\sum_{k=0}^{M_{N}}q_{n-k}S_{k}f (x ) \Biggr)=0, $$

which implies that

$$ \int_{I_{N}} \frac{1}{Q_{n}} \Biggl( \sum _{k=0}^{M_{n}}q_{n-k}D_{k} (x-t ) \Biggr)f(t)\,d\mu (t )=0. $$

Hence,

$$\begin{aligned}& \bigl\vert t^{*}f(x) \bigr\vert \\& \quad \leq \sup_{n>M_{N}} \Biggl\vert \int_{I_{N}} \frac{1}{Q_{n}} \Biggl(\sum _{k=0}^{M_{N}}q_{n-k}D_{k} (x-t ) \Biggr)f(t)\,d\mu (t ) \Biggr\vert \\& \qquad {} + \sup_{n>M_{N}} \Biggl\vert \int_{I_{N}} \frac{1}{Q_{n}} \Biggl(\sum _{k=M_{N}+1}^{n}q_{n-k}D_{k} (x-t ) \Biggr)f(t)\,d\mu (t ) \Biggr\vert \\& \quad = \sup_{n>M_{N}} \Biggl\vert \int_{I_{N}} \frac{1}{Q_{n}} \Biggl(\sum _{k=M_{N}+1}^{n}q_{n-k}D_{k} (x-t ) \Biggr)f(t)\,d\mu (t ) \Biggr\vert . \end{aligned}$$
(14)

Let \(t\in I_{N}\) and \(x\in \overline{I_{N}}\). Then \(x-t\in \overline{I_{N}}\) and (14) implies that

$$\begin{aligned}& \int _{\overline{I_{N}}} \bigl\vert t^{*}f(x) \bigr\vert \,d \mu (x) \\& \quad \leq \int _{\overline{I_{N}}}{\sup_{n>M_{N}}} \int_{I_{N}} \Biggl\vert \frac{1}{Q_{n}} \Biggl( \sum _{k=M_{N}+1}^{n}q_{n-k}D_{k} (x-t ) \Biggr)f(t) \Biggr\vert \,d\mu (t )\,d\mu (x ) \\& \quad \leq \int _{\overline{I_{N}}} \int_{I_{N}}{\sup_{n>M_{N}}} \Biggl\vert \frac{1}{Q_{n}} \Biggl( \sum_{k=M_{N}+1}^{n}q_{n-k}D_{k} (x-t ) \Biggr)f(t) \Biggr\vert \,d\mu (t )\,d\mu (x ) \\& \quad \leq \int _{I_{N}} \int_{\overline{I_{N}}}{\sup_{n>M_{N}}} \Biggl\vert \frac{1}{Q_{n}} \Biggl( \sum_{k=M_{N}+1}^{n}q_{n-k}D_{k} (x-t ) \Biggr)f(t) \Biggr\vert \,d\mu (x )\,d\mu (t ) \\& \quad \leq \int _{I_{N}} \int_{\overline{I_{N}}}{\sup_{n>M_{N}}} \Biggl\vert \frac{1}{Q_{n}} \Biggl( \sum_{k=M_{N}+1}^{n}q_{n-k}D_{k} (x ) \Biggr)f(t) \Biggr\vert \,d\mu (x )\,d\mu (t ) \\& \quad \leq \int _{I_{N}} \bigl\vert f(t) \bigr\vert \,d\mu (t ) \int_{\overline{I_{N}}}{\sup_{n>M_{N}}} \Biggl\vert \frac{1}{Q_{n}} \Biggl( \sum_{k=M_{N}+1}^{n}q_{n-k}D_{k} (x ) \Biggr) \Biggr\vert \,d\mu (x ) \\& \quad = \Vert f \Vert _{1} \int_{\overline{I_{N}}}{ \sup_{n>M_{N}}} \Biggl\vert \frac{1}{Q_{n}} \Biggl( \sum_{k=M_{N}+1}^{n}q_{n-k}D_{k} (x ) \Biggr) \Biggr\vert \,d\mu (x ) \\& \quad \leq c \Vert f \Vert _{1}. \end{aligned}$$

Thus (13) holds, so the proof is complete. □

By using the same technique of proof, we obtain in a similar way the following result.

Theorem 2

Let \(t_{n}\) be Nörlund means and \(F_{n}\) be the corresponding Nörlund kernels such that

$$ \int_{\overline{I_{N}}}\sup_{n>M_{N}} \bigl\vert F_{n} (t ) \bigr\vert \,d\mu (t )< c< \infty . $$

If the maximal operator \(t^{*}\) of the Nörlund means is bounded from \(L^{p_{1}}\) to \(L^{p_{1}}\) for some \(1< p_{1}\leq \infty \), then the operator \(t^{*}\) is of weak type \(( 1,1 ) \), i.e., the following weak type inequality

$$ \sup_{y>0}y\mu \bigl\{ t^{*}f>y \bigr\} \leq \Vert f \Vert _{1} $$

holds for all \(f\in L^{1}(G_{m})\).

Next, we present a new related result concerning almost everywhere convergence of some summability methods. The study of almost everywhere convergence is one of the most difficult topics in Fourier analysis.

Theorem 3

Let \(f\in L^{1}(G_{m})\) and \(t_{n}\) be the regular Nörlund means with nondecreasing sequences \(\{q_{k}:k\in \mathbb{N}\}\). Then

$$ t_{n}f\rightarrow f\quad \textit{a.e. as } n\rightarrow \infty . $$

Proof

Since

$$ S_{n}P=P\quad \text{for every }P\in \mathcal{P} $$

according to the regularity of Nörlund means with nondecreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\), we obtain that

$$ t_{n}P\rightarrow P \quad \text{a.e. as } n \rightarrow \infty , $$

where \(P\in \mathcal{P}\) is dense in the space \(L^{1}\).

On the other hand, by combining Lemma 4, Lemma 7, and Theorem 1, we obtain that the maximal operator \(t^{\ast}\) of the Nörlund means with nondecreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\) is bounded from the space \(L^{1}\) to the space \(weak-L^{1}\), that is, the following weak type inequality holds:

$$ \sup_{y>0}y \mu \bigl\{ x\in G_{m}: \bigl\vert t^{\ast} f (x ) \bigr\vert >y \bigr\} \leq \Vert f \Vert _{1}. $$

Hence, according to Lemma 6, we obtain the claimed almost everywhere convergence of Nörlund means with nondecreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\):

$$ t_{n}f\rightarrow f\quad \text{a.e. as }n\rightarrow \infty . $$

The proof is complete. □

Theorem 4

Let \(f\in L^{1}\) and \(t_{n}\) be the Nörlund means with nondecreasing sequence \(\{q_{k}:k\geq 0\}\) satisfying the conditions in (3). Then

$$ t_{n}f\rightarrow f,\quad \textit{a.e., as }n\rightarrow \infty . $$

Proof

The proof is similar to the proof of Theorem 3 if we instead apply Lemma 4, Lemma 9, and Theorem 1, so we omit the details. □

Next we consider almost everywhere convergence of Nörlund means with nonincreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\).

Theorem 5

Let \(f\in L^{1}\) and \(t_{n}\) be the Nörlund means with nonincreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\) satisfying condition (1). Then

$$ t_{n}f\rightarrow f\quad \textit{a.e. as } n\rightarrow \infty . $$

Proof

The proof is quite analogous to that of Theorem 3 if we apply Lemma 4, Lemma 8, and Theorem 1, so we omit the details. □

Theorem 6

Let \(f\in L^{1}\) and \(t_{n}\) be Nörlund means with nonincreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\) satisfying the conditions in (4). Then

$$ t_{n}f\rightarrow f\quad \textit{a.e. as } n\rightarrow \infty . $$

Proof

The proof is similar to the proof of Theorem 3 if we instead apply Lemma 4, Lemma 10, and Theorem 1, so we omit the details. □

Theorem 7

Let \(f\in L^{1}\) and \(t_{n}\) be Nörlund means with nonincreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\). Then

$$ t_{M_{n}}f\rightarrow f\quad \textit{a.e. as }n\rightarrow \infty . $$

Proof

If we apply the fact that (see [810], and [25])

$$\begin{aligned} F_{M_{n}}(x)=D_{M_{n}}(x)-\psi _{M_{n}-1}(x) \overline{F^{-1}}_{M_{n}}(x), \end{aligned}$$

we can prove that if \(\{q_{k}:k\in \mathbb{N}\}\) is a sequence of nonincreasing numbers, then, for any \(N\in \mathbb{N_{+}}\),

$$\begin{aligned} & \int _{G_{m}} F_{M_{n}}(x)\,d\mu (x)=1, \\ &\sup_{n\in \mathbb{N}} \int _{G_{m}} \bigl\vert F_{M_{n}}(x) \bigr\vert \,d\mu (x)\leq c< \infty , \\ &\sup_{n\in \mathbb{N}} \int _{G_{m} \backslash I_{N}} \bigl\vert F_{M_{n}}(x) \bigr\vert \,d\mu (x)\rightarrow 0 \quad \text{as } n\rightarrow \infty , \end{aligned}$$

and

$$\begin{aligned} & \int _{\overline{I_{N}}}\sup_{n>N} \Biggl\vert \frac{1}{Q_{n}} \sum_{j=M_{N}}^{M_{n}}q_{j}D_{n-j} (x ) \Biggr\vert \,d\mu (x)\leq c< \infty , \end{aligned}$$

and also in this case the proof is absolutely analogous to that of Theorem 3, so we can omit the details. □

A number of special cases of our results are of particular interest and give both well-known and new information. We just give the following examples of such corollaries.

In particular, since \(\sigma _{n}\) and \(\sigma _{n}^{\alpha }\) are regular Nörlund means with nondecreasing sequence \(\{q_{k}:k\in \mathbb{N}\}\), we have the following consequences of our Theorems:

Corollary 1

(see [23] and [32])

Let \(f\in L^{1}\). Then

$$\begin{aligned} \sigma _{n}f \rightarrow &f,\quad \textit{a.e., as }n \rightarrow \infty \end{aligned}$$

and

$$\begin{aligned} \sigma _{n}^{\alpha }f \rightarrow &f,\quad \textit{a.e., as }n\rightarrow \infty ,\textit{when }0< \alpha < 1. \end{aligned}$$

Corollary 2

(see [2] and [11])

Let \(f\in L^{1}\). Then

$$ L_{M_{n}}f\rightarrow f\quad \textit{a.e. as }n\rightarrow \infty . $$

We also give the following examples of new consequences.

Corollary 3

Let \(f\in L^{1}\) and the summability method \(V_{n}^{\alpha}\) be defined by

$$ V_{n}^{\alpha}f:=\frac{1}{Q_{n}}\sum _{k=1}^{n} (n-k-1 )^{\alpha -1}S_{k}f. $$

Then

$$ V_{n}^{\alpha}f\rightarrow f\quad \textit{a.e. as }n \rightarrow \infty ,\textit{as } 0< \alpha < 1. $$

Proof

Since \(V_{n}^{\alpha }\) are Nörlund means with nonincreasing sequences \(\{q_{k}:k\in \mathbb{N}\}\) satisfying the conditions in (4). Hence, the proof is complete by just using Theorem 6. □

Corollary 4

Let \(f\in L^{1}\) and the summability method \(\beta _{n}^{\alpha}\) be defined by

$$ \beta _{n}^{\alpha}f:=\frac{1}{Q_{n}}\sum _{k=1}^{n}\log ^{\alpha} ( n-k-1 )S_{k}f. $$

Then

$$ \beta _{n}^{\alpha }f\rightarrow f \quad \textit{a.e. as } n \rightarrow \infty . $$

Proof

We note that \(\beta _{n}^{\alpha}\) are Nörlund means with nondecreasing sequences \(\{q_{k}:k\in \mathbb{N}\}\). Hence, the proof is complete by just using Theorem 3. □

Corollary 5

Let \(f\in L^{1}\) and \(B_{n}\) be the Nörlund means with monotone and bounded sequence \(\{ q_{k}:k\in \mathbb{N} \} \). Then

$$ B_{n}f\rightarrow f \quad \textit{a.e. as } n\rightarrow \infty . $$

Proof

The proof follows from Theorems 4 and 5. □

Corollary 6

Let \(f\in L^{1}\) and the summability method \(U_{n}^{\alpha}\) be defined by

$$ U_{n}^{\alpha }f:=\frac{1}{Q_{n}}\sum _{k=1}^{n} \frac{S_{k}f}{ (n-k-3 )\ln ^{\alpha} (n-k-3 )}. $$

Then

$$ U^{\alpha}_{M_{n}}f\rightarrow f\quad \textit{a.e., as }n \rightarrow \infty . $$

Proof

Obviously, \(U^{\alpha}_{n}\) are regular Nörlund means with nonincreasing sequences \(\{q_{k}:k\in \mathbb{N}\}\), the proof follows from Theorem 7. □

Availability of data and materials

Not applicable.

References

  1. Agaev, G., Vilenkin, N., Dzahafarly, G., Rubinstein, A.: Multiplicative Systems of Functions and Harmonic Analysis on Zero-Dimensional Groups. Ehim, Baku (1981)

    Google Scholar 

  2. Baramidze, D., Persson, L.-E., Singh, H., Tephnadze, G.: Some new results and inequalities for subsequences of Nörlund logarithmic means of Walsh-Fourier series. J. Inequal. Appl. (2022). https://doi.org/10.1186/s13660-022-02765-5

    Article  MATH  Google Scholar 

  3. Baramidze, L., Persson, L.E., Tephnadze, G., Wall, P.: Sharp \(H_{p}-L_{p}\) type inequalities of weighted maximal operators of Vilenkin-Nörlund means and its applications. J. Inequal. Appl. (2016). https://doi.org/10.1186/s13660-016-1182-1

    Article  MATH  Google Scholar 

  4. Billard, P.: Sur la convergence presque partout des séries de Fourier-Walsh des fonctions de l’espace \(L^{2} (0, 1)\). Stud. Math. 28, 363–388 (1967)

    Article  MATH  Google Scholar 

  5. Blahota, I., Nagy, K., Tephnadze, G.: Approximation by Marcinkiewicz Θ-means of double Walsh-Fourier series. Math. Inequal. Appl. 22(3), 837–853 (2019)

    MathSciNet  MATH  Google Scholar 

  6. Blahota, I., Persson, L.E., Tephnadze, G.: On the Nörlund means of Vilenkin-Fourier series. Czechoslov. Math. J. 65(4), 983–1002 (2015)

    Article  MATH  Google Scholar 

  7. Demeter, C.: A guide to Carleson’s theorem. Rocky Mt. J. Math. 45(1), 169–212 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  8. Gát, G.: Investigations of certain operators with respect to the Vilenkin system. Acta Math. Hung. 61, 131–149 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  9. Gát, G., Goginava, U.: Uniform and L-convergence of logarithmic means of Walsh-Fourier series. Acta Math. Sin. Engl. Ser. 22(2), 497–506 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  10. Gát, G., Goginava, U.: On the divergence of Nörlund logarithmic means of Walsh-Fourier series. Acta Math. Sin. Engl. Ser. 25(6), 903–916 (2009). (English summary)

    Article  MathSciNet  MATH  Google Scholar 

  11. Goginava, U.: Almost everywhere convergence of subsequences of logarithmic means of Walsh-Fourier series. Acta Math. Acad. Paedagog. Nyházi. 21, 169–175 (2005)

    MathSciNet  MATH  Google Scholar 

  12. Golubov, B.I., Efimov, A.V., Skvortsov, V.A.: Walsh Series and Transforms. Mathematics and Its Applications, vol. 64. Nauka, Moscow (1987). (Russian), English transl, Kluwer Academic, Dordrecht, 1991

    MATH  Google Scholar 

  13. Gosselin, J.: Almost everywhere convergence of Vilenkin-Fourier series. Trans. Am. Math. Soc. 185, 345–370 (1973)

    Article  MathSciNet  MATH  Google Scholar 

  14. Marcinkiewicz, I., Zygmund, A.: On the summability of double Fourier series. Fundam. Math. 32, 112–132 (1939)

    Article  MATH  Google Scholar 

  15. Moore, C.N.: Summable Series and Convergence Factors, Summable Series and Convergence Factors. Dover, New York (1966)

    MATH  Google Scholar 

  16. Móricz, F., Siddiqi, A.: Approximation by Nörlund means of Walsh-Fourier series. J. Approx. Theory 70(3), 375–389 (1992). (English summary)

    Article  MathSciNet  MATH  Google Scholar 

  17. Nagy, K.: Approximation by Nörlund means of quadratical partial sums of double Walsh-Fourier series. Anal. Math. 36(4), 299–319 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  18. Nagy, K.: Approximation by Cesáro means of negative order of Walsh-Kaczmarz-Fourier series. East J. Approx. 16(3), 297–311 (2010)

    MathSciNet  MATH  Google Scholar 

  19. Nagy, K.: Approximation by Nörlund means of Walsh-Kaczmarz-Fourier series. Georgian Math. J. 18(1), 147–162 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  20. Nagy, K.: Approximation by Nörlund means of double Walsh-Fourier series for Lipschitz functions. Math. Inequal. Appl. 15(2), 301–322 (2012)

    MathSciNet  MATH  Google Scholar 

  21. Pál, J., Simon, P.: On a generalization of the concept of derivative. Acta Math. Acad. Sci. Hung. 29(1–2), 155–164 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  22. Persson, L.-E., Schipp, F., Tephnadze, G., Weisz, F.: An analogy of the Carleson-Hunt theorem with respect to Vilenkin systems. J. Fourier Anal. Appl. (2022). https://doi.org/10.1007/s00041-022-09938-2

    Article  MathSciNet  MATH  Google Scholar 

  23. Persson, L.E., Tephnadze, G., Wall, P.: On the maximal operators of Vilenkin-Nörlund means. J. Fourier Anal. Appl. 21(1), 76–94 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  24. Persson, L.E., Tephnadze, G., Wall, P.: Some new \((H_{p},L_{p})\) type inequalities of maximal operators of Vilenkin-Nörlund means with non-decreasing coefficients. J. Math. Inequal. 9(4), 1055–1069 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  25. Persson, L.E., Tephnadze, G., Weisz, F.: Martingale Hardy Spaces and Summability of One-Dimensional Vilenkin-Fourier Series. Springer, Basel (2022)

    Book  MATH  Google Scholar 

  26. Schipp, F., Certain, F.: Rearrangements of series in the Walsh system. Mat. Zametki 18(2), 193–201 (1975). (Russian)

    MathSciNet  Google Scholar 

  27. Schipp, F., Wade, W.R., Simon, P., Pál, J.: Walsh Series. An Introduction to Dyadic Harmonic Analysis. Hilger, Bristol (1990)

    MATH  Google Scholar 

  28. Simon, P., Weisz, F.: Weak inequalities for Cesáro and Riesz summability of Walsh-Fourier series. J. Approx. Theory 151(1), 1–19 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  29. Sjölin, P.: An inequality of Paley and convergence a.e. of Walsh-Fourier series. Ark. Mat. 7, 551–570 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  30. Tephnadze, G.: Martingale Hardy Spaces and Summability of the One Dimensional Vilenkin-Fourier Series. PhD thesis, Department of Engineering Sciences and Mathematics, LuleåUniversity of Technology (2015). ISSN 1402-1544

  31. Vilenkin, N.Y.: On a class of complete orthonormal systems. Am. Math. Soc. Transl. 28(2), 1–35 (1963)

    MathSciNet  MATH  Google Scholar 

  32. Weisz, F.: Martingale Hardy Spaces and Their Applications in Fourier Analysis. Lecture Notes in Mathematics, vol. 1568. Springer, Berlin (1994)

    MATH  Google Scholar 

  33. Weisz, F.: Cesáro summability of one- and two-dimensional Walsh-Fourier series. Anal. Math. 22(3), 229–242 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  34. Weisz, F.: \(( C,\alpha ) \) summability of Walsh-Fourier series, summability of Walsh-Fourier series. Anal. Math. 27(2), 141–155 (2001)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The work of Davit Baramidze was supported by Shota Rustaveli National Science Foundation grant PHDF-21-1702. The publication charges for this article have been funded by a grant from the publication fund of UiT The Arctic University of Norway. We thank both careful referees for some generous advice, which has improved the final version of this paper.

Funding

The publication charges for this manuscript are funded by UiT The Arctic University of Norway. Open Access funding provided by UiT The Arctic University of Norway (incl University Hospital of North Norway).

Author information

Authors and Affiliations

Authors

Contributions

DB and NN gave the idea and initiated the writing of this paper. LEP and GT followed up this with some complementary ideas. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Lars-Erik Persson.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

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

Rights and permissions

Open Access This article is 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

Baramidze, D., Nadirashvili, N., Persson, LE. et al. Some weak type inequalities and almost everywhere convergence of Vilenkin–Nörlund means. J Inequal Appl 2023, 66 (2023). https://doi.org/10.1186/s13660-023-02970-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13660-023-02970-w

MSC

Keywords