1 Introduction
As is well known, for the real-valued local integrable functions $f$ on $\mathbb{R}^n, $ the classical Hardy-Littlewood maximal operator $M$ is defined by
$
Mf(x)=\sup\limits_{x\in Q}\frac{1}{|Q|}\displaystyle\int_Q|f(y)|dy,
$ |
where $Q$ is a non-degenerate cube with its sides paralleled to the coordinate axes and $|Q|$ is the Lebesgue measure of $Q.$
Suppose that $ {\it\Phi}: [0, \infty)\rightarrow \mathbb{R} $ is nondecreasing and continuous with $ {\it\Phi}(0)=0$ and $\lim\limits_{t\rightarrow+\infty}{\it\Phi}(t)=\infty.$ Let
$
L^{{\it\Phi}}=\Big\{f:~\displaystyle\int_{R^n}
{\it\Phi}(\varepsilon|f|)dx <\infty~
\hbox{ for some }\varepsilon>0 \Big\}.
$ |
Then $L^{\it{\Phi}}$ is called an Orlicz space. In Orlicz spaces, Kokilashvili and Krbec [1, Theorem 1.2.1] stated that the following statements are equivalent:
(a) there exists a positive constant $C$ such that
$
\begin{equation}
\label{Booka1}
\displaystyle\int_{{\mathbb{R}^n}}{\it\Phi}(Mf)dx\leq
C\displaystyle\int_{{\mathbb{R}^n}}{\it\Phi}(C|f|)dx, ~~\forall
f\in L^1_{\hbox{loc}};
\end{equation}
$ |
(1.1) |
(b) the function ${\it\Phi}^\alpha$ is quasi-convex for some $\alpha\in(0, 1).$
When $ {\it\Phi}(t)=\frac{t^p}{p}, ~~p>1, $ $L^{\it{\Phi}}$ is the space $L^p.$ The above result implies that $M$ is bounded on $L^p$ for $1 <p < \infty$ (see Grafakos [2, Theorem 2.1.6]) and unbounded on $L^1$(see Grafakos [2, Example 2.1.2]). In fact, $M$ maps $L^1$ to $L^{1, \infty}$ (see Grafakos [2, Theorem 2.1.6]).
Let $\mathbb{T}$ be the group of real numbers modulo $2\pi$ and $f(x)$ a real-valued integrable function denoted on $\mathbb{T}$ with period $2\pi.$ The Hardy-Littlewood maximal function $M_{\mathbb{T}}f(x)$ is defined by $M_{\mathbb{T}}f(x)=\sup\limits_{x\in
I}\frac{1}{|I|}\displaystyle\int_I|f(y)|dy, $ where the supremum taken over all open $I\subseteq\mathbb{T}$ with $x\in I.$
Let $a(s)$ and $b(s)$ be two positive nondecreasing continuous functions defined on $[0, \infty)$ satisfying
$
\displaystyle\int_0^1\frac{a(s)}{s}ds=K <+\infty,
$ |
(1.2) |
$
\displaystyle\int_1^\infty\frac{a(s)}{s}ds=+\infty
$ |
(1.3) |
and
$
\begin{equation}
\label{a2}\lim\limits_{s\rightarrow\infty}b(s)=+\infty.
\end{equation}
$ |
(1.4) |
Then, we define
$
\begin{equation}
\label{a3}{\it\Phi}(t)=\displaystyle\int_0^ta(s)ds,
~~~{\it\Psi}(t)=\displaystyle\int_0^tb(s)ds, ~\forall t\geq0.
\end{equation}
$ |
(1.5) |
Under hypotheses (1.3)–(1.5), a series of inequalities involving $M_{\mathbb{T}}f$ and $f$ were obtained. Kita [3, Theorem 2.1] gave necessary and sufficient conditions for $a(\cdot)$ and $b(\cdot)$ in order to guarantee that $M_{\mathbb{T}}f$ is in $L^{\it{\Phi}}$ whenever $f$ is in $L^{\it{\Psi}}.$ Conversely, Kita [4, Theorem 2.1] stated necessary and sufficient conditions for $a(\cdot)$ and $b(\cdot)$ in order to guarantee that a function $f$ is in $L^{\it{\Psi}}$ whenever the function $M_{\mathbb{T}}f$ is in $L^{\it{\Phi}}, $ which is a generalization of [5, Theorem 5.4] to functions in an Orlicz space $L^{\it{\Psi}}$ and also a generalization of the result of [6, Theorem 3.1]. These are related to the class of $L\log L.$ The class was introduced by Zygmund to give a sufficient condition on the local integrability of the Hardy-Littlewood maximal operator. The necessity of this condition was observed by Stein [7].
Under the additional hypothesis $(1.2), $ Kita [8, Theorem 2.3, Theorem 2.5] extended the above results [3, Theorem 2.1] and [4, Theorem 2.1] to the cases on $\mathbb{R}^n.$ Moreover, Kita [9, 10] obtained the related weighted theory on $\mathbb{R}^n.$ Specially, Kita [10] also dealt with the iterated maximal operator.
Comparing with the results without weights on $\mathbb{R}^n$ and $\mathbb{T}$, Mei and Liu [11, 12] considered the martingale cases. In our paper, we prove weighted inequalities for maximal operators in Orlicz martingale classes. Let us introduce some preliminaries.
Let $(\Omega, \mathcal {F}, \mu)$ be a complete probability space, and $(\mathcal {F}_n)_{n\geq0}$ be an increasing sequence of sub-$\mathcal {F}$-fields of $\mathcal{F}$ with $\mathcal{F}=\bigvee\mathcal{F}_n.$ If $f=(f_n)_{n\geq0}$ is a martingale adapted to $(\mathcal {F}_n)_{n\geq0}, $ the maximal function (operator) $Mf$ for martingale $f=(f_n)$ is defined by $Mf=\sup\limits_{n\geq 0}|f_n|.$ In this paper, the limit of $f_n$ is also denoted by $f.$ Specifically, $\big({\Omega}, (\mathcal
{F}_n)_{n\geq0}, \mathcal{F}, \nu\big)$ is said to satisfy the $R\hbox{-condition}$ (or simply $\nu\in R$), if for all $n\geq1, ~F_n\in\mathcal {F}_n, $ there exists a $G_n\in \mathcal
{F}_{n-1}$ such that $F_n\subset G_n$ and $\nu(G_n)\leq
d\cdot\nu(F_n).$ In our paper, we assume that $\mathcal
{F}_{0}=\{\emptyset, \Omega\}.$
A weight $\omega$ is a positive random variable with $E \omega <\infty.$ For convenience, we assume $E \omega=1$ in this paper. In addition, we say $\omega\in A_1$ (or $B_1$), if $\omega^{-1}\in
L^{\infty}(d\mu)$ and there exists a constant $C\geq1$ such that $\omega_n\leq C \omega$ (or $C \omega_n\geq\omega$). Moreover, for the weights as above, a function $f\in L^1(\omega d\mu)$ implies $f\in L^1(d\mu).$
Let $a(s), $ $b(s), $ $\it{\Phi(t)}$ and $\it{\Psi(t)}$ satisfy (1.2)-(1.5). It is clear that $\it{\Phi}$ and $\it{\Psi}$ are convex. Recall that $L^{\it{\Phi}}(d\mu)\cup
L^{\it{\Psi}}(d\mu)\subset L^1(d\mu)$ when $\mu({\Omega}) <\infty.$
Then we have the following Theorems 1.1 and 1.2, which are martingale versions of [8, Theorem 2.3]. Kita [8, Theorem 2.3] gave an assumption that the function $\frac{a(s)}{s}$ is bounded in a neighborhood of zero. However, we do not need the assumption in our Theorems 1.1 and 1.2.
Theorem 1.1 Let $\omega\in A_1.$ Suppose that there exists a positive constant $C_1$ such that
$
\begin{equation}
\label{2b1} \displaystyle\int_1^s\frac{a(t)}{t}dt\leq
C_1b(C_1s), ~s\geq1.
\end{equation}
$ |
(1.6) |
Then there exist positive constants $C_2$ and $C_3$ such that
$
\begin{equation}
\label{2bb1}\displaystyle\int_{{\Omega}}{\it\Phi}(Mf)\omega
d\mu\leq
C_2\|f\|_{L^1(\omega d\mu)}+C_2\displaystyle\int_{{\Omega}}{\it\Psi}(C_3|f|)\omega
d\mu,
\end{equation}
$ |
(1.7) |
where $C_2$ and $C_3$ are constants independent of $f=(f_n).$
Theorem 1.2 Let $(\Omega, \mathcal
{F}, \omega)$ be non-atomic and $\omega\in R\cap B_1.$ Suppose that $f\in L^{\it{\Psi}}(\omega d\mu)$ implies $Mf\in
L^{{\it\Phi}}(\omega d\mu).$ Then $(1.6)$ holds.
Following from Theorems 1.1 and 1.2, we easily have Corollaries 1.3 and 1.4, respectively.
Corollary 1.3 Let $\omega\in A_1.$ Suppose that $(1.6)$ is valid, then $Mf\in
L^{\it{\Phi}}(\omega d\mu)$ for all $f\in L^{\it{\Psi}}(\omega
d\mu).$
Corollary 1.4 Let $(\Omega, \mathcal
{F}, \omega)$ be non-atomic and $\omega\in R\cap B_1.$ Suppose that $(1.7)$ is valid, then we have $(1.6).$
We consider the converse inequality of maximal functions. Theorems 1.5 and 1.6 are martingale versions of [8, Theorem 2.5] and involve weights.
Theorem 1.5 Let $\omega\in B_1\cap R.$ Suppose that there exist positive constants $C_6$ and $s_0>1$ such that
$
\begin{equation}
\label{add2} \displaystyle\int_1^s\frac{a(t)}{t}dt\geq
C_6b(C_6s), ~s\geq s_0.
\end{equation}
$ |
(1.8) |
Then there exist positive constants $C_7$ and $C_8$ such that
$
\begin{equation}
\label{add3}\displaystyle\int_{{\Omega}}{\it\Psi}(C_7|f|)\omega
d\mu\leq C_8\|f\|_{L^1(\omega d\mu)}+C_8
\displaystyle\int_{{\Omega}}{\it\Phi}(M(|f|))\omega d\mu,
\end{equation}
$ |
(1.9) |
where $C_7$ and $C_8$ are constants independent of $f=(f_n)$ and $\|f\|_{L^1(\omega d\mu)}\leq1.$
Theorem 1.6 Let $(\Omega, \mathcal
{F}, \omega)$ be non-atomic and $\omega\in A_1.$ Suppose that $(1.9)$ is valid. Then $(1.8)$ holds.
We easily obtain Corollary 1.7 by Theorems 1.5. In addition, following the proof of Theorem 1.6, we have Corollary 1.8.
Corollary 1.7 Let $\omega\in B_1\cap R.$ Suppose that $(1.8)$ is valid. Then $f\in
L^{\it{\Psi}}(\omega d\mu)$ for all $f$ satisfying $Mf\in
L^{\it{\Phi}}(\omega d\mu).$
Corollary 1.8 Let $(\Omega, \mathcal
{F}, \omega)$ be non-atomic and $\omega\in A_1.$ Suppose that $Mf\in L^{\it{\Phi}}(\omega d\mu)$ implies $f\in
L^{\it{\Psi}}(\omega d\mu).$ Then $(1.8)$ holds.
2 Proofs of Theorems
First, we give some lemmas which will be used in our proofs several times.
Lemma 2.1 Suppose that $\nu$ is an finite measure on $({\Omega}, \mathcal {F}), $ then for each function $f\in
L^1(d\nu), $ we have
$
\begin{equation}
\label{3a1}\displaystyle\int_{\{|f|>t\}}|f|d\nu=t\cdot\nu\{|f|>t\}+\displaystyle\int^\infty_t\nu\{|f|>s\}ds, ~t>0.
\end{equation}
$ |
(2.1) |
Proof The proof follows in the same way as the proof of [9, Lemma 3.1].
Lemma 2.2 Let $\omega\in A_1$. Then there exists a positive constant $C_0$ such that
$
\begin{equation}\omega\{Mf>\lambda\}
\leq\frac{C_0}{\lambda}\displaystyle\int_{\{|f|>\frac{\lambda}{2}\}}|f|\omega
d\mu
\end{equation}
$ |
(2.2) |
for all $\lambda>0$ and $f\in L^1(\omega d\mu).$
Proof By the assumption that $\omega\in A_1, $ there exists a positive constant $C$ such that
$
\omega\{Mg>\lambda\}
\leq\frac{C}{\lambda}\|g\|_{L^1(\omega)}, ~~g\in L^1(\omega)
$ |
in view of [13, Theorem 6.6.2]. For $\lambda>0, $ we set $f_\lambda=f\cdot\chi_{\{|f|\leq\frac{\lambda}{2}\}}$ and $f^\lambda=f-f_\lambda.$ Then $f^\lambda\in L^1(\omega)$ and $
Mf_\lambda=\sup\limits_{n\geq0}|E_n(f_\lambda)|
\leq\sup\limits_{n\geq0}E_n(|f_\lambda|) \leq\frac{\lambda}{2}.$ It follows that
$
\omega\{Mf>\lambda\}
\leq\omega\{Mf_\lambda>\frac{\lambda}{2}\}+
\omega\{Mf^\lambda>\frac{\lambda}{2}\}
\leq\frac{C_0}{\lambda}\displaystyle\int_{\{|f|>\frac{\lambda}{2}\}}|f|\omega
d\mu,
$ |
where $C_0=2C.$
Lemma 2.3 Let $\omega\in A_1$. Then there exists a positive constant $C$ such that
$
\begin{equation}\label{add3d1}\omega\{Mf>t\}
\leq\frac{C}{t}\displaystyle\int_{\frac{t}{2}}^\infty\omega\{|f|>s\}
ds
\end{equation}
$ |
(2.3) |
for all $t>0$ and $f\in L^{\it{\Psi}}(\omega).$
Proof Fix $f\in L^{\it{\Psi}}(\omega d\mu).$ For $t>0, $ set $ f_t=(|f|\wedge\frac{t}{2})\cdot\mathop{\hbox{sign}}f$ and $f^t=f-f_t.$ Trivially, we have $f^t\in L^1(\omega).$ By Lemma $2.2, $ we also have
$
\begin{equation}\label{3e1}\omega\{Mf^t>\lambda\}
\leq\frac{C_0}{\lambda}\displaystyle\int_{\{|f^t|>\frac{\lambda}{2}\}}|f^t|\omega
d\mu.
\end{equation}
$ |
(2.4) |
Let $\lambda=\frac{t}{2}$ in $(2.4).$ It follows that
$
\begin{equation}\label{3f1}\omega\{Mf^t>\frac{t}{2}\}
\leq\frac{2C_0}{t}\displaystyle\int_{{\Omega}}|f^t|\omega d\mu.
\end{equation}
$ |
(2.5) |
Hence,
$
\omega\{Mf>t\}
\leq \omega\{Mf^t>\frac{t}{2}\}+\omega\{Mf_t>\frac{t}{2}\}
=\omega\{Mf^t>\frac{t}{2}\}\\
\leq \frac{2C_0}{t}\displaystyle\int_{{\Omega}}|f^t|\omega d\mu
=\frac{2C_0}{t}\displaystyle\int_{\{|f|>\frac{t}{2}\}}
|f-\frac{t}{2}\cdot\mathop{\hbox{sign}}f|\omega d\mu\\
= \frac{2C_0}{t}\displaystyle\int_{\{|f|>\frac{t}{2}\}}
\Big||f|-\frac{t}{2}\Big|\omega d\mu
=\frac{2C_0}{t}\displaystyle\int_{\{|f|>\frac{t}{2}\}}
(|f|-\frac{t}{2})\omega d\mu.
$ |
Combining with (2.1), we have (2.3) is valid with $C=2C_0.$
Proof of Theorem 1.1 For $C_3>0$ (which will be determined later), fix $f\in L^{\it{\Psi}}(\omega
d\mu)$ such that $\displaystyle\int_{{\Omega}}{\it\Psi}(C_3|f|)\omega d\mu <\infty.$ With the constant $C$ in Lemma $2.3, $ it follows from (1.2), (1.3) and (1.6) that
$
\displaystyle\int_{{\Omega}}\it{\Phi}(Mf)\omega d\mu
= \displaystyle\int_0^\infty\omega\{Mf>t\}a(t)dt\\
\leq C\displaystyle\int_0^\infty\frac{a(t)}{t}
\big(\displaystyle\int_{\frac{t}{2}}^\infty\omega\{|f|>s\} ds\big) dt\\
= C\displaystyle\int_0^\infty\omega\{|f|>s\}
\big(\displaystyle\int_0^{2s}\frac{a(t)}{t}dt\big)ds\\
= C\displaystyle\int_0^\infty\omega\{|f|>s\}
\big(\displaystyle\int_0^{1}\frac{a(t)}{t}dt+\displaystyle\int_1^{2s}\frac{a(t)}{t}dt\big)ds\\
\leq KC\|f\|_{L^1(\omega d\mu)}+CC_1\displaystyle\int_0^\infty\omega\{|f|>s\}
b(2C_1s)ds\\
= KC\|f\|_{L^1(\omega d\mu)}+
\frac{C}{2}\displaystyle\int_{{\Omega}}{\it\Psi}(2C_1|f|)\omega
d\mu.
$ |
Let $C_2=(KC)\vee\frac{C}{2}$ and $C_3=2C_1.$ The we have (1.7).
In order to give the proof of Theorem $1.2, $ we need the following lemmas.
Lemma 2.4 Let $\hat{E}_n(\cdot)$ be the conditional expectation with respect to $(\Omega, \mathcal
({\mathcal {F}}_n)_{n\geq0}, $ $\mathcal {F}, \omega d\mu)$ and $\hat{M}(\cdot)=\sup\limits_{n\geq0}\hat{E}_n(\cdot).$ Suppose $\omega\in R.$ Then
$
\begin{equation}\label{3j1}\displaystyle\int_{\{|f|>\lambda\}}|f|\omega d\mu\leq
d\lambda\cdot\omega\{\hat{M}(|f|)>\lambda\}, ~f\in L^1(\omega
d\mu), ~ \lambda\geq\|f\|_{L^1(\omega d\mu)}.
\end{equation}
$ |
(2.6) |
Proof For $f\in L^1(\omega d\mu), $ the sequence $\big(\hat{E}_n(|f|)\big)_{n\geq0}$ is a uniformly integral martingale with respect to $(\Omega, \mathcal ({\mathcal
{F}}_n)_{n\geq0}, \mathcal {F}, \omega d\mu)$. In view of [13, Theorem 1.3.2.9], we have $\lim\limits_{n\rightarrow\infty}\hat{E}_n(|f|)=|f|~~{\rm a.e.}.$ Then it follows that $|f|\leq \hat{M}(|f|).$
Because of $\omega\in R, $ we have
$
\begin{equation*}\displaystyle\int_{\{\hat{M}f
>\lambda\}}|f|\omega d\mu\leq
d\lambda\cdot\omega\{\hat{M}f>\lambda\}, ~f\in L^1(\omega d\mu), ~
\lambda\geq\|f\|_{L^1(\omega d\mu)},
\end{equation*}
$ |
where we use [13, Theorem 7.1.2]. Combining this with $|f|\leq \hat{M}(|f|), $ we have
$
\begin{equation*}\displaystyle\int_{\{|f|>\lambda\}}|f|\omega d\mu
\leq\displaystyle\int_{\{\hat{M}f
>\lambda\}}|f|\omega d\mu\leq
d\lambda\cdot\omega\{\hat{M}(|f|)>\lambda\}, ~f\in L^1(\omega
d\mu), ~ \lambda\geq\|f\|_{L^1(\omega d\mu)}.
\end{equation*}
$ |
Lemma 2.5 Let $\omega$ be a weight. Then
$
\hat{E}(f\cdot\omega^{-1}|\mathcal{F}_n)\cdot
E(\omega|\mathcal{F}_n)=E(f|\mathcal{F}_n), ~f\in L^1(d\mu).
$ |
Proof For all $ A\in \mathcal{F}_n$, we have
$
\displaystyle\int_Afd\mu
= \displaystyle\int_Af\cdot\omega^{-1}\omega
d\mu
=\displaystyle\int_A\hat{E}(f\cdot\omega^{-1}|
\mathcal{F}_n)\cdot\omega d\mu\\
= \displaystyle\int_A\hat{E}(f\cdot\omega^{-1}|
\mathcal{F}_n)\cdot{E}(\omega|
\mathcal{F}_n)d\mu.
$ |
It follows that $\hat{E}(f\cdot\omega^{-1}|\mathcal{F}_n)\cdot
E(\omega|\mathcal{F}_n)=E(f|\mathcal{F}_n)$.
Reforming Lemma $2.5, $ we have
$
\begin{equation}\label{lemma1}\hat{E}(g|\mathcal{F}_n)\cdot
E(\omega|\mathcal{F}_n)=E(g\omega|\mathcal{F}_n), ~g\in L^1(\omega
d\mu).
\end{equation}
$ |
(2.7) |
Lemma 2.6 Suppose $\omega \in B_1\cap R.$ Then there exist constants $C_4$ and $C_5$ such that
$
\begin{equation}\label{3k1}\displaystyle\int_{\{|f|>C_4\lambda\}}|f|\omega d\mu\leq
C_5\lambda\cdot\omega\{M(|f|)>\lambda\}, ~~f\in L^1(\omega
d\mu), ~\lambda\geq\frac{1}{C}\|f\|_{L^1(\omega d\mu)},
\end{equation}
$ |
(2.8) |
where the constant $C$ is the one in the definition of $B_1.$
Proof Fix $f\in L^1(\omega d\mu)$ and $\lambda\geq
\frac{1}{C}\|f\|_{L^1(\omega d\mu)}.$ Since $\omega\in B_1, $ there exists a constant $C\geq1$ such that $C\cdot\omega_n\geq\omega.$ Combining this with $(2.7), $ we have
$
\hat{E}_n(|f|)=\omega_n^{-1}{E}_n(|f|\omega)={E}_n(|f|\omega_n^{-1}\omega)
\leq{E}_n(|f|\omega_n^{-1}\omega)\leq C{E}_n(|f|).
$ |
Thus
$
\begin{equation}\label{add4}\hat{M}(|f|)\leq CM(|f|).\end{equation}
$ |
(2.9) |
Since $C\lambda\geq\|f\|_{L^1(\omega d\mu)}, $ we have
$
\begin{equation}\label{3k2}\displaystyle\int_{\{|f|>C\lambda\}}|f|\omega
d\mu\leq C d\cdot\lambda\cdot\omega\{CM(|f|)>C\lambda\}
\end{equation}
$ |
(2.10) |
in view of Lemma $2.4.$ Thus, $(2.8)$ is valid with $C_4=C$ and $C_5=Cd.$
Proof of Theorem 1.2 Assume for contradiction that $(1.6)$ does not hold. Then there exists a sequence of positive numbers $s_k\geq1$ such that
$
\begin{equation}\label{3l2}\displaystyle\int_1^{s_k}\frac{a(t)}{t}dt>
2^kb(2^ks_k), ~k\geq1.
\end{equation}
$ |
(2.11) |
Set $\alpha_k=\frac{1}{\alpha2^k{\it\Psi}(2^ks_k)}.$ It is clear that we can choose $\alpha$ big enough to make $\sum\limits_{k=1}\limits^{\infty}\alpha_k\leq1$ and $\frac{1}{\alpha}\leq{\it\Psi}(1).$ By the assumption that $(\Omega, \mathcal {F}, \omega)$ is non-atomic, we have a family of measurable sets $Q_k\in\mathcal {F}$ such that
$
\begin{equation}\label{3m1}
\omega(Q_k)=\alpha_k~\hbox{and}~
Q_{k_1}\cap Q_{k_2}=\emptyset, ~k_1\neq k_2.
\end{equation}
$ |
(2.12) |
Set $f=\sum\limits_{k=1}\limits^{\infty}2^ks_k\cdot\chi_{Q_k}, $ then
$
\displaystyle\int_{{\Omega}}{\it\Psi}(|f|)\omega d\mu
= \sum\limits_{k=1}\limits^{\infty}
\displaystyle\int_{Q_k}{\it\Psi}(|f|)\omega d\mu\\
= \sum\limits_{k=1}\limits^{\infty}
{\it\Psi}(2^ks_k)\omega(Q_k)\\
= \sum\limits_{k=1}\limits^{\infty}
{\it\Psi}(2^ks_k)\frac{1}{\alpha2^k{\it\Psi}(2^ks_k)}\\
= \frac{1}{\alpha}\leq{\it\Psi}(1).
$ |
Thus $f\in L^{\it{\Psi}}(\omega d\mu).$ By Jensen's inequality, we have $f\in L^1(\omega d\mu) \hbox{~and~}
\|f\|_{L^1(\omega d\mu)}\leq1.$
On the other side, we claim that the function $Mf$ is not in $L^{\it{\Phi}}(\omega d\mu).$ Then we get a contradiction. To show the claim in the following way: fix a constant $\varepsilon\in(0, 1]$ for $\varepsilon f, $ with $C, ~C_4 \hbox{ and
}C_5$ in Lemma 2.6, we have
$
\begin{equation}\label{3n1}\frac{\varepsilon}{\lambda}
\displaystyle\int_{\{\varepsilon f>C_4\lambda\}}f\omega d\mu\leq
C_5\cdot\omega\{\varepsilon
Mf>\lambda\}, ~\lambda\geq\frac{\varepsilon}{C}\|f\|_{L^1(\omega
d\mu)}.
\end{equation}
$ |
(2.13) |
Note that $1\geq\frac{\varepsilon}{C}\|f\|_{L^1(\omega d\mu)}, $ we have
$
\displaystyle\int_{\Omega}\it{\Phi}(\varepsilon\cdot
Mf)\omega d\mu
= \displaystyle\int_0^\infty\omega\{\varepsilon\cdot Mf>\lambda\}
a(\lambda)d\lambda\\
\geq \displaystyle\int_1^\infty\omega\{\varepsilon\cdot Mf>\lambda\}
a(\lambda)d\lambda \\
\geq \displaystyle\int_1^\infty\Big(\frac{\varepsilon}{C_5\lambda}
\displaystyle\int_{\{\varepsilon\cdot |f|>C_4\lambda\}}|f|\omega d\mu
\Big)a(\lambda)d\lambda\\
= \frac{\varepsilon}{C_5}\displaystyle\int_\Omega|f|
\Big(\displaystyle\int_1^{\frac{\varepsilon\cdot
|f|}{C_4}}\frac{a(\lambda)}{\lambda}d\lambda\Big)\omega
d\mu\\
\geq \frac{\varepsilon}{C_5}\sum\limits_{k=1}\limits^{\infty}\displaystyle\int_{Q_k}2^ks_k
\Big(\displaystyle\int_1^{\frac{\varepsilon\cdot
2^ks_k}{C_4}}
\frac{a(\lambda)}{\lambda}d\lambda\Big)\omega d\mu
$ |
For the above $\varepsilon, $ we can choose $k(\varepsilon)$ such that $\frac{\varepsilon\cdot 2^k}{C_4}>1, ~k\geq
k(\varepsilon).$ Put
$
L=\frac{\varepsilon}{C_5}\sum\limits_{k=1}\limits^{k(\varepsilon)-1}\displaystyle\int_{Q_k}2^ks_k
\Big(\displaystyle\int_1^{\frac{\varepsilon\cdot
2^ks_k}{C_4}}
\frac{a(\lambda)}{\lambda}d\lambda\Big)\omega d\mu
,
$ |
then $L$ is a real number. Therefore, it follows from $(2.11)\hbox{ and }(2.12)$ that
$
\displaystyle\int_{\Omega}{\it\Phi}(\varepsilon\cdot
Mf)\omega d\mu
\geq L+\frac{\varepsilon}{C_5}\sum\limits_{k=k(\varepsilon)}\limits^{\infty}
\displaystyle\int_{Q_k}2^ks_k
\Big(\displaystyle\int_1^{\frac{\varepsilon\cdot
2^ks_k}{C_4}}\frac{a(\lambda)}{\lambda}d\lambda\Big)\omega
d\mu\\
\geq L+\frac{\varepsilon}{C_5}\sum\limits_{k=k(\varepsilon)}\limits^{\infty}
\displaystyle\int_{Q_k}2^ks_k
\Big(\displaystyle\int_1^{s_k}\frac{a(\lambda)}{\lambda}d\lambda\Big)\omega
d\mu
\geq L+\frac{\varepsilon}{C_5}\sum\limits_{k=k(\varepsilon)}\limits^{\infty}
\displaystyle\int_{Q_k}2^ks_k
\big(2^kb(2^ks_k)\big)\omega
d\mu\\
= L+\frac{\varepsilon}{C_5}\sum\limits_{k=k(\varepsilon)}
\limits^{\infty}2^k2^ks_kb(2^ks_k)\omega(Q_k)
=L+\frac{\varepsilon}{C_5}\sum\limits_{k=k(\varepsilon)}
\limits^{\infty}2^k
2^ks_kb(2^ks_k)\frac{1}{\alpha2^k{\it\Psi}(2^ks_k)}\\
\geq L+\frac{\varepsilon}{C_5}\sum\limits_{k=k(\varepsilon)}
\limits^{\infty}2^k2^ks_k
b(2^ks_k)\frac{1}{\alpha2^k(2^ks_k)b(2^ks_k)}
=L+\frac{\varepsilon}{C_5}\sum\limits_{k=k(\varepsilon)}
\limits^{\infty}\frac{1}{\alpha}=\infty.
$ |
The claim is proved and the proof is finished.
Proof of Theorem 1.5 For $C_7=\frac{C_6}{C}, $ fix $f\in
L^{\it{\Phi}}(\omega
d\mu)$ such that $\displaystyle\int_{{\Omega}}{\it\Phi}(M(|f|))\omega
d\mu <\infty$ and $\|f\|_{L^1(\omega
d\mu)}\leq1, $ where the constant $C$ is the one in the definition of $B_1.$ Then
$
\displaystyle\int_{{\Omega}}{\it\Psi}(\frac{C_6}{C}|f|)\omega
d\mu
= \frac{C_6}{C}\displaystyle\int_0^\infty\omega\{|f|>t\}b(\frac{C_6}{C}t)dt\\
= C_6\displaystyle\int_0^\infty\omega\{|\frac{f}{C}|>t\}b(C_6t)dt\\
= C_6\displaystyle\int_0^{s_0}\omega\{|\frac{f}{C}|>t\}b(C_6t)dt
+C_6\displaystyle\int_{s_0}^\infty\omega\{|\frac{f}{C}|>t\}b(C_6t)dt.
$ |
Denote
$
I_0=C_6\displaystyle\int_0^{s_0}\omega\{|\frac{f}{C}|>t\}b(C_6t)dt
~\hbox{and}~I_1=C_6\displaystyle\int_{s_0}^\infty\omega\{\frac{f}{C}|>t\}b(C_6t)dt.
$ |
Then, for $I_0, $ we have
$
I_0 \leq
C_6b(C_6s_0)\displaystyle\int_0^{s_0}\omega\{|\frac{f}{C}|>t\}dt
\leq C_6b(C_6s_0)\displaystyle\int_0^\infty\omega\{|\frac{f}{C}|>t\}dt\\
= \frac{1}{C}\cdot C_6b(C_6s_0)\|f\|_{L^1(\omega d\mu)}.
$ |
It follows from $(1.8)$ that
$
I_1 \leq \displaystyle\int_{s_0}^\infty\omega\{|\frac{f}{C}|>t\}
\Big(\displaystyle\int_1^t\frac{a(s)}{s} ds\Big) dt
\leq\displaystyle\int_1^\infty\omega\{|\frac{f}{C}|>t\}\Big(\displaystyle\int_1^t\frac{a(s)}{s} ds\Big) dt \\
= \displaystyle\int_1^\infty\frac{a(s)}{s}\Big(\displaystyle\int_s^\infty\omega\{|\frac{f}{C}|>t\}dt\Big)ds.
$ |
It follows from (2.1), (2.6) and (2.9) that
$
I_1 \leq \displaystyle\int_1^\infty\frac{a(s)}{s}
\Big(\displaystyle\int_{\{|\frac{f}{C}|>s\}}|\frac{f}{C}|\omega d\mu\Big)ds
\leq\displaystyle\int_1^\infty\frac{a(s)}{s}
\Big(d\cdot s\cdot\omega\{\hat{M}(|\frac{f}{C}|)>s\}\Big)ds\\
= d\cdot\displaystyle\int_1^\infty a(s)
\cdot\omega\{\hat{M}(|\frac{f}{C}|)>s\}ds
\leq d\cdot\displaystyle\int_1^\infty a(s)
\cdot\omega\{C\cdot M(\frac{|f|}{C})>s\}ds\\
\leq d\cdot\displaystyle\int_\Omega{\it\Phi}(M(|f|)) \omega d\mu,
$ |
where we have used $\|\frac{f}{C}\|_{L^1(\omega
d\mu)}\leq\|f\|_{L^1(\omega d\mu)}\leq1.$ Therefore,
$(1.9)$ is valid with $C_7=\frac{C_6}{C}$ and $C_8=\frac{C_6b(C_6s_0)}{C}\vee d.$
Proof of Theorem 1.6 We proceed by contradiction and assume that $(1.8)$ dose not hold. Then we get a sequence $s_k>1$ satisfying
$
\frac{4}{3} <s_1 <s_2 <\cdot\cdot\cdot <s_k <s_{k+1} <\cdot\cdot\cdot
~\hbox{and}~\lim\limits_{k\rightarrow\infty}s_k=\infty;
$ |
(2.14) |
$
\displaystyle\int_1^{s_k}\frac{a(t)}{t}dt <
\frac{1}{2^k}b(\frac{1}{2^k}s_k), ~k\geq1;
$ |
(2.15) |
$
s_{k+1}>4s_k, ~k\geq1;
$ |
(2.16) |
$
b(\frac{s_k}{2^k})\geq k2^k, ~k\geq1.
$ |
(2.17) |
Let $\alpha_k=\frac{1}{\frac{s_k}{2^k}b(\frac{s_k}{2^k})}, $ then
$
\sum\limits_{k=1}\limits^{\infty}\alpha_k
=\sum\limits_{k=1}\limits^{\infty}
\frac{1}{\frac{s_k}{2^k}b(\frac{s_k}{2^k})}
\leq\sum\limits_{k=1}\limits^{\infty}
\frac{1}{\frac{s_k}{2^k}k2^k}
=\sum\limits_{k=1}\limits^{\infty}
\frac{1}{ks_k}\leq\sum\limits_{k=1}\limits^{\infty}
\frac{1}{s_k}
\leq \frac{1}{s_1}\sum\limits_{k=1}\limits^{\infty}
\frac{1}{4^{k-1}}\leq1.
$ |
By the assumption that $(\Omega, \mathcal {F}, \omega)$ is non-atomic, we have a family of measurable sets $Q_k\in\mathcal
{F}$ such that
$
\begin{equation}\label{extra7}
\omega(Q_k)=\alpha_k~\hbox{and}~
Q_{k_1}\cap Q_{k_2}=\emptyset, ~k_1\neq k_2.
\end{equation}
$ |
(2.18) |
Set $f=\sum\limits_{k=1}\limits^{\infty}\frac{ks_k}{2^k}\cdot\chi_{Q_k}, $ then
$
\displaystyle\int_{{\Omega}}|f|\omega d\mu
=\sum\limits_{k=1}\limits^{\infty}
\displaystyle\int_{Q_k}|f|\omega d\mu
=\sum\limits_{k=1}\limits^{\infty}
\frac{ks_k}{2^k}\omega(Q_k)
=\sum\limits_{k=1}\limits^{\infty}
\frac{ks_k}{2^k}\frac{1}{\frac{s_k}{2^k}b(\frac{s_k}{2^k})}
\leq\sum\limits_{k=1}\limits^{\infty}\frac{1}{2^k}=1.
$ |
Thus $f\in L^1(\omega d\mu)$ and $\|f\|_{L^1(\omega
d\mu)}\leq1.$ Moreover, we also have $\displaystyle\int_{\Omega}\it{\Phi}(Mf)\omega d\mu <\infty.$ In fact, for the sequence $s_k, $ we have
$
\frac{(k+1)s_{k+1}}{2^{k+1}}
\geq\frac{(k+1)4s_k}{2^{k+1}}
>\frac{4ks_k}{2^{k+1}}\geq2\frac{ks_k}{2^k}
>\frac{ks_k}{2^k}, ~k\geq1.
$ |
It is clear that $\frac{s_{k+1}}{2^{k+1}}>\frac{s_k}{2^k}, $ thus
$
\omega(Q_{k+1})
=\frac{1}{\frac{s_{k+1}}{2^{k+1}}b(\frac{s_{k+1}}{2^{k+1}})}
\leq\frac{1}{\frac{4s_k}{2^{k+1}}b(\frac{s_k}{2^k})}
\leq\frac{1}{\frac{2s_k}{2^k}b(\frac{s_k}{2^k})}=\frac{1}{2}\omega(Q_k), ~k\geq1.
$ |
Let $\beta_k=\frac{ks_k}{2^k}, ~k\geq1$ and $\beta_0=0.$ Obviously,
$\beta_k\uparrow\infty.$ For $k\geq1, $ when $s\in[\beta_{k-1}, \beta_k), $ we have
$
\omega\{|f|>s\}=\sum\limits_{i=0}\limits^{\infty}\omega(Q_{k+i})
\leq\omega(Q_k)\sum\limits_{i=0}
\limits^{\infty}\frac{1}{2^i}=2\omega(Q_k)
$ |
and
$
\displaystyle\int_1^{2s}\frac{a(t)}{t}dt\leq0\leq\frac{1}{2^k}b(\frac{s_k}{2^k}), ~\hbox{if
}2s\leq1;\\ \displaystyle\int_1^{2s}\frac{a(t)}{t}dt
\leq\displaystyle\int_1^{2\beta_k}\frac{a(t)}{t}dt
\leq\displaystyle\int_1^{s_k}\frac{a(t)}{t}dt
<\frac{1}{2^k}b(\frac{s_k}{2^k}), ~\hbox{if
}2s>1.
$ |
Combining (2.3), (1.2) and (1.3), we have
$
\displaystyle\int_{\Omega}\it{\Phi}(Mf)\omega d\mu
=\displaystyle\int_0^\infty\omega\{Mf>t\}
a(t)dt\\
\leq \displaystyle\int_0^\infty\Big(\frac{C}{t}
\displaystyle\int_{\frac{t}{2}}^\infty \omega\{|f|>s\}
ds \Big)a(t)dt \\
= C\displaystyle\int_0^\infty \omega\{|f|>s\}\Big(
\displaystyle\int_{0}^{2s}\frac{a(t)}{t}
dt \Big) ds\\
= C\displaystyle\int_0^\infty \omega\{|f|>s\}\Big(
\displaystyle\int_{0}^{1}\frac{a(t)}{t}
dt +\displaystyle\int_{1}^{2s}\frac{a(t)}{t}
dt \Big) ds\\
= KC+C\sum\limits_{k=1}^\infty\displaystyle\int_{\beta_{k-1}}^{\beta_k} \omega\{|f|>s\}\Big(
\displaystyle\int_{1}^{2s}\frac{a(t)}{t}
dt \Big) ds\\
\leq KC+2C\sum\limits_{k=1}^\infty
\omega(Q_k)\Big(\frac{1}{2^k}b(\frac{s_k}{2^k})
\Big)\beta_k \\
= KC+2C\sum\limits_{k=1}^\infty
\frac{1}{\frac{s_k}{2^k}b(\frac{s_k}{2^k})}
\Big(\frac{1}{2^k}b(\frac{s_k}{2^k})
\Big)\frac{ks_k}{2^k}\\
= KC+2C\sum\limits_{k=1}^\infty\frac{k}{2^k} <\infty.
$ |
On the other hand, we have
$
\displaystyle\int_{\Omega}{\it\Psi}(C_7\cdot |f|)\omega d\mu
=\sum\limits_{k=1}\limits^{\infty}{\it\Psi}
(C_7\cdot
\frac{ks_k}{2^k})\omega({Q_k})
=\sum\limits_{k=1}\limits^{\infty}{\it\Psi}
(C_7\cdot
\frac{ks_k}{2^k})\frac{1}{\frac{s_k}{2^k}b(\frac{s_k}{2^k})}.
$ |
For the above $C_7, $ we can choose a constant $k(C_7)>0$ such that $k\cdot C_7\geq2, ~k>k(C_7).$ Therefore,
$
\displaystyle\int_{\Omega}{\it\Psi}(C_7\cdot |f|)\omega
d\mu \geq \sum\limits_{k=k(C_7)}\limits^{\infty}{\it\Psi}
(C_7\cdot
\frac{ks_k}{2^k})\frac{1}{\frac{s_k}{2^k}b(\frac{s_k}{2^k})}\\
\geq \sum\limits_{k=k(C_7)}
\limits^{\infty}\displaystyle\int_{\frac{s_k}{2^k}}^{2\frac{s_k}{2^k}}
\frac{b(t)}{\frac{s_k}{2^k}b(\frac{s_k}{2^k})}dt
\geq\sum\limits_{k=k(C_7)}\limits^{\infty}1=\infty.
$ |
The result contradicts assumption $(1.9).$ This completes the proof.