1 Introduction

Let \({\mathbb {R}}^n\) be the n-dimensional Euclidean space and \(({\mathbb {R}}^n)^m={\mathbb {R}}^n\times \ldots {\mathbb {R}}^n.\) The multilinear fractional operator is defined by

$$\begin{aligned} I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)=\int _{({\mathbb {R}}^n)^{m}} \frac{f_{1}(y_{1})\ldots f_{m}(y_{m})}{|(x-y_{1}), \ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}, \end{aligned}$$
(1.1)

where \(0<\alpha <mn,m\in {\mathbb {N}}.\) 

The multilinear fractional integral \(I_{\alpha ,m}\) is a natural generalization of the classical fractional integral. Kenig and Stein [9] as well as Grafakos and Kalton [4] considered the boundedness of a family of related fractional integrals. In 2012, Iida, Sato, Sawano and Tanaka in [6] obtained the boundedness property of the Adams type for multilinear fractional integral operators. Let \(0<\alpha <mn\), \(1<p_{1},p_{2},\ldots ,p_{m}<\infty , 0<p\le p_{0}<\frac{n}{\alpha }, 0<q\le q_{0}<\infty .\) Suppose that

$$\begin{aligned} \frac{1}{p}=\frac{1}{p_{1}}+\ldots +\frac{1}{p_{m}}, \frac{1}{q_{0}}=\frac{1}{p_{0}}-\frac{\alpha }{n} ~\text {and}~\frac{q}{q_{0}}=\frac{p}{p_{0}}. \end{aligned}$$

Then there exists a constant \(C>0\) such that

$$\begin{aligned} \begin{aligned} \Vert I_{\alpha ,m}(\vec {f})\Vert _{{\mathcal {M}}^{q_{0}}_{q}}&\le C\sup _{Q\in {\mathcal {Q}}}|Q|^{1/p_{0}}\prod _{i=1}^{m} \bigg (\frac{1}{|Q|}\int _{Q}|f_{i}(y_{i})|^{p_{i}}dy_{i} \bigg )^{1/p_{i}}\\&=:\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}_{\vec {P}}^{p_{0}}}, \end{aligned} \end{aligned}$$
(1.2)

where \({\mathcal {M}}^{q_{0}}_{q}\) stands for the Morrey spaces. The right-hand side of (1.2) is named by “the multi-Morrey norm”, which is strictly smaller than m-fold product of the Morrey norms.

An interesting question arises. Can we obtain the similar results for the endpoint cases, that is \(p_{0}=\frac{n}{\alpha }\), \(p_{0}>\frac{n}{\alpha }\) and \(p_{i} = 1\) for some \(i\in \{1,\ldots , m\}\). In this paper, we prove that the questions above have a affirmative solution.

To state the main results of this paper, we need first to recall some necessary notations and notion.

Definition 1

For \(0< q\le p <\infty \), we say that a function f belongs to Morrey space \({\mathcal {M}}^{p}_{q}\) if

$$\begin{aligned} \Vert f\Vert _{{\mathcal {M}}^{p}_{q}}:=\sup _{x\in {\mathbb {R}}^{n},r>0}\frac{1}{|B(x,r)|^{1/q-1/p}} \bigg (\int _{B(x,r)}|f(y)|^{q}dy\bigg )^{1/q}<\infty ; \end{aligned}$$

a function f belongs to weak Morrey space \(WM_{q}^{p}\) if

$$\begin{aligned} \Vert f\Vert _{W{\mathcal {M}}_{q}^{p}}:=\sup _{x\in {\mathbb {R}}^{n},r>0}\frac{1}{|B(x,r)|^{1/q-1/p}} \sup _{\lambda>0}\bigg (\lambda ^{q}|\{y\in B(x,r):|f(y)|>\lambda \}|\bigg )^{1/q}<\infty . \end{aligned}$$

Morrey spaces seem to describe precisely the boundedness property of fractional integral operators. Morrey spaces describe local regularity more precisely than \(L^{p}\) spaces and can be seen as a complement of \(L^{p}\). In fact, \(L^p=M^{p}_{p}\subset M^{p}_{q}\) and \(WM^{p}_{p}=L^{p,\infty }\) for \(0<q \le p<\infty \).

Recall that the fractional integral operator (or the Riesz potential) \(I_{\alpha }\), \(0<\alpha <n\), is given by

$$\begin{aligned} I_{\alpha }(f)(x)=\int _{{\mathbb {R}}^n}\frac{f(y)}{|x-y|^{n-\alpha }}dy. \end{aligned}$$

The inequality

$$\begin{aligned} \Vert I_{\alpha }(f)\Vert _{{\mathcal {M}}^{q_{0}}_{q}}\le C\Vert f\Vert _{{\mathcal {M}}^{p_{0}}_{p}} \end{aligned}$$
(1.3)

with \(1/q_{0}=1/p_{0}-\alpha /n\) and \(1/q=1/p-\alpha /n\) was obtained by Spanne. The result was improved by Adams [1](see also [2]), the inequality (1.3) holds if \(1/q_{0}=1/p_{0}-\alpha /n\) and \(q/q_{0}=p/p_{0}\). And, Olsen [10] showed by an example that the result of Adams is optimal. Tang [13] extended the result to the multilinear fractional integral operator \({\mathcal {I}}_\alpha \). They also established some endpoint estimates for the multilinear fractional integral.

In the endpoint case \(p_{0} = n/\alpha \), the exponential integrability of \(I_{\alpha ,m}\) was proved by Strichartz [11] for \(m = 1\), by Tang [13] for \(m\ge 2\).

Theorem

(cf. [13]) Let \(m\in {\mathbb {N}}\), \(0<\alpha <mn, 1/p=1/p_{1}+\ldots +1/p_{m}=\alpha /n\) with \(1<p_{i}<\infty \) for \(i=1,\ldots , m\). Let B be a ball of radius R in \({\mathbb {R}}^n\) and let \(f_{j}\in L^{p_{j}}(B)\) be supported in B. Then there exist constants \(k_{1},k_{2}\) depending only on \(n,m,\alpha ,p\) and the \(p_{j}\) such that

$$\begin{aligned} \int _{B}\exp \bigg (k_{1}\Big (\frac{|I_{\alpha ,m}(f_{1}, \ldots ,f_{m})(x)|}{\prod _{j=1}^{k}\Vert f_{j}\Vert _{L^{p_{j}}(B)}} \Big )^{n/(mn-\alpha )}\bigg )dx\le k_{2}R^{n}. \end{aligned}$$
(1.4)

For the case \(p_{0}\ge n/\alpha \), we also study the boundedness for multilinear fractional integrals on spaces as BMO space and Lipschitz spaces. Campanato spaces are a useful tool in the regularity theory of PDEs due to their better structures, which allows us to give an integral characterization of the spaces of Hölder continuous functions.

Definition 2

Let \(0<p<\infty \) and \(-n/p<\beta <n\). A locally integrable function f is said to belong to Campanato space \({\mathcal {C}}_{\beta ,q}\) if there exists a constant \(C > 0\) such that for any ball \(B\subset {\mathbb {R}}^n\),

$$\begin{aligned} \frac{1}{|B|^{\beta /n}}\bigg (\frac{1}{|B|}\int _{B}|f(x) -f_{B}|^{p}dx\bigg )^{1/p}\le C, \end{aligned}$$

where \(f_{B}=\frac{1}{|B|}\int _{B}f(x)dx\) and the minimal constant C is defined by \(\Vert f\Vert _{{\mathcal {C}}_{\beta ,p}}\).

The Lipschitz (Hölder) and Campanato spaces are related by the following equivalences:

$$\begin{aligned} \Vert f\Vert _{Lip_{\beta }}:=\sup _{x,h\in {\mathbb {R}}^n,h\ne 0}\frac{|f(x+h)-f(x)|}{|h|^{\beta }}\approx \Vert f\Vert _{{\mathcal {C}}_{\alpha ,q}},\quad 0<\alpha <1. \end{aligned}$$

The equivalence can be found in [3] for \(q=1\), [7] for \(1<q<\infty \) and [15] for \(0<q<1\). Specially, \({\mathcal {C}}_{0,q}={{\textrm{BMO}}}\), the spaces of bounded mean oscillation. The crucial property of \({{\textrm{BMO}}}\) functions is the John-Nirenberg inequality [8],

$$\begin{aligned} |\{x\in Q: |f(x)-f_{Q}|>\lambda \}|\le c_{1}|Q|e^{-\frac{c_{2}\lambda }{\Vert f\Vert _{{\textrm{BMO}}}}}, \end{aligned}$$

where \(c_{1}\) and \(c_{2}\) depend only on the dimension. A well-known immediate corollary of the John-Nirenberg inequality as follows:

$$\begin{aligned} \Vert f\Vert _{{\textrm{BMO}}}\approx \sup _{Q}\frac{1}{|Q|}\Big (\int _{Q}|f(x)-f_{Q}|^{p}dx\Big )^{1/q}, \end{aligned}$$

for all \(1<q<\infty \). In fact, the equivalence also holds for \(0<q<1\). See, for example, the work of Strömberg [12](or [5] and [16] for the general case). In addition, we also proved in [14] that \(f\in {{\textrm{BMO}}}\) if and only if for \(0<q<\infty \),

$$\begin{aligned} \sup _{Q}\frac{\big \Vert (f-f_{Q})\chi _{Q} \big \Vert _{L^{q,\infty }}}{|Q|^{1/q}}<\infty . \end{aligned}$$
(1.5)

The main result of this paper are stated as follows.

Theorem 1.1

Let \(m\in {\mathbb {N}}\)\(0<\alpha <mn\) and \(I_{\alpha ,m}\) be as in (1.1). Let \(\vec {P}=(p_{1},\ldots ,p_{m}), 1\le p_{1},\ldots ,p_{m}<\infty \) with \(1/p=1/p_{1}+\ldots +1/p_{m}\) and \(0<p\le p_{0}<\infty \). If \(p_{0}=n/\alpha \) and \(f_{i}\in L^{p_{i}}\) with compact support, then for any ball \(B=B(x_{0},R)\), there exist constants \({\tilde{C}}, k_{1},k_{2}\) such that \(f_{1},\ldots , f_{m}\) is supported in \(B(x_{0},{\tilde{C}}R)\) and

$$\begin{aligned} \frac{1}{|B|}\int _{B}\exp \Big (\frac{k_{1}|I_{\alpha ,m}(f_{1}, \ldots ,f_{m})(x)|}{\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}} _{\vec {P}}}}\Big )dx\le k_{2}({\tilde{C}}+1)^{\alpha p}. \end{aligned}$$
(1.6)

Theorem 1.2

Let \(m\in {\mathbb {N}}\)\((m-1)n<\alpha <mn\) and \(I_{\alpha ,m}\) be as in (1.1). Let \(\vec {P}=(p_{1},\ldots ,p_{m}), 1\le p_{1},\ldots ,p_{m}<\infty , 1/p=1/p_{1}+\ldots +1/p_{m}\) and \(0<p\le p_{0}<\infty \). If \(p_{0}=n/\alpha ,\) then

$$\begin{aligned} \Vert I_{\alpha ,m}(f_{1},\ldots ,f_{m})\Vert _{{\textrm{BMO}}}\le C\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$

For the case \(n/\alpha<p_{0}<\infty \), we obtain the result as follows. We remark that when \(p_{0}=\infty \) in Theorem 1.3, the conclusion also holds. Indeed, the proof is similar to the case \(p_{0} < \infty \), moreover the proof is simpler.

Theorem 1.3

Let \(m\in {\mathbb {N}}\)\(0<\alpha <mn\) and \(I_{\alpha ,m}\) be as in (1.1). Let \(\vec {P}=(p_{1},\ldots ,p_{m}), 1\le p_{1},\ldots ,p_{m}<\infty \) and \(0<p\le p_{0}<\infty \). If \(p_{0}>n/\alpha \) and \(0<\alpha -n/p_{0}<1,\) then

$$\begin{aligned} \Vert I_{\alpha ,m}(f_{1},\ldots ,f_{m})\Vert _{Lip_{\alpha -n/p_{0}}}\le C\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {p}}}. \end{aligned}$$

If there is at least \(p_{i}\) is equiv to 1, one has the following weak type estimate.

Theorem 1.4

Let \(m\in {\mathbb {N}}\)\(0<\alpha <mn\) and \(I_{\alpha ,m}\) be as in (1.1). Let \(\vec {P}=(p_{1},\ldots ,p_{m})\), \(1\le p_{j}<\infty \) for \(j=1,\ldots , m\), \(1/p=1/p_{1}+\ldots +1/p_{m}\), \(1/q=1/q_{1}+\ldots +1/q_{m}\) with \(1/q=1/p-\alpha /n\) and \(1/q_{0}=1/p_{0}-\alpha /n\). If \(p_{0}<n/\alpha \) and there is at least one \(q_{i}\) which is equal to 1, we have

$$\begin{aligned} \Vert I_{\alpha ,m}(f_{1},\ldots ,f_{m})\Vert _{W{\mathcal {M}}^{q_{0}}_{q}}\le C\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$

Let |E| denote the Lebesgue measure of a measurable set \(E\subset {\mathbb {R}}^n\). Throughout this paper, the letter C denotes constants which are independent of main variables and may change from one occurrence to another. B(xr) denotes a ball centered at x, with side length r.

2 Proofs of Theorems 1.11.2

Proof of Theorem 1.1

Let us first assume \(\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}=1.\) For some \(\delta >0\), we have

$$\begin{aligned}&|I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\\&\quad \le \int _{|(x-y_{1},\ldots , x-y_{m})|\le \delta } \frac{|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|}{|(x-y_{1}, \ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}\\&\qquad +\int _{|(x-y_{1},\ldots , x-y_{m})|> \delta } \frac{|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|}{|(x-y_{1}, \ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}\\&\quad =:F_{1}+F_{2}. \end{aligned}$$

For \(F_{1}\), by a direct argument, we see that

$$\begin{aligned} F_{1}{} & {} =\sum _{j=1}^{m}\int _{2^{-j}\delta <|(x-y_{1}, \ldots , x-y_{m})|\le 2^{-j+1}\delta } \frac{|f(y_{1})|\ldots |f(y_{m})|}{|(x-y_{1}, \ldots , x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}\nonumber \\{} & {} \le \sum _{j=1}^{m}(2^{-j}\delta )^{\alpha -mn} \int _{|(x-y_{1},\ldots , x-y_{m})|\le 2^{-j+1} \delta }|f(y_{1})|\ldots |f(y_{m})|dy_{1}\ldots dy_{m}\nonumber \\{} & {} \le \sum _{j=1}^{m}(2^{-j}\delta )^{\alpha -mn} (2^{-j+1}\delta )^{mn}{\mathcal {M}}(f_{1},\ldots ,f_{m})(x)\nonumber \\{} & {} \le \frac{2^{mn}\delta ^{\alpha }}{2^{\alpha } -1}{\mathcal {M}}(f_{1},\ldots ,f_{m})(x),\nonumber \\ \end{aligned}$$
(2.1)

where \({\mathcal {M}}\) denotes the multilinear Hardy-Littlewood maximal function

$$\begin{aligned} {\mathcal {M}}(f_{1},\ldots ,f_{m})(x)=\sup _{B\ni x}\prod _{i=1}^{m}\frac{1}{|B|}\int _{B}|f_{i}(y_{i})|dy_{i}. \end{aligned}$$

For any \(x\in B=B(x_{0},R)\), there exist a ball \({\tilde{B}}=B(x_{0}, {\tilde{C}}R)\) with \({\tilde{C}}>0\) such that the functions \(f_{1}, \ldots , f_{m}\) are supported in \({\tilde{B}}\). We conclude that for any \(y_{i}\in {\tilde{B}}\), \(|x-y_{i}|\le |x-x_{0}|+|x_{0}-y_{i}|\le ({\tilde{C}}+1)R\) and

$$\begin{aligned} |(x-y_{1},\ldots ,x-y_{m})|\le ({\tilde{C}}+1)\sqrt{m}R, \end{aligned}$$

this shows that

$$\begin{aligned} F_{2}{} & {} \le \int _{\delta<|(x-y_{1},\ldots ,x-y_{m})|\le ({\tilde{C}}+1)\sqrt{m}R}\frac{|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|}{|(x-y_{1},\ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\cdots dy_{m}\nonumber \\{} & {} \le \sum _{j=1}^{J}\int _{2^{j}\delta <|(x-y_{1},\ldots ,x-y_{m})|\le 2^{j+1}\delta }\frac{|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|}{|(x-y_{1},\ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}\nonumber \\{} & {} \le \sum _{j=1}^{J}(2^{j}\delta )^{\alpha -mn}\int _{|(x-y_{1},\ldots ,x-y_{m})|\le 2^{j+1}\delta }|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|dy_{1}\ldots dy_{m}\nonumber \\{} & {} \le \sum _{j=1}^{J}(2^{j}\delta )^{\alpha -mn}\prod _{i=1}^{m}\int _{|x-y_{i}|\le 2^{j+1}\delta }|f_{i}(y_{i})|dy_{i}\nonumber \\{} & {} \le \sum _{j=1}^{J}(2^{j}\delta )^{\alpha -mn}(2^{j+1} \delta )^{\frac{n}{p'_{i}}}\prod _{i=1}^{m}\Big (\int _{|x-y_{i}|\le 2^{j+1}\delta }|f_{i}(y_{i})|^{p_{i}}dy_{i}\Big )^{1/p_{i}}\nonumber \\{} & {} \le 2^{mn-\alpha }J, \nonumber \\ \end{aligned}$$
(2.2)

where \(J=\log \frac{({\tilde{C}}+1)\sqrt{m}R}{\delta }.\)

Combining (2.1) and (2.2), we obtain for any \(x\in B\) and \(0<\delta \le ({\tilde{C}}+1)\sqrt{m}R\),

$$\begin{aligned} |I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\le \frac{2^{mn}}{2^{\alpha }-1}\delta ^{\alpha }{\mathcal {M}}(f_{1},\ldots ,f_{m})(x)+ 2^{mn-\alpha }\log \frac{({\tilde{C}}+1)\sqrt{m}R}{\delta }. \end{aligned}$$
(2.3)

In particular, the choice of \(\delta =({\tilde{C}}+1)\sqrt{m}R\) yields for all \(x\in B\),

$$\begin{aligned} |I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\le \frac{2^{mn}}{2^{\alpha }-1}{\mathcal {M}}(f_{1},\ldots ,f_{m})(x). \end{aligned}$$

Therefore, the election of

$$\begin{aligned} \delta =\delta (x)=\epsilon \Big (\frac{|I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|}{ \frac{2^{mn}}{2^{\alpha }-1}{\mathcal {M}}(f_{1}, \ldots ,f_{m})(x)}\Big )^{\frac{1}{\alpha }} \end{aligned}$$

for all \(\epsilon <1\). Now, (2.3) implies that

$$\begin{aligned}{} & {} |I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\nonumber \\{} & {} \quad \le \epsilon ^{\alpha }|I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\nonumber \\{} & {} \qquad + \frac{2^{mn-\alpha }}{\alpha p}\log \frac{\big (\frac{2^{mn} \sqrt{m}}{\epsilon (2^{\alpha }-1)}\big )^{p}({\tilde{C}}+1)^{\alpha p} R^{\alpha p}{\mathcal {M}}(f_{1},\ldots ,f_{m})(x)^{p}}{|I_{\alpha ,m} (f_{1},\ldots ,f_{m})(x)|^{p}} \nonumber \\ \end{aligned}$$
(2.4)

If we use the notation \(k_{1}=\frac{(1-\epsilon )^{\alpha }\alpha p}{2^{mn-\alpha }}\) and \(C_{1}=\big (\frac{2^{mn}\sqrt{m}}{\epsilon (2^{\alpha }-1)}\big )^{p}\), (2.4) is equivalent to

$$\begin{aligned} k_{1}|I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\le \log \frac{C_{1} ({\tilde{C}}+1)^{\alpha p}R^{\alpha p}{\mathcal {M}}(f_{1},\ldots ,f_{m})(x)^{p}}{|I_{\alpha ,m} (f_{1},\ldots ,f_{m})(x)|^{p}}. \end{aligned}$$
(2.5)

By exponentiating (2.5), we get

$$\begin{aligned} \exp \{k_{1}|I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\}\le \frac{C_{1}({\tilde{C}}+1)^{\alpha p}R^{\alpha p}{\mathcal {M}}(f_{1},\ldots ,f_{m})(x)^{p}}{|I_{\alpha ,m} (f_{1},\ldots ,f_{m})(x)|^{p}}. \end{aligned}$$
(2.6)

Let \(B_{1}=\{x\in B: |I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\ge 1\}\) and \(B_{2}=B\backslash B_{1}\). By the inequality (2.6),

$$\begin{aligned} \int _{B_{1}}\exp \{k_{1}|I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)|\}dx&\le C_{1}({\tilde{C}}+1)^{\alpha p}R^{\alpha p}\int _{B_{1}} {\mathcal {M}}(f_{1},\ldots ,f_{m})(x)^{p}dx\\&\le C_{1}({\tilde{C}}+1)^{\alpha p} R^{n}\Vert {\mathcal {M}}(f_{1}, \ldots ,f_{m})\Vert ^{p}_{{\mathcal {M}}^{p_{0}}_{p}}\\&\le C_{2}({\tilde{C}}+1)^{\alpha p}R^n, \end{aligned}$$

where \(C_{2}=C_{1}\Vert {\mathcal {M}}\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}\rightarrow {\mathcal {M}}^{p_{0}}_{p}}\). On the other hand,

$$\begin{aligned} \int _{B_{2}}\exp \{k_{1}|I_{\alpha ,m}(f_{1},\ldots ,f_{m}) (x)|\}dx\le \exp \{k_{1}\}|B_{2}|. \end{aligned}$$
(2.7)

Thus, adding the integrals above over \(B_1\) and \(B_2\),

$$\begin{aligned} \frac{1}{|B|}\int _{B}\exp \{k_{1}|I_{\alpha ,m}(f_{1}, \ldots ,f_{m})(x)|\}dx\le k_{2}({\tilde{C}}+1)^{\alpha p}, \end{aligned}$$

where \(k_{2}=(\exp \{k_{1}\}+C_{2})\omega _{n}\) and \(\omega _{n}=|B(0,1)|\).

For the general \(\vec {f}=(f_{1},\ldots ,f_{m})\), we can use the natation \({\tilde{f}}_{1}=f_{1}/\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}} _{\vec {P}}}\). Obviously,

$$\begin{aligned} I_{\alpha ,m}(\vec {f}_{1},f_{2},\ldots ,f_{m})=I_{\alpha ,m} (f_{1},f_{2},\ldots ,f_{m})/\Vert (f_{1},\ldots ,f_{m}) \Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$

Therefore, we obtain the desired inequality (1.6). \(\square \)

Proof of Theorem 1.2

The inequality (1.5) implies that for \(0<q<\infty ,\) \(f\in {{\textrm{BMO}}}\) if and only if there exists a constant \(c_{B}\) related to the ball B,  such that

$$\begin{aligned} \sup _{B}\frac{\big \Vert |f(x)-c_{B}|\chi _{B} \big \Vert _{L^{q,\infty }}}{|B|^{1/q}}<\infty . \end{aligned}$$

Then, we need only to prove that for any ball \(B=B(x_{0},r)\),

$$\begin{aligned} \frac{\big \Vert |I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)-c_{B}| \chi _{B}\big \Vert _{L^{q,\infty }}}{|B|^{1/q}}\le C\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$
(2.8)

Set \(\Omega (x_{0},r):=\{(y_{1},\ldots ,y_{m}): |x_{0}-y_{1}|+\ldots +|x_{0}-y_{m}|\le r\},\) 

$$\begin{aligned} {{\textrm{I}}}_{1}^{B}(x)&:=\int _{\Omega (x_{0},2^mr)}\frac{f_{1}(y_{1}) \ldots f_{m}(y_{m})}{|(x-y_{1},\ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m},\\ {{\textrm{I}}}_{2}^{B}(x)&:=\int _{\Omega ^{c}(x_{0},2^mr)}\frac{f_{1}(y_{1}) \ldots f_{m}(y_{m})}{|(x-y_{1},\ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m} \end{aligned}$$

and

$$\begin{aligned} c_{B}:=\frac{1}{|B|}\int _{B}I_{2}^{B}(z)dz. \end{aligned}$$

Then \(I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)=I_{1}^{B}(x)+I_{2}^{B}(x).\)

Since \(\Omega (x_{0},2r)\subset B(x_{0},2r)\times \ldots \times B(x_{0},2r)\) and the boundedness of \(I_{\alpha ,m}\) from \(L^{p_{1}}\times \ldots \times L^{p_{m}}\) to \(L^{q,\infty }\) with \(1/p_{1}+\ldots +1/p_{m}-1/q=\alpha /n\), we arrive at

$$\begin{aligned} \Vert {{\textrm{I}}}_{1}^{B}\Vert _{L^{q,\infty }}{} & {} \le \big \Vert |I_{\alpha ,m} (|f_{1}|\chi _{4B},\ldots ,|f_{m}|\chi _{4B})|\big \Vert _{L^{q,\infty }}\nonumber \\{} & {} \le C\prod _{i=1}^{m}\Vert f_{1}\chi _{4B}\Vert _{L^{p_{i}}}\nonumber \\{} & {} \le C|B|^{-1/q}\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}.\nonumber \\ \end{aligned}$$
(2.9)

On the other hand, for \(x,z\in B, (y_{1},\ldots ,y_{m})\in \Omega ^{c}(x_{0},2^mr)\), by the direct calculation, we get

$$\begin{aligned} \begin{aligned}&\bigg |\frac{1}{|(x-y_{1},\ldots ,x-y_{m})|^{mn-\alpha }} -\frac{1}{|(z-y_{1},\ldots ,z-y_{m})|^{mn-\alpha }}\bigg |\\&\le \frac{C|x-z|}{|(z-y_{1},\ldots ,z-y_{m})|^{mn-\alpha +1}} \le \frac{Cr}{|(z-y_{1},\ldots ,z-y_{m})|^{mn-\alpha +1}}. \end{aligned} \end{aligned}$$

Using this observation, we see that

$$\begin{aligned} \begin{aligned}&|{{\textrm{I}}}_{2}^{B}(x)-c_{B}|\\&\quad \le \frac{Cr}{|B|}\int _{B}\int _{\Omega ^{c}(x_{0},2^mr)} \frac{|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|}{|(z-y_{1}, \ldots ,z-y_{m})|^{mn-\alpha +1}}dy_{1}\ldots dy_{m}dz\\&\quad \le \frac{Cr}{|B|}\int _{B}\sum _{k=1}^{\infty }\int _{\Omega (x_{0}, 2^{(k+1)m}r)\backslash \Omega (x_{0},2^{km}r)} \frac{|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|}{|(z-y_{1}, \ldots ,z-y_{m})|^{mn-\alpha +1}}dy_{1}\ldots dy_{m}dz\\&\quad \le C\sum _{k=1}^{\infty }\frac{r}{(2^{km}r)^{mn-\alpha +1}} \prod _{i=1}^{m}\int _{B(x_{0},2^{(k+1)m}r)}|f_{i}(y_{i})|dy_{i}\\&\quad \le C\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}\nonumber \\ \end{aligned}$$
(2.10)

The inequality (2.10) shows that

$$\begin{aligned} \big \Vert |{{\textrm{I}}}_{2}^{B}(x)-c_{B}|\chi _{B}\big \Vert _{L^{q,\infty }}\le C|B|^{1/q}\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$
(2.11)

Combining with the estimates (2.9) and (2.11), we obtain the inequality  (2.8). \(\square \)

3 Proofs of Theorems 1.31.4

Proof of Theorem 1.3

For any \(x_{1},x_{2}\in {\mathbb {R}}^n,\) and \(r=|x_{1}-x_{2}|\), we write

$$\begin{aligned} \Omega (x_{1},r):=\{(y_{1},\ldots ,y_{m}): |x_{1}-y_{1}|+\ldots +|x_{1}-y_{m}|\le r\}, \end{aligned}$$

it follows that

$$\begin{aligned} \begin{aligned}&|I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x_{1})-I_{\alpha ,m}(f_{1}, \ldots ,f_{m})(x_{2})|\\&\quad \le \Big |\int _{\Omega (x_{1},2^{m}r)}\frac{f_{1}(y_{1}) \ldots f_{m}(y_{m})dy_{1}\ldots dy_{m}}{|(x_{1}-y_{1}, \ldots ,x_{1}-y_{m})|^{mn-\alpha }}\Big |\\&\qquad +\Big |\int _{\Omega (x_{1},2^{m}r)}\frac{f_{1}(y_{1}) \ldots f_{m}(y_{m})dy_{1}\ldots dy_{m}}{|(x_{2}-y_{1},\ldots ,x_{2}-y_{m})|^{mn-\alpha }}\Big |\\&\qquad +\Big |\int _{\Omega ^{c}(x_{1},2^{m}r)}\frac{f_{1}(y_{1}) \ldots f_{m}(y_{m})dy_{1}\ldots dy_{m}}{|(x_{1}-y_{1}, \ldots ,x_{1}-y_{m})|^{mn-\alpha }}\\&\qquad -\int _{\Omega ^{c}(x_{1},2^{m}r)} \frac{f_{1}(y_{1})\ldots f_{m}(y_{m})dy_{1}\ldots dy_{m}}{|(x_{2}-y_{1}, \ldots ,x_{2}-y_{m})|^{mn-\alpha }}\Big |\\&\quad =:{{\textrm{II}}}_{1}+{{\textrm{II}}}_{2}+{{\textrm{II}}}_{3}. \end{aligned} \end{aligned}$$
(3.1)

The same argument as (2.9), we have

$$\begin{aligned} {{\textrm{II}}}_{1}, {{\textrm{II}}}_{2}\le Cr^{\alpha -n/p_{0}}\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$
(3.2)

For the last term \({{\textrm{II}}}_{3},\) we also obtain

$$\begin{aligned} \begin{aligned} {{\textrm{II}}}_{3}&\le Cr\int _{\Omega ^{c}(x_{1},2^mr)} \frac{|f_{1}(y_{1})|\ldots |f_{m}(y_{m})|}{|(x_{1}-y_{1}, \ldots ,x_{1}-y_{m})|^{mn-\alpha +1}}dy_{1}\ldots dy_{m}\\&\le C\sum _{k=1}^{\infty }\frac{r}{(2^{km}r)^{mn-\alpha +1}} \prod _{i=1}^{m}\int _{B(x_{0},2^{(k+1)m}r)}|f_{i}(y_{i})|dy_{i}\\&\le Cr^{\alpha -n/p_{0}}\Vert (f_{1},\ldots ,f_{m}) \Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned} \end{aligned}$$
(3.3)

Combining with the estimates (3.1),(3.2) and (3.3), we arrive at

$$\begin{aligned}{} & {} |I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x_{1})-I_{\alpha ,m} (f_{1},\ldots ,f_{m})(x_{2})|\\{} & {} \quad \le C|x_{1}-x_{2}|^{\alpha -n/p}\Vert (f_{1},\ldots ,f_{m}) \Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$

Thus, Theorem 1.3 is proved. \(\square \)

Proof of Theorem 1.4

For any \(x_{0}\in {\mathbb {R}}^n\) and \(B=B(x_{0},r)\) centered at \(x_{0}\) with radius r, set \(\Omega (x_{0},r):=\{(y_{1},\ldots ,y_{m}): |x_{0}-y_{1}|+\cdots +|x_{0}-y_{m}|\le r\}.\) For any \(x\in B\), we split \(I_{\alpha ,m}(f_{1},\ldots ,f_{m})(x)\) with

$$\begin{aligned} {{\textrm{III}}}_{1}^{B}(x):=\int _{\Omega (x_{0},4r)}\frac{f_{1}(y_{1})\ldots f_{m}(y_{m})}{|(x-y_{1},\ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}, \end{aligned}$$

and

$$\begin{aligned} {{\textrm{III}}}_{2}^{B}(x):=\int _{\Omega ^{c}(x_{0},4r)} \frac{f_{1}(y_{1})\ldots f_{m}(y_{m})}{|(x-y_{1},\ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}. \end{aligned}$$

By the boundedness of \(I_{\alpha ,m}\) from \(L^{p_{1}}\times \cdots \times \cdots L^{p_{m}}\) to \(L^{q,\infty }\) and the fact that

$$\begin{aligned} {{\textrm{III}}}_{1}^{B}(x)\le I_{\alpha }(|f^{0}_{1}|,\ldots ,|f^{0}_{m}|)(x) \end{aligned}$$

with \(f^{0}_{i}=f_{i}\chi _{4B}\), we can easily get

$$\begin{aligned} \Vert {{\textrm{III}}}_{1}^{B}\Vert _{L^{q,\infty }(B(x_{0},r))}&\le \Vert I_{\alpha }(f^{0}_{1},\ldots ,f^{0}_{m})\Vert _{L^{q,\infty }(B(x_{0},r))}\\&\le \Vert I_{\alpha ,m}(|f^{0}_{1}|,\ldots ,|f^{0}_{m}|)\Vert _{L^{q,\infty }} \le C\prod _{i=1}^{m}\Vert f_{i}\Vert _{L^{p_{i}}(B(x_{0},4r))}. \end{aligned}$$

By \(\frac{1}{q_{0}}=\frac{1}{p_{0}}-\frac{\alpha }{n}\) and \(\frac{1}{q}=\frac{1}{p_{1}}+\ldots +\frac{1}{p_{m}}-\frac{\alpha }{n}\), one has

$$\begin{aligned} |B|^{\frac{1}{q_{0}}-\frac{1}{q}}\Vert {{\textrm{III}}}_{1}^{B}\Vert _{L^{q,\infty }(B(x_{0},r))}\le C|4B|^{\frac{1}{p_{0}}-\frac{1}{p}}\prod _{i=1}^{m} \Vert f_{i}\Vert _{L^{p_{i}}(B(x_{0},4r))} \end{aligned}$$

and \(\Vert {{\textrm{III}}}_{1}^{B}\Vert _{W{\mathcal {M}}^{q_{0}}_{q}}\le C\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}\).

For the term \({{\textrm{III}}}^{B}_{2}\), Hölder inequality shows that

$$\begin{aligned} |{{\textrm{III}}}_{2}^{B}(x)|&\le \sum _{j=1}^{\infty } \int _{\Omega (x,2^{j+1}r)\backslash \Omega (x,2^{j}r)} \frac{|f_{1}(y_{1})\ldots f_{m}(y_{m})|}{|(x-y_{1}, \ldots ,x-y_{m})|^{mn-\alpha }}dy_{1}\ldots dy_{m}\\&\le \sum _{j=1}^{\infty }(2^{j}r)^{-mn+\alpha } \int _{\Omega (x,2^{j+1}r)\backslash \Omega (x,2^{j}r)}|f_{1}(y_{1})| \ldots |f_{m}(y_{m})|dy_{1}\ldots dy_{m}\\&\le C \sum _{j=1}^{\infty }(2^{j}r)^{-\frac{n}{p}+\alpha } \Vert f_{j}\Vert _{L^{p_{i}}(B(x,2^{j+1}r)}\\&\le C\sum _{j=1}^{\infty }(2^{j}r)^{\alpha -\frac{n}{p_{0}}} \Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}\\&\le C r^{\alpha -\frac{n}{p_{0}}}\Vert (f_{1},\ldots ,f_{m}) \Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}. \end{aligned}$$

Thus, \(\Vert {{\textrm{III}}}_{2}^{B}\Vert _{W{\mathcal {M}}^{q_{0}}_{q}}\le C\Vert (f_{1},\ldots ,f_{m})\Vert _{{\mathcal {M}}^{p_{0}}_{\vec {P}}}\) and the theorem has been proved. \(\square \)