中国科学院大学学报  2024, Vol. 41 Issue (6): 721-727   PDF    
Continuity of truncated Hardy-Littlewood maximal function
WANG Yidong, WU Jia, YAN Dunyan     
School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
Abstract: This paper focuses on the continuity of the truncated Hardy-Littlewood maximal function. We first show that the truncated Hardy-Littlewood maximal function is lower semi-continuous. Then by investigating the behavior of the truncated Hardy-Littlewood maximal function when the truncated parameter γ changes, we obtain an equivalent condition of the continuity of the truncated Hardy-Littlewood maximal function.
Keywords: Hardy-Littlewood maximal function    truncation    continuity    truncated parameter    
截断Hardy-Littlewood极大函数的连续性
王一栋, 武嘉, 燕敦验     
中国科学院大学数学科学学院, 北京 100049
摘要: 主要研究截断Hardy-Littlewood极大函数的连续性。首先证明截断Hardy-Littlewood极大函数是下半连续的, 随后通过研究截断参数γ的变化对截断Hardy-Littlewood极大函数的影响, 得到截断Hardy-Littlewood极大函数连续性的一个等价条件。
关键词: Hardy-Littlewood极大函数    截断    连续性    截断参数    

Let ƒ be a complex-valued measurable function that is locally integrable, denoted by $f \in L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$. The non-centered Hardy-Littlewood maximal function and the centered Hardy-Littlewood maximal function Mcƒ are defined by

$ M f(x)=\sup _{B_r \ni x} \frac{1}{\left|B_r\right|} \int_{B_r}|f(y)| \mathrm{d} y, $

and

$ M^{\mathrm{c}} f(x)=\sup _{r>0} \frac{1}{\left|B_r(x)\right|} \int_{B_r(x)}|f(y)| \mathrm{d} y . $

Here Br is a ball with radius r, and Br(x) is a ball with radius r and centered at the point x. For every subset $A \subset \mathbb{R}^n$, we use |A| to denote its Lebesgue measure of the set A.

The Hardy-Littlewood maximal function was introduced by Hardy and Littlewood[1] for n=1, and by Wiener[2] for n≥2. It is well-known that the Hardy-Littlewood maximal function plays an important role in various fields of analysis.

The operator M: $f \mapsto M f$ generated by the Hardy-Littlewood maximal function is named as Hardy-Littlewood maximal operator. It maps Lp($\mathbb{R}^n$) to itself for 1 < p ≤ ∞, and L1($\mathbb{R}^n$) to L1, ∞($\mathbb{R}^n$).M is a classical mean operater, and usually used to control some other important operators in harmonic analysis. More properties of Hardy-Littlewood maximal functions can be found in Refs.[3-4].

For some real number γ$\mathbb{R}$ satisfying 0 < γ < ∞, the truncated non-centered Hardy-Littlewood maximal function Mγƒ and the truncated centered Hardy-Littlewood maximal function $M_γ^{\mathrm{c}} f$ can be defined by

$ M_\gamma f(x)=\sup _{\substack{B_r \ni x \\ 0<r<\gamma}} \frac{1}{\left|B_r\right|} \int_{B_r}|f(y)| \mathrm{d} y, $ (1)

and

$ M_\gamma^{\mathrm{c}} f(x)=\sup _{0<r<\gamma} \frac{1}{\left|B_r(x)\right|} \int_{B_r(x)}|f(y)| \mathrm{d} y . $

Although both operators Mγ and $M_\gamma^c$ are obviously controlled by M and Mc, they have the same norms. Wei et al.[5] proved the (p, p) norm of Mγ equals to that of M for 1 < p≤∞ and 0 < γ < ∞. For the case of p=1, they also proved that the weak (1, 1) norm of Mγ equals to that of M for 0 < γ < ∞. As for $M_\gamma^{\mathrm{c}}$, the result is true as well.

The regularity properties of and Mcƒ are of significant interest due to their successful applications in the study of Sobolev functions and partial differential equations. A deal of work has been done since Kinnunen[6] proved that Mc is bounded in the Sobolev space W1, p($\mathbb{R}^n$) for 1 < p≤∞. Aldaz and Pérez[7] proved that for a nondegenerate interval I if ƒ: I$\mathbb{R}$ is of bounded variation, then is absolutely continuous.

In 2020, Wu and Yan[8] obtained a characterization of continuity of and Mcƒ, and investigated the relationships between the continuity of and the continuity or bounded variation of ƒ.

Theorem A   Let ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right) $ and (a) < ∞. The non-centered Hardy-Littlewood maximal function is continuous at the point a$\mathbb{R}^n$ if and only if there exists $\tilde{f}$ such that

$ \tilde{f}(x)=f(x) $

for almost every x$\mathbb{R}^n$, and

$ \limsup _{x \rightarrow a}|\tilde{f}(x)| \leqslant M^{\mathrm{c}} f(a) . $

Theorem B   Let ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right) $ and Mcƒ(a) < ∞. The centered Hardy-Littlewood maximal function is continuous at the point a$\mathbb{R}^n$ if and only if there exists $\tilde{f}$ such that

$ \tilde{f}(x)=f(x) $

for almost every x$\mathbb{R}^n$, and

$ \limsup _{x \rightarrow a}|\tilde{f}(x)| \leqslant M^{\mathrm{c}} f(a) . $
1 Main results

Noticing the same properties on norms of Mγ and M, it is natural to consider whether Mγƒ and $M_\gamma^{\mathrm{c}} f$ have similar continuity characterizations to that in Theorem A and B. Finally, we obtain the following results.

Theorem 1.1   Ifƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$ and Mγƒ(x0) < ∞, then Mγƒ is continuous at the point x0 if and only if there exists a function $\tilde{f}$ such that

$ \tilde{f}(x)=f(x) $ (2)

for almost every x$\mathbb{R}^n$, and

$ \limsup _{x \rightarrow x_0}|\tilde{f}(x)| \leqslant M_\gamma f\left(x_0\right) . $ (3)

Theorem 1.1 characterizes an equivalent condition of continuity of Mγƒ. For $M_\gamma^{\mathrm{c}} f$, we have a similar characterization in Theorem 1.2.

Theorem 1.2   If ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$ and $M_\gamma^{\mathrm{c}} f\left(x_0\right)$ < ∞, then $M_\gamma^{\mathrm{c}} f$ is continuous at the point x0 if and only if there exists a function $\tilde{f}$ such that

$ \tilde{f}(x)=f(x), $

for almost every x$\mathbb{R}^n$, and

$ \limsup _{x \rightarrow x_0}|\tilde{f}(x)| \leqslant M_γ^{\mathrm{c}} f\left(x_0\right) . $
2 Some facts and lemmas

In this section, we give two useful lemmas. In Lemma 2.1, we prove the lower semi-continuity of Mγƒ and $M_\gamma^{\mathrm{c}} f$. In Lemma 2.2 and 2.3, we investigate the upper semi-continuity with respect to γ of the value Mγƒ(x0) and $M_\gamma^{\mathrm{c}} f\left(x_0\right)$ for some fixed x0$\mathbb{R}^n$.

It is well-known that and Mcƒ are lower semi-continuous. So are Mγƒ and $M_\gamma^{\mathrm{c}} f$. Now we provide the proof of the lower semi-continuity of Mγƒ and $M_\gamma^{\mathrm{c}} f$.

Lemma 2.1   If ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$, 0 < γ < ∞ and Mγƒ(x0) < ∞, then Mγƒ and $M_\gamma^{\mathrm{c}} f$ are lower semi-continuous at the point x0.

Proof   For any ε>0, according to the definition of Mγƒ, there exists a ball $B_r \ni x_0$ with radius r∈(0, γ) such that

$ \frac{1}{\left|B_r\right|} \int_{B_r}|f(y)| \mathrm{d} y>M_\gamma f\left(x_0\right)-\frac{\varepsilon}{2} . $

Take δ∈(0, γ-r) such that

$ \frac{\left|B_{r+\delta}\right|}{\left|B_r\right|}<\frac{2 M_\gamma f\left(x_0\right)-\varepsilon}{2 M_\gamma f\left(x_0\right)-2 \varepsilon} . $

Let $\hat{x}_0$ denote the center of Br. For any xBδ(x0), we have

$ \left|x-\hat{x}_0\right| \leqslant\left|x-x_0\right|+\left|x_0-\hat{x}_0\right|<r+\delta, $

thus $B_\delta\left(x_0\right) \subset B_{r+\delta}\left(\hat{x}_0\right)$. For any xBδ(x0), we conclude that

$ \begin{aligned} M_\gamma f(x) & \geqslant \frac{1}{\left|B_{r+\delta}\left(\hat{x}_0\right)\right|} \int_{B_{r+\delta\left(\hat{x}_0\right)}}|f(y)| \mathrm{d} y \\ & \geqslant \frac{\left|B_r\right|}{\left|B_{r+\delta}\right|} \cdot \frac{1}{\left|B_r\right|} \int_{B_{r\left(\hat{x}_0\right)}}|f(y)| \mathrm{d} y \\ & >\frac{2 M_\gamma f\left(x_0\right)-2 \varepsilon}{2 M_\gamma f\left(x_0\right)-\varepsilon} \cdot\left(M_\gamma f\left(x_0\right)-\frac{\varepsilon}{2}\right) \\ & =M_\gamma f\left(x_0\right)-\varepsilon . \end{aligned} $ (4)

The calculation (4) implies that Mムis lower semi-continuous at the point x0.

The proof of $M_\gamma^{\mathrm{c}} f$ is quite similar. We just need to require $\hat{x}_0=x_0$ as an extra condition.

In order to prove Theorem 1.1, we also need the following crucial lemma that proves the upper semi-continuity with respect to γ of the value Mγƒ(x0) for fixed x0.

Lemma 2.2   Let ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$. For fixed x0$\mathbb{R}^n$, define

$ \varphi(\gamma)=M_\chi f\left(x_0\right)=\sup _{\substack{B_r \ni x_0 \\ 0<r<\gamma}} \frac{1}{\left|B_r\right|} \int_{B_r}|f(y)| \mathrm{d} y . $ (5)

Then one of the following two cases is true:

(i) φ(γ)=∞ for any γ∈(0, ∞);

(ii) φ is upper semi-continuous on (0, ∞).

Proof    Suppose that there exists γ0>0 such that φ(γ0)=∞. Now we prove that (i) is true.

Notice the observation that γ>γ0 implies

$ \begin{aligned} \varphi(\gamma) & =\sup _{\substack{B_r \ni x_0 \\ 0<r<\gamma}} \frac{1}{\left|B_r\right|} \int_{B_r}|f(y)| \mathrm{d} y \\ & \geqslant \sup _{\substack{B_r \ni x_0 \\ 0<r<\gamma}} \frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y \\ & =\varphi\left(\gamma_0\right) . \end{aligned} $ (6)

It immediately follows from (6) that φ(γ)=∞ for any γ>γ0.

Next we just need to prove that φ(γ)=∞ for any γ∈(0, γ0). By the definition of , we can choose a sequence of balls $\left\{B_k\right\}_{k=1}^{\infty}$ such that each of them contains x0, the radius rk of Bk is less than γ0, and

$ \frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y=\frac{1}{V_n r_{k B_k}^n}|f(y)| \mathrm{d} y \geqslant k, $ (7)

where Vn is the Lebesgue measure of the ndimensional unit ball.

Then we claim that

$ \lim _{k \rightarrow \infty} r_k=0 . $ (8)

Assume that the assertion (8) is true. Then for any γ∈(0, γ0), there exists an integer K>0 such that rk < γ for any k>K. It follows from (7) that

$ \begin{aligned} \varphi(\gamma) & =\sup _{\substack{B_r \ni x_0 \\ 0<r<\gamma}} \frac{1}{\left|B_r\right|} \int_{B_r}|f(y)| \mathrm{d} y \\ & \geqslant \sup _{k>K} \frac{1}{\left|B_r\right|} \int_{B_r}|f(y)| \mathrm{d} y \\ & =\infty . \end{aligned} $

Now let us prove the assertion (8) is true. Recalling that ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$ and noticing that $B_k \subset B_{2 \gamma_0}\left(x_0\right)$

$ \int\limits_{{B_k}} {f(y)|{\rm{d}}y} \int\limits_{{B_{2{\gamma _0}}}\left( {{x_0}} \right)} {|f(y)|{\rm{d}}y: = C < \infty } . $ (9)

It follows from (7) and (9) that

$ r_k \leqslant \frac{C}{V_n k} . $ (10)

The inequality (10) implies the assertion (8).

Next we investigate the other case. If φ(γ) < ∞ for any γ∈(0, ∞), we can prove case (ii) by contradiction. Assume that case (ii) is false. Then there exists a point γ0 such that φ is not upper semi-continuous at the point. Furthermore, there exists α>0 such that

$ \limsup _{\gamma \rightarrow\gamma_0} \varphi(\gamma)>\varphi\left(\gamma_0\right)+\alpha $ (11)

Recalling the observation (6), the statement (11) implies that there exists a sequence $\left\{\boldsymbol{\gamma}_k\right\}_{k=1}^{\infty}$ such that γ0 < γk < 2γ0, monotonically decreasing,

$ \lim _{k \rightarrow \infty} \gamma_k=\gamma_0, $ (12)

and

$ \varphi\left(\gamma_k\right)>\varphi\left(\gamma_0\right)+\frac{\alpha}{2} $

holds for any k≥1. By the definition (5) of φ, there exists a sequence of balls $\left\{B_k\right\}_{k=1}^{\infty}$ such that each of them contains x0,

$ \frac{1}{{\left| {{B_k}} \right|}}\int\limits_{{B_k}} {f(y)|{\rm{d}}y > \varphi \left( {{\gamma _0}} \right) + \frac{\alpha }{4}, } $ (13)

and 0 < rk < γk, where rk denotes the radius of Bk. The statements (12) and (13) imply that γ0 < rk < γk and rkγ0 as k→∞. We denote the center of Bk by $\hat{x}_k$. Let B′k be the ball with radius r′k=2γ0-rk centered at the point

$ \begin{aligned} \hat{x}_k^{\prime} & =\left(1-\frac{r_k^{\prime}}{r_k}\right) x_0+\frac{r_k^{\prime}}{r_k} \hat{x}_k \\ & =2\left(1-\frac{\gamma_0}{r_k}\right) x_0+\left(\frac{2 \gamma_0}{r_k}-1\right) \hat{x}_k \end{aligned} $

Then we can deduce that x0$B_k^{\prime} \subset B_k$, 0 < r′k < γ0 and r′kγ0 as k→∞. By the definition (5) of φ, the statement (13) implies that

$ \begin{aligned} \frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y & >\varphi\left(\gamma_0\right)+\frac{\alpha}{4} \\ & >\frac{1}{\left|B_k^{\prime}\right|} \int_{B_k^{\prime}}|f(y)| \mathrm{d} y+\frac{\alpha}{4} . \end{aligned} $

Then we have that

$ \begin{aligned} \limsup _{k \rightarrow \infty} & \left(\frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y-\right. \\ & \left.\frac{1}{\left|B_k^{\prime}\right|} \int_{B_k^{\prime}}|f(y)| \mathrm{d} y\right) \geqslant \frac{\alpha}{4} . \end{aligned} $ (14)

On the other hand, denote the characteristic function of $A \subset \mathbb{R}^n \text { by } \chi_A$. For any γ0∈(0, ∞), let

$ \begin{aligned} g(x) & =|f(x)| \chi_{B_{4 \gamma_0}\left(x_0\right)}(x) \\ & =\left\{\begin{array}{cc} |f(x)|, & x \in B_{B_{4 \gamma_0}}\left(x_0\right), \\ 0, & x \notin B_{4 \gamma_0}\left(x_0\right) . \end{array}\right. \end{aligned} $

Then we obtain that gL1($\mathbb{R}^n$) is implied just by $f \in L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$. For k≥1, set

$ f_k(x)=|f(x)| \chi_{B_k \backslash B_k^{\prime}}(x), \quad x \in \mathbb{R}^n . $

Due to 0 < r′k < γ0 < rk < 2γ0 and the choice of B′k, we have $B_k^{\prime} \subset B_k \subset B_{4 \gamma_0}\left(x_0\right)$. It follows that $B_k \backslash B^{\prime}{ }_k \subset B_{4 \gamma_0}\left(x_0\right)$, which implies that ƒk is controlled by g. Using the Lebesgue dominated convergence theorem, we obtain that

$ \begin{aligned} & \lim _{k \rightarrow \infty}\left(\frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y-\frac{1}{\left|B_k^{\prime}\right|} \int_{B_k^{\prime}}|f(y)| \mathrm{d} y\right) \\ & \leqslant \lim _{k \rightarrow \infty} \frac{1}{\left|B_k\right|}\left(\int_{B_k}|f(y)| \mathrm{d} y-\int_{B_k^{\prime}}|f(y)| \mathrm{d} y\right) \\ &= \lim _{k \rightarrow \infty} \frac{1}{\left|B_k\right|} \int_{B_k \backslash B_k^{\prime}}|f(y)| \mathrm{d} y \\ &= \lim _{k \rightarrow \infty} \frac{1}{\left|B_k\right|} \int_{\mathbf{R}^n}|f(y)| \chi_{B_k \backslash B_k^{\prime}}(y) \mathrm{d} y \\ &= \lim _{k \rightarrow \infty} \frac{1}{\left|B_k\right|} \int_{\mathbf{R}^n} f_k(y) \mathrm{d} y \\ &= \frac{1}{\left|B_k\right|} \int_{\mathbf{R}^n } \lim _{k\rightarrow \infty} f_k(y) \mathrm{d} y=0, \text { thus } \\ & \limsup _{k \rightarrow \infty}\left(\frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y-\frac{1}{\left|B_k^{\prime}\right|} \int_{B_k^{\prime}}|f(y)| \mathrm{d} y\right) \leqslant 0, \end{aligned} $

which contradicts with (14).

Therefore case (ii) is true.

For the centered version, we have a similar result, and the proof can be repeated almost word for word. We just need an additional condition that the center of the ball is at the point x0 every time we take balls.

Lemma 2.3   Let $f \in L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$. For fixed x0$\mathbb{R}^n$, define

$ \varphi_{\mathrm{c}}(\gamma)=M_{\gamma}^{\mathrm{c}} f\left(x_0\right)=\sup _{0<r<\gamma} \frac{1}{\left|B_r\left(x_0\right)\right|} \int_{B_r\left(x_0\right)}|f(y)| \mathrm{d} y . $

Then one of the following two cases is true:

(i) φc(γ)=∞ for any γ∈(0, ∞);

(ii) φc is upper semi-continuous on (0, ∞).

3 Proof of main theorems and discussions

Now we present the proof of Theorem 1.1.

Proof of Theorem 1.1   First we consider the necessity. Define

$ \tilde{f}(x)= \begin{cases}M_γ f(x), & |f(x)|>M_\gamma f(x), \\ f(x), & |f(x)| \leqslant M_\gamma f(x) .\end{cases} $ (15)

If x is one of the Lebesgue points of ƒ, then we have that

$ |f(x)| \leqslant M_\gamma^{\mathrm{c}} f(x) \leqslant M_\gamma f(x) . $ (16)

Due to ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$, almost every x$\mathbb{R}^n$ is a Lebesgue point of ƒ. It follows that (16) holds almost everywhere. Furthermore, $\tilde{f}$=f holds almost everywhere on $\mathbb{R}^n$.

On the other hand, (15) implies that $|\tilde{f}(x)| \leqslant M_\gamma f(x)$. Therefore the continuity of Mムyields that

$ \limsup _{x \rightarrow x_0}|\tilde{f}(x)| \leqslant \limsup _{x \rightarrow x_0} M_\gamma f(x)=M_\gamma f\left(x_0\right) . $

Next we prove the sufficiency by contradiction. Suppose that Mγƒ is not continuous at the point x0. Then Mγƒ is not upper semi-continuous at x0, since Mγƒ is lower semi-continuous at the point x0. Thus there exists a positive real number α>0 such that

$ \limsup _{x \rightarrow x_0} M_γ f(x)>M_γ f\left(x_0\right)+\alpha . $ (17)

The statement (17) implies that there exists a sequence of balls $\left\{\boldsymbol{B}_k\right\}_{k=1}^{\infty}$ such that xkx0 as k→∞ and

$ M_γ f\left(x_k\right)>M_\gamma f\left(x_0\right)+\alpha $

holds for any k≥1. By the definition (1) of Mγƒ, for any xk, there exists a ball $B_k \ni x_k$ such that the radius rk of Bk is less than γ and

$ \begin{aligned} \frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y & >M_\gamma f\left(x_k\right)-\frac{\alpha}{2} \\ & >M_\gamma f\left(x_0\right)+\frac{\alpha}{2} . \end{aligned} $ (18)

We claim that

$ \lim _{k \rightarrow \infty} r_k=0 . $ (19)

Assume that the assertion (19) is true. It follows from (2) and (18) that

$ \begin{aligned} M_\gamma f\left(x_0\right)+\frac{\alpha}{2} & <\frac{1}{\left|B_k\right|} \int_{B_k}|f(y)| \mathrm{d} y \\ & =\frac{1}{\left|B_k\right|} \int_{B_k}|\tilde{f}(y)| \mathrm{d} y . \end{aligned} $ (20)

For any Bk, there exists xkBk such that

$ \left|\tilde{f}\left(\bar{x}_k\right)\right| \geqslant \frac{1}{\left|B_k\right|} \int_{B_k}|\tilde{f}(y)| \mathrm{d} y>M_\gamma f\left(x_0\right)+\frac{\alpha}{2} . $

We can deduce that xkx0 as k→∞ from (19) since xkx0 as k→∞. It follows that

$ \limsup _{x \rightarrow x_0}|\tilde{f}(x)| \geqslant \limsup _{k \rightarrow \infty}\left|\tilde{f}\left(\bar{x}_k\right)\right| \geqslant M_\gamma f\left(x_0\right)+\frac{\alpha}{2}, $

which contradicts with (3). Therefore Mムis continuous at the point x0.

Now we prove the assertion (19) by contradiction again. Suppose that the assertion (19) is false, then we have

$ 0<c:=\limsup _{k \rightarrow \infty} r_k \leqslant \gamma . $

Take a subsequence of $\left\{r_k\right\}_{k=1}^{\infty}$, still denoted by $\left\{r_k\right\}_{k=1}^{\infty}$ without loss of clarity, such that

$ \lim _{k \rightarrow \infty} r_k=c>0 . $

Lemma 2.2 implies that Mγƒ(x0) is upper semi-continuous with respect to γ. Thus there exists γ′> γ such that

$ M_{\gamma^{\prime}} f\left(x_0\right) \leqslant M_\gamma f\left(x_0\right)+\frac{\alpha}{4} . $ (21)

Take ε=γ′-γ. Then there exists K>0 such that |xk-a| < ε for any k>K. Recalling that 0 < rk < γ, we have

$ \rho_k:=r_k+\left|x_k-x_0\right|<\gamma^{\prime} . $ (22)

Denote the center of Bk by $\hat{x}_k $. Then we have

$ \begin{aligned} \left|\hat{x}_k-x_0\right| & \leqslant\left|\hat{x}_k-x_k\right|+\left|x_k-x_0\right| \\ & <r_k+\left|x_k-x_0\right|=\rho_k \end{aligned} $

which implies that x0Bρk($\hat{x}_k $). For any k>K, it follows from (20), (21), and (22) that

$ \begin{aligned} M_\gamma f\left(x_0\right)+\frac{\alpha}{4} & \geqslant M_\gamma f\left(x_0\right) \\ & \geqslant \frac{1}{\left|B_{\rho_k}\left(\hat{x}_k\right)\right|} \int_{B_{\rho_k}\left(\hat{x}_k\right)}|f(y)| \mathrm{d} y \\ & \geqslant \frac{\left|B_k\right|}{\left|B_{\rho_k}\left(\hat{x}_k\right)\right|} \cdot \frac{1}{\left|B_k\right|} \int_{R_k}|f(y)| \mathrm{d} y \\ & >\frac{r_k^n}{\left(r_k+\left|x_k-x_0\right|\right)^n} \cdot\left(M_\gamma f\left(x_0\right)+\frac{\alpha}{2}\right) . \end{aligned} $

Let k→∞, then we deduce that

$ M_\gamma f\left(x_0\right)+\frac{\alpha}{4} \geqslant M_\gamma f\left(x_0\right)+\frac{\alpha}{2}, $

which is a contradiction. Therefore the assertion (19) is true.

The proof of Theorem 1.2 is quite similar to the proof of Theorem 1.1. In fact, we just need an additional condition that the center of the ball is at the point x0 every time we take balls. So we omit the proof of Theorem 1.2.

The applications and discussions of the classical Hardy-Littlewood maximal functions presented by Wu and Yan[8] make sense for the truncated version as well, and the proof just need to be adjusted as we did in Theorem 1.1. Now we give the results without proof.

Theorem 3.1   Let ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$. If ƒ is continuous at the point x0, then Mγƒ and $M_\gamma^{\mathrm{c}} f$ is continuous at x0.

Theorem 3.2   Let ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$. If ƒ is of locally bounded variation at the point x0, then Mγƒ is continuous at x0.

Remark 3.1   The same as the classical version, Theorem 3.2 is not true on $\mathbb{R}^n$ for n≥2. We can use the function ƒ showed by Wu and Yan[8] to explain it as well. For the convenience of reading, we repeat the definition of ƒ. We just consider the case of n=2.

Define a set $E \subset \mathbb{R}^2$ by

E: ={(x, y)∈ $\mathbb{R}^2$ |x < 1, x+y>0, x-y>0}.

For k=1, 2, …, define Bk, $C_k \subset \mathbb{R}$ by

$ \begin{aligned} & B_k:=\left\{x \in \mathbb{R} \left\lvert\, \frac{1}{2^k} \leqslant x<\frac{1}{2^k}+\frac{1}{2^{k+2}}\right.\right\}, \\ & C_k:=\left\{x \in \mathbb{R} \left\lvert\, -\frac{1}{2^{k+1}} \leqslant <\frac{1}{2^{k+1}}\right.\right\}, \end{aligned} $

and let

Ak=Bk×Ck.

Then $A_k \subset E$ for k≥1.

Furthermore, define ƒ: $\mathbb{R}^2 \rightarrow \mathbb{R}$ by

$ f(x, y):=\sum\limits_{k=1}^{\infty} \chi_{A_k}(x, y), $

where χAk is the characteristic function of Ak. Then ƒ is of locally bounded variation at the coordinate origin (0, 0). Take γ=2, and consider the point sequence $\left\{p_k=\left(\frac{1}{2^k}, 0\right)\right\}_{k=1}^{\infty} $. It is obvious that pk→(0, 0) as k→∞. By the definition (1) of Mγƒ, we have

Mγƒ(pk)=1,

and

$ M_\gamma f(0, 0) \geqslant \frac{1}{2} . $

Therefore, Mムis not continuous at (0, 0).

Remark 3.2   The converse of the Theorem 3.2 is false as well. It is equivalent to that there exists a function ƒ$L_{\mathrm{loc}}^1\left(\mathbb{R}^n\right)$, a real number γ>0 and a point x0$\mathbb{R}$ such that Mγƒ is continuous at x0 but ƒ does not equal to a function g almost everywhere, where g is of locally bounded variation at x0. Here we give a counterexample. Consier the function

$ f(x)= \begin{cases}x \sin \left(\frac{1}{x}\right), & 0<x<1 \\ 0, & x \leqslant 0 \text { or } x \geqslant 1 .\end{cases} $

Notice that ƒ is continuous. By the Theorem 3.1, Mγƒ is continuous at the point 0. On the other hand, the continuity of ƒ implies that if g=ƒ almost everywhere, then g=ƒ everywhere. Thus we just need to show that ƒ is not of locally bounded variation at 0. For x≠0 and x≠1, calculate the derivative of ƒ, and we have

$ f^{\prime}(x)= \begin{cases}\sin \left(\frac{1}{x}\right)-\frac{1}{x} \cos \left(\frac{1}{x}\right), & 0<x<1, \\ 0, & x<0 \text { or } x>1 .\end{cases} $

Then for any δ>0, there exists an integer K≥1 such that

$ \frac{1}{2 K \pi+2 \pi / 3}<\delta . $

Then we have

$ \begin{aligned} & \int_{-\delta}^\delta\left|f^{\prime}(x)\right| \mathrm{d} x \\ \geqslant & \sum\limits_{k=K}^{\infty} \int_{\frac{1}{2 k\mathsf{π}++\mathsf{π}}}^{ \frac{1}{2 k\mathsf{π}+2 \mathsf{π} / 3}}\left|\sin \left(\frac{1}{x}\right)-\frac{1}{x} \cos \left(\frac{1}{x}\right)\right| \mathrm{d} x \\ \geqslant & \sum\limits_{k=K}^{\infty} \int_{\frac{1}{2 k \mathsf{π}+\mathsf{π}}}^{\frac{1}{2 k \mathsf{π}+\frac{2 \mathsf{π}}{3}}} k \mathsf{π} \mathrm{~d} x=\sum\limits_{k=K}^{\infty} \frac{k}{2(2 k+1)(3 k+1)} \\ \geqslant & \sum\limits_{k=K}^{\infty} \frac{1}{12(k+1)}=\infty . \end{aligned} $

Thus ƒ is not of locally bounded variation at 0.

Remark 3.3   Observe that the codomain of Mγƒ is [0, ∞]. We can consider taking ∞ as a valid value when investigating the pointwise continuity of Mγƒ. If we generalize the definition of pointwise continuity in the following way: Mγƒ is continuous at the point x0$\mathbb{R}^n$, if

$ \lim _{x \rightarrow x_0} M_γ f(x)=M_γ f\left(x_0\right) . $

This would not cause confusions. Then Mγƒ is continuous at the point x0 if Mγƒ(x0)=∞, since Mγƒ is lower continuous. Thus in this sense, the condition Mγƒ(x0) < ∞ can be removed from theorems or lemmas.

References
[1]
Hardy G H, Littlewood J E. A maximal theorem with function-theoretic applications[J]. Acta Mathematica, 1930, 54(1): 81-116. DOI:10.1007/BF02547518
[2]
Wiener N. The ergodic theorem[J]. Duke Mathematical Journal, 1939, 5(1): 1-18. DOI:10.1215/s0012-7094-39-00501-6
[3]
Grafakos L. Classical Fourier analysis[M]. Graduate texts in mathematics: vol 249. 3rd ed. New York, NY: Springer New York, 2014.
[4]
Rudin W. Real and complex analysis[M]. 3rd ed. New York, NY: McGraw-Hill, 1987.
[5]
Wei M Q, Nie X D, Wu D, et al. A note on Hardy-Littlewood maximal operators[J]. Journal of Inequalities and Applications, 2016, 2016(1): 21. DOI:10.1186/s13660-016-0963-x
[6]
Kinnunen J. The Hardy-Littlewood maximal function of a Sobolev function[J]. Israel Journal of Mathematics, 1997, 100(1): 117-124. DOI:10.1007/BF02773636
[7]
Aldaz J M, Pérez Lázaro J. Functions of bounded variation, the derivative of the one dimensional maximal function, and applications to inequalities[J]. Transactions of the American Mathematical Society, 2007, 359(5): 2443-2461. DOI:10.1090/s0002-9947-06-04347-9
[8]
Wu D, Yan D Y. Continuity of Hardy-Littlewood maximal function[J]. Acta Mathematicae Applicatae Sinica, English Series, 2020, 36(4): 982-990. DOI:10.1007/s10255-020-0983-z