1 Introduction

It is well known that Hardy space \(H^{p}(\mathbb R^{n})\) has much better functional properties than the space \(L^{p}(\mathbb R^{n})\) for \(p<1\) [5]. However, as Goldberg in [14] pointed out that \(H^{p}(\mathbb {R}^{n})\) space is well suited only to the Fourier analysis, and is not stable under multiplications by the Schwartz test functions. One reason is that \(H^{p}(\mathbb {R}^{n})\) does not contain \(\mathcal {S}(\mathbb {R}^{n})\), the space of the Schwartz test functions. To circumvent those drawbacks, Goldberg in [14] introduced the local Hardy spaces \(h^{p}(\mathbb R^{n}),0<p<\infty \). Let \(\Phi \in \mathcal {S}(\mathbb {R}^{n})\) with \(\int \Phi \ne 0,\Phi _{t}(x)=t^{-n}\Phi (\frac{x}{t})\) and set

$$\begin{aligned} \mathcal {M}_{\Phi }(f)(x)=\sup _{0<t<1}|\Phi _{t}*f(x)|. \end{aligned}$$

Then \(h^{p}(\mathbb {R}^{n})=\{f\in \mathcal {S}'(\mathbb {R}^{n}),\mathcal {M}_{\Phi }(f)\in L^{p}(\mathbb {R}^{n})\},\) where \(\mathcal {S}'(\mathbb {R}^{n})\) is the dual of \(\mathcal {S}(\mathbb {R}^{n})\).

Goldberg showed that the dual of \(h^{1}(\mathbb {R}^{n})\) is \(bmo(\mathbb {R}^{n})\), which is defined as the set of \(f\in L^{1}_{\mathrm {loc}}(\mathbb {R}^{n})\) such that

$$\begin{aligned} S_{1}(f)=\sup _{|Q|<1}\frac{1}{|Q|}\int _{Q}|f(x)-f_{Q}|\mathrm{d}x<\infty \end{aligned}$$

and

$$\begin{aligned} S_{2}(f)=\sup _{|Q|\ge 1}\frac{1}{|Q|}\int _{Q}|f(x)|\mathrm{d}x<\infty , \end{aligned}$$

equipped with the norm \(\Vert f\Vert _{bmo(\mathbb {R}^{n})}=\max \{S_{1}(f),S_{2}(f)\}\), where \(f_{Q}\) is the mean of f over Q, i.e., \(f_{Q}=\frac{1}{|Q|}\int _{Q}f(x)\mathrm{d}x\). For \(0<r<1,\) let \(\Lambda _{r}=\{f\in L^{\infty }(\mathbb {R}^{n}): \sup _{x\ne y}\frac{|f(x)-f(y)|}{|x-y|^{r}}<\infty \}\). Then, it is well known that the dual of \(h^{p}(\mathbb {R}^{n})\) is \(\Lambda _{n(\frac{1}{p}-1)}\) for \(\frac{n}{n+1}<p<1\). We refer the reader to [14] for more details of \(\Lambda _{r}\) when \(r\ge 1\). It is convenient to denote \(\Lambda _{0}=bmo(\mathbb {R}^{n})\). For more results about local Hardy spaces, we refer the reader to [1,2,3, 18, 19, 25,26,27,28,29, 31,32,33]. Some recent developments of multi-parameter local Hardy spaces can be seen in [4, 10, 11].

In [31], Rychkov obtained that \(h^{p}(\mathbb {R}^{n})\) can be characterized by a continuous Littlewood–Paley–Stein square function. More precisely, let

$$\begin{aligned} \varphi _{0}\in \mathcal {D}({\mathbb {R}^{n}}) \hbox { with nonzero integral, and } \varphi (x)=\varphi _{0}(x)-2^{-n}\varphi _{0}(\frac{x}{2}), \end{aligned}$$
(1.1)

where \(\mathcal {D}({\mathbb {R}^{n}})\) is the set of all smooth functions with compact support on \(\mathbb {R}^{n}\), then for any \(N\ge 0\), there exist two functions \(\phi _{0}, \phi \in \mathcal {D}({\mathbb {R}^{n}})\) such that \(\phi \) has vanishing moments up to order N (i.e., \(\int x^{\alpha }\phi (x)\mathrm{d}x=0\) for all multi-indices with \(|\alpha |\le N\)) and

$$\begin{aligned} f(x)=\sum \limits _{j\in \mathbb {N}}\phi _{j}*\varphi _{j}*f(x),\ \hbox { in}\ \mathcal {D}'({\mathbb {R}^{n}}), \end{aligned}$$
(1.2)

where \(\mathcal {D}'({\mathbb {R}^{n}})\) is dual space of \(\mathcal {D}({\mathbb {R}^{n}})\), \(\mathbb {N}\) is the set of all natural numbers. Here and the following, we use dyadic dilations defined by \(g_{j}(x)=2^{jn} g(2^jx)\) for \(j\in \mathbb {N}, j\ge 1\), and \(g_{j}(x)\) for \(j=0\) is just the value of a function \(g_{0}\). For any \(j\in \mathbb Z\), denote \(\Pi _{j}=\{Q: \) Q are dyadic cubes in \(\mathbb R^{n}\) with the side length \(l(Q)=2^{-j}\), and the left lower corners of Q are \(x_{Q}=2^{-j}\ell \), \(\ell \in \mathbb Z^{n}\}\), and \(\Pi =\cup _{j\in \mathbb {N}}\Pi _{j}\). At last, denote \(\bar{\Pi }=\{Q: \) Q are cubes in \(\mathbb R^{n}\}\). Using continuous local Calderón’s identity (1.2), Rychkov proved that \(f\in h^{p}(\mathbb {R}^{n})\) if and only if

$$\begin{aligned} \left\| \left( \sum _{j=0}^{\infty }|\varphi _{j}*f|^{2}\right) ^{1/2}\right\| _{L^{p}(\mathbb {R}^{n})}<\infty . \end{aligned}$$
(1.3)

Moreover, the continuous local Calderón’s identity (1.2) also holds in \(\mathcal {S}({\mathbb {R}^{n}})\), \(\mathcal {S}'({\mathbb {R}^{n}})\). We want to remark that using the fact pointed out in p160 of [31], the vanishing moments of \(\varphi \) in (1.3) can be up to any fixed order.

Recently, the authors in [9] obtained the discrete Littlewood–Paley–Stein characterization of \(h^{p}(\mathbb {R}^{n})\). More precisely, let \(\psi _{0},\psi \in {\mathcal S}(\mathbb R^{n})\) with

$$\begin{aligned} \hbox {supp} \widehat{\psi _{0}}\subseteq \{\xi \in \mathbb R^{n}: |\xi |\le 2\};\ \widehat{\psi _{0}}(\xi )=1,\ \hbox {if}\ |\xi |\le 1, \end{aligned}$$
(1.4)

and

$$\begin{aligned} \hbox {supp} \widehat{\psi }\subseteq \{\xi \in \mathbb R^{n}: \frac{1}{2}\le |\xi |\le 2\}, \end{aligned}$$
(1.5)

and

$$\begin{aligned} |\widehat{\psi _{0}}(\xi )|^{2}+\sum \limits ^{\infty }_{j =1}|\widehat{\psi }(2^{-j}\xi )|^2=1, \ \hbox { for all } \xi \in \mathbb R^{n}, \end{aligned}$$
(1.6)

then one has the following continuous Calderón reproducing formula

$$\begin{aligned} \qquad f(x)= & {} \sum \limits _{j=0}^{+\infty } \psi _{j}*\psi _j*f(x), \end{aligned}$$
(1.7)

and the discrete Calderón reproducing formula is the following

$$\begin{aligned} \qquad f(x)= & {} \sum \limits _{j=0}^{+\infty } \sum \limits _{Q\in \Pi _{j}}|Q|(\psi _{j}*f)(x_{Q})\times \psi _{j}(x-x_{Q}), \end{aligned}$$
(1.8)

where the both series converge in \(L^{2}(\mathbb R^{n})\), \(\mathcal S (\mathbb R^{{n}})\) and \(\mathcal S'(\mathbb R^{{n}})\) [13]. It is proved that in [9] \(f\in h^{p}(\mathbb {R}^{n})\) if and only if

$$\begin{aligned} \left\| \Big (\sum _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j}}|\psi _{j}*f(x_{Q})|^2\chi _{Q}(x)\Big )^{\frac{1}{2}}\right\| _{L^p(\mathbb {R}^{n})}<+\infty , \end{aligned}$$

where \(\psi _{0}\) and \(\psi \in {\mathcal S}(\mathbb R^{n})\) are functions satisfying conditions (1.4)–(1.6). Moreover, they proved that \(cmo^{p}(\mathbb {R}^{n})\) defined as following is also the dual space of \(h^{p}(\mathbb {R}^{n})\).

Definition 1.1

Let \(0<p\le 1\). Suppose that \(\psi _{0}\) and \(\psi \in {\mathcal S}(\mathbb R^{n})\) are functions satisfying conditions (1.4)–(1.6). \(cmo^{p}(\mathbb {R}^{n})\) is defined by

$$\begin{aligned} cmo^{p}(\mathbb {R}^{n})=\{f\in \mathcal S'(\mathbb {R}^{n}):\ \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}<\infty \}, \end{aligned}$$

where

$$\begin{aligned} \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}= \sup \limits _{P\in \bar{\Pi }}\Big (\frac{1}{|P|^{\frac{2}{p}-1}} \int _{P}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}(|\psi _{j}*f(x_{Q})|^2 \chi _{Q}(x)) \mathrm{d}x\Big )^{1/2}.\nonumber \\ \end{aligned}$$
(1.9)

Obviously, \(cmo^{p}(\mathbb {R}^{n})\) is coincident with \(\Lambda _{n(\frac{1}{p}-1)}\) for \(\frac{n}{n+1}<p\le 1\) since they are all the dual spaces of \(h^{p}(\mathbb {R}^{n})\). A natural question is: can \(cmo^{p}(\mathbb {R}^{n})\) be coincident with \(\Lambda _{n(\frac{1}{p}-1)}\) for \(\frac{n}{n+1}<p\le 1\) by norms? In this paper, we give a positive answer.

Theorem 1.1

Suppose that \(\psi _{0}\) and \(\psi \in {\mathcal S}(\mathbb R^{n})\) are functions satisfying conditions (1.4)–(1.6). Then for \(\frac{n}{n+1}<p\le 1\), \(f\in \Lambda _{n(\frac{1}{p}-1)}\) if and only if \(f\in cmo^{p}(\mathbb {R}^{n})\). Moreover, \(\Vert f\Vert _{\Lambda _{n(\frac{1}{p}-1)}}\approx \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}.\)

Furthermore, we give another equivalent norm of \(cmo^{p}(\mathbb {R}^{n})\) which has a very simple form.

Definition 1.2

Let \(0<p\le 1\). Suppose that \(\psi _{0}\) and \(\psi \in {\mathcal S}(\mathbb R^{n})\) are functions satisfying conditions (1.4)–(1.6). \(Lip^{p}(\mathbb {R}^{n})\) is defined by

$$\begin{aligned} Lip^{p}(\mathbb {R}^{n})=\{f\in \mathcal S'(\mathbb {R}^{n}):\ \Vert f\Vert _{Lip^{p}(\mathbb {R}^{n})}<\infty \}, \end{aligned}$$

where

$$\begin{aligned} \Vert f\Vert _{Lip^{p}(\mathbb {R}^{n})}= \sup \limits _{j\ge 0,x\in \mathbb {R}^n}2^{jn(\frac{1}{p}-1)}|\psi _{j}*f(x)|. \end{aligned}$$
(1.10)

We remark that using continuous local Calderón’s identity (1.2), the test functions in (1.10) can be replaced by those given in (1.1).

Theorem 1.2

Suppose that \(\psi _{0}\) and \(\psi \in {\mathcal S}(\mathbb R^{n})\) are functions satisfying conditions (1.4)–(1.6). Then for \(0<p\le 1\), \(f\in Lip^{p}(\mathbb {R}^{n})\) if and only if \(f\in cmo^{p}(\mathbb {R}^{n})\). Moreover, \(\Vert f\Vert _{Lip^{p}(\mathbb {R}^{n})}\approx \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}.\)

Since \(Lip^{p}(\mathbb {R}^{n})\) is decreasing as p increases, we have the following corollary.

Corollary 1.3

For \(0<p\le 1\), \(cmo^{p}(\mathbb {R}^{n})\) is increasing as p increases. Precisely, for \(0<p_{1}<p_{2}\le 1\), we have

$$\begin{aligned} \Vert f\Vert _{cmo^{p_{2}}(\mathbb {R}^{n})}\le C\Vert f\Vert _{cmo^{p_{1}}(\mathbb {R}^{n})}, \end{aligned}$$

where C is a constant independent of f.

The organization of this paper is as follows. In Sect. 2 we establish the equivalence between \(cmo^{p}(\mathbb {R}^{n})\) and \(Lip^{p}(\mathbb {R}^{n})\) for \(0<p\le 1\). The proof of Theorem 1.1 is presented in Sect. 3. For this, we prove that \(Lip^{p}(\mathbb {R}^{n})\) identifies \(\Lambda _{n(\frac{1}{p}-1)}\) with equivalent norms for \(\frac{n}{n+1}<p<1\) and \(\Lambda _{0}\) equals to \(cmo^{1}(\mathbb {R}^{n})\) with equivalent norms, that is, Theorems 3.1 and 3.2, respectively. As an application of equivalent theorems, we discuss the boundedness of inhomogeneous para-product operators on \(h^{p}(\mathbb {R}^{n})\) in last section.

Finally, we make some conventions. Throughout the paper, C denotes a positive constant that is independent of the main parameters involved, but whose value may vary from line to line. Constants with subscript, such as \(C_{1}\), do not change in different occurrences. We denote \(f\le Cg\) by \(f\lesssim g\). If \(f\lesssim g \lesssim f\), we write \(f\approx g\).

2 Equivalence Between \(cmo^{p}(\mathbb {R}^{n})\) and \(Lip^{p}(\mathbb {R}^{n})\)

In this section, we prove the equivalence between \(cmo^{p}(\mathbb {R}^{n})\) and \(Lip^{p}(\mathbb {R}^{n})\) for \(0<p\le 1\), that is Theorem 1.2.

We first prove

$$\begin{aligned} \Vert f \Vert _{cmo^{p}(\mathbb {R}^{n})}\lesssim \Vert f \Vert _{Lip^{p}(\mathbb {R}^{n})}. \end{aligned}$$
(2.1)

Firstly for any fixed \(P\in \bar{\Pi }\),

$$\begin{aligned}&\frac{1}{|P|^{\frac{2}{p}-1}} \int _{P}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}(|\psi _{j}*f(x_{Q})|^2 \chi _{Q}(x)) \mathrm{d}x \\&\quad =\frac{1}{|P|^{\frac{2}{p}-1}}\sum _{j\in N}\sum _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f(x_{Q})|^{2}. \end{aligned}$$

We consider two cases \(|P|<1\) and \(|P|\ge 1\). When \(|P|<1\) and \(P\in \bar{\Pi }\), there exists \(j_{0}\in \mathbb {N}\) such that \(2^{-j_{0}}\le \ell (P)<2^{-j_{0}+1}\). Then

$$\begin{aligned}&\frac{1}{|P|^{\frac{2}{p}-1}}\sum _{j\in N}\sum _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f(x_{Q})|^{2}\\&\quad \le \Vert f\Vert ^{2}_{Lip^{p}(\mathbb {R}^{n})}|P|^{1-\frac{2}{p}}\sum _{j\ge j_{0}}\sum _{Q\in \Pi _{j},Q\subseteq P} |Q|2^{-2jn(\frac{1}{p}-1)}:=A. \end{aligned}$$

Fix \(\tau \in (0,1)\), then for any \(p\in (0,1]\),

$$\begin{aligned} A\le & {} \Vert f\Vert _{Lip^{p}(\mathbb {R}^{n})}|P|^{2-\tau -\frac{2}{p}}\sum _{j\ge j_{0}}2^{-jn\tau }2^{-2jn(\frac{1}{p}-1)}\\\lesssim & {} \Vert f\Vert _{Lip^{p}(\mathbb {R}^{n})}|P|^{2-\tau -\frac{2}{p}}2^{-j_{0}n\tau }2^{-2j_{0}n(\frac{1}{p}-1)}\lesssim \Vert f\Vert ^{2}_{Lip^{p}(\mathbb {R}^{n})}, \end{aligned}$$

If \(|P|\ge 1\),

$$\begin{aligned}&\frac{1}{|P|^{\frac{2}{p}-1}}\sum _{j\in \mathbb {N}}\sum _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f(x_{Q})|^{2}\\&\quad =\frac{1}{|P|^{\frac{2}{P}-1}}\sum _{j\in \mathbb {N}}\sum _{Q\in \Pi _{j},Q\subseteq P} |Q|(2^{jn(\frac{1}{p}-1)}|\psi _{j}*f(x_{Q})|)^{2}2^{-2jn(\frac{1}{p}-1)}\\&\quad \le \Vert f\Vert ^{2}_{Lip^{p}(\mathbb {R}^{n})}|P|^{1-\tau -\frac{2}{p}}\sum _{j\in \mathbb {N}}2^{-2jn\tau }2^{-2jn(\frac{1}{p}-1)} \lesssim \Vert f\Vert ^{2}_{Lip^{p}(\mathbb {R}^{n})}. \end{aligned}$$

Thus, we obtain (2.1).

Conversely, first we claim that

$$\begin{aligned} |\psi _{j'}*f(x)|\lesssim & {} \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}\Vert \psi _{j'}(x-\cdot )\Vert _{h^{p}(\mathbb {R}^{n})}. \end{aligned}$$
(2.2)

Then, if we obtain that

$$\begin{aligned} \Vert \psi _{j'}(x-\cdot )\Vert _{h^{p}(\mathbb {R}^{n})}\lesssim 2^{j'n(1-\frac{1}{p})}, \end{aligned}$$
(2.3)

one can complete the proof. Now we prove (2.2). For any fixed \(x\in \mathbb {R}^{n}\), set \(h(t)=\psi _{j'}(x-t)\). Define

$$\begin{aligned} S(h)(\cdot )=\big (\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j}}|(\psi _{j}*h)(x_{Q})|^{2}\chi _{Q}(\cdot )\big )^{1/2}. \end{aligned}$$

For any \(i\in \mathbb {Z}\), let

$$\begin{aligned} \Omega _i =\left\{ x \in \mathbb R^{n}: S(h)(x)> 2^i\right\} ,\tilde{\Omega }_i=\left\{ x \in \mathbb R^{n}: M(\chi _{\Omega _i}) (x)> \frac{1}{(10)^{n}}\right\} \end{aligned}$$

and

$$\begin{aligned} \mathcal {B}_i=\left\{ Q: Q\in \cup _{j\ge 0}\Pi _{j}, \ |Q \cap \Omega _i| > \frac{1}{2} |Q|, \ \ |Q \cap \Omega _{i+1}| \le \frac{1}{2}|Q|\right\} . \end{aligned}$$

Then, \(\cup _{Q\in \mathcal {B}_i}Q\subseteq \tilde{\Omega }_i\). By local reproducing formula (1.8), one has

$$\begin{aligned} \psi _{j'}*f(x)= & {} \sum _{j\in \mathbb {N}}\sum _{Q\in \Pi _{j}}\psi _{j}*f(x_{Q})\cdot \psi _{j}*\psi _{j'}(x-x_{Q})\\= & {} \sum _{Q\in \Pi }\psi _{Q}*f(x_{Q}) \psi _{Q}*\psi _{j'}(x-x_{Q})\\= & {} \sum ^{+\infty }_{i=-\infty }\sum _{Q^{*}\in B_{i}}\sum _{Q\subseteq Q^{*},Q\in B_{i}}|Q||\psi _{Q}*f(x_{Q})||\psi _{Q}*\psi _{j'}(x-x_{Q})|. \end{aligned}$$

in \(L^{2}(\mathbb {R}^{n})\), where \(Q^{*}\) are the maximal dyadic cubes in \(\mathcal {B}_i\), and \(\varphi _{j}\) are denoted by \(\varphi _{Q}\).

Hence,

$$\begin{aligned} |\psi _{j'}*f(x)|\le & {} \big \{\sum ^{+\infty }_{i=-\infty }\sum _{Q^{*}\in B_{i}}|Q^{*}|^{1-\frac{p}{2}}\big (\sum _{Q\subseteq Q^{*},Q\in B_{i}} |Q|(\psi _{Q}*\psi _{j'}(x-x_{Q}))^{2}\big )^{\frac{p}{2}}\\&\cdot \big (\frac{1}{|Q^{*}|^{\frac{2}{p}-1}} \sum _{Q\subseteq Q^{*},Q\in B_{i}}|Q||\psi _{Q}*f(x_{Q})|^{2}\big )^{\frac{p}{2}}\big \}^{\frac{1}{p}}\\\le & {} \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}\big \{\sum ^{+\infty }_{i=-\infty }\sum _{Q^{*}\in B_{i}}|Q^{*}|^{1-\frac{p}{2}}\nonumber \\&\big (\sum _{Q\subseteq Q^{*},Q\in B_{i}}|Q||\psi _{Q}*\psi _{j'}(x-x_{Q})|^{2}\big )^{\frac{p}{2}}\big \}^{\frac{1}{p}}. \end{aligned}$$

Since \(Q\in \mathcal {B}_i\), one has \(|Q|\le 2|Q\cap \widetilde{\Omega }_{i}\backslash \Omega _{i+1}|\). Hence,

$$\begin{aligned} \sum _{Q\subseteq Q^{*},Q\in \mathcal {B}_i}|Q||(\varphi _{Q}*h)(x_{Q})|^{2}\le & {} 2\sum _{Q\in \mathcal {B}_i}|Q\cap \widetilde{\Omega }_{i}\backslash \Omega _{i+1}||(\varphi _{Q}*h)(x_{Q})|^{2}\\= & {} 2\int _{\widetilde{\Omega }_{i}\backslash \Omega _{i+1}}\sum _{Q\in \mathcal {B}_i}|(\varphi _{Q}*h)(x_{Q})|^{2}\chi _{Q}(x)dx\\\le & {} 2\int _{\widetilde{\Omega }_{i}\backslash \Omega _{i+1}}(S(h)(x))^{2}dx\lesssim 2^{2i}|\widetilde{\Omega }_{i}|, \end{aligned}$$

which implies that

$$\begin{aligned}&|\psi _{j'}*f(x)|\quad \lesssim \quad \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}\left\{ \sum \limits _{i=-\infty }^{+\infty }|\tilde{\Omega } _i|^{1-\frac{p}{2}}(2^{2i} |\tilde{\Omega }_i|)^{p/2}\right\} ^{\frac{1}{p}}\nonumber \\&\quad \lesssim \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})}\{\sum \limits _{i=-\infty }^{+\infty }2^{ip}|\Omega _{i}|\}^{\frac{1}{p}}\\&\quad \lesssim \quad \Vert f\Vert _{cmo^{p}(\mathbb {R}^{n})} \Vert h\Vert _{h^{p}(\mathbb {R}^{n})}. \end{aligned}$$

It is well known that

$$\begin{aligned} \Vert S(h)\Vert _{L^p(\mathbb {R}^{n})}\approx \Vert \big (\sum \limits _{j\in \mathbb {N}} |\psi _{j}*h(\cdot )|^{2}\big )^{1/2}\Vert _{L^{p}(\mathbb {R}^{n})}. \end{aligned}$$

Hence,

$$\begin{aligned} \Vert \psi _{j'}(x-\cdot )\Vert ^{p}_{h^{p}(\mathbb {R}^{n})}\approx & {} \int \big (\sum \limits _{j\in \mathbb {N}} |\psi _{j}*[\psi _{j'}(x-\cdot )](y)|^{2}\big )^{p/2}\mathrm{d}y\\= & {} \int \big (\sum \limits _{j\in \mathbb {N}} |\psi _{j}*\psi _{j'}(y)|^{2}\big )^{p/2}\mathrm{d}y\\\lesssim & {} \int \sum _{j\in \mathbb {N}}\frac{2^{(j\wedge j')np}2^{-|j-j'| Np}}{(1+2^{(j\wedge j')}|y|)^{Lp}}dy\\\lesssim & {} \sum _{j\in \mathbb {N}}2^{(j\wedge j')(np-n)}\cdot 2^{-|j-j'| Np}\\= & {} 2^{j'n(p-1)}\sum _{j\in \mathbb {N}}2^{(j\wedge j')n(p-1)}\cdot 2^{j'n(1-p)}\cdot 2^{-|j-j'| Np}\\= & {} 2^{j'n(p-1)}\sum _{j\in \mathbb {N}}2^{[(j'-j\wedge j')]n(1-p)}\cdot 2^{-|j-j'| Np}. \end{aligned}$$

Since\(|j'-j\wedge j'|\le | j-j'|\), one has

$$\begin{aligned} \Vert \psi _{j'}(x-\cdot )\Vert ^{p}_{h^{p}(\mathbb {R}^{n})}\lesssim & {} 2^{j'n(p-1)}\sum _{j\in \mathbb {N}}2^{|j-j'|n(1-p)}\cdot 2^{-|j-j'| Np}\\= & {} 2^{j'n(p-1)}\sum _{j\in \mathbb {N}}2^{-|j-j'|(Np-n(1-p))}\\= & {} 2^{j'n(p-1)}\sum _{j\in \mathbb {N}}2^{-|j-j'|p(N-n(\frac{1}{p}-1))}. \end{aligned}$$

Choosing N big enough such that \(N-n(\frac{1}{p}-1)>0\), one obtain (2.3).

Thus we complete the proof. \(\square \)

3 Equivalence Between \(Lip^{p}(\mathbb {R}^{n})\) and \(\Lambda _{r}\)

Theorem 3.1

\(f\in \Lambda _r, 0<r<1\) if and only if \(f\in \mathcal S'(\mathbb {R}^{n})\) and \(\sup _{j\in \mathbb N, x\in \mathbb R^n}2^{jr}|\psi _j*f(x)|<+\infty .\) Moreover, \(\Vert f\Vert _{\Lambda _r}\approx \sup _{j\in \mathbb N, x\in \mathbb R^n}2^{jr}|\psi _j*f(x)|.\)

Proof

Suppose that \(f\in \Lambda _r, 0<r<1,\) that is, f is a bounded and continuous function. Then it is easy to check that \(f\in \mathcal S'(\mathbb {R}^{n}), |\psi _0*f(x)|\le C \Vert f\Vert _{L^{\infty }(\mathbb {R}^{n})}\), and for \(j\ge 1,|\psi _j*f(x)|=|\int \psi _j(x-y)f(y)\mathrm{d}y|=|\int \psi _j(x-y)[f(y)-f(x)]\mathrm{d}y|\) since \(\int \psi _j(x)\mathrm{d}x=0.\) Thus, \(|\psi _j*f(x)|\le C 2^{-rj}\Vert f\Vert _{\Lambda _r}.\) These estimates imply that \(\sup _{j\in \mathbb N, x\in \mathbb R^n}2^{jr}|\psi _j*f(x)|\le C\Vert f\Vert _{\Lambda _r}.\) Conversely, if \(f\in \mathcal S'(\mathbb R^{n}),\) by the continuous Calderón reproducing formula (1.7), \(f(x)=\sum \limits _{j=0}^{\infty } \psi _{j}*\psi _j*f(x)\) in \(\mathcal S'(\mathbb R^{n}).\) Note that if \(\sup _{j\in \mathbb N, x\in \mathbb R^n}2^{jr}|\psi _j*f(x)|\le C,\) then \(|\psi _{j}*\psi _j*f(x)|\le C2^{-rj}\) and hence, the series \(\sum \limits _{j=0}^{\infty } \psi _{j}*\psi _j*f(x)\) converges uniformly. This implies that \(\sum \limits _{j=0}^{\infty } \psi _{j}*\psi _j*f(x)\) is a bounded and continuous function, thus \(f(x)=\sum \limits _{j=0}^{\infty } \psi _{j}*\psi _j*f(x)\) for all \(x\in \mathbb R^n.\) To see that \(f\in \Lambda _r, 0<r<1,\) we only need to show that for any given x and y with \(|x-y|\le 1, |f(x)-f(y)|\le C|x-y|^r\sup _{j\in \mathbb N, u\in \mathbb R^n}2^{jr}|\psi _j*f(u)|.\) To do this, let \(j_0\in \mathbb N\) such that \(2^{-j_0}\le |x-y|<2^{1-j_0}.\) Split the series into two parts: \(I(x)=\sum \limits _{j=0}^{j_0}\psi _{j}*\psi _j*f(x)\) and \(II(x)=\sum \limits _{j=j_0+1}^{\infty }\psi _{j}*\psi _j*f(x).\) Write \(I(x)-I(y)=\sum \limits _{j=0}^{j_0}\int [\psi _{j}(x-z)-\psi _j(y-z)]\psi _j*f(z)\mathrm{d}z.\) Applying the estimate that \(\int |[\psi _{j}(x-z)-\psi _j(y-z)]\psi _j*f(z)\mathrm{d}z|\le C(2^j|x-y|)2^{-rj}\sup _{j\in \mathbb N, u\in \mathbb R^n}2^{jr}|\psi _j*f(u)|,\) we obtain that

$$\begin{aligned} |I(x)-I(y)|\le & {} C2^{j_0(1-r)}|x-y|\sup _{j\in \mathbb N, u\in \mathbb R^n}2^{jr}|\psi _j*f(u)|\\\le & {} C |x-y|^r\sup _{j\in \mathbb N, u\in \mathbb R^n}2^{jr}|\psi _j*f(u)| . \end{aligned}$$

Applying the size condition, it follows that

$$\begin{aligned} |II(x)-II(y)|\le & {} C2^{-rj_0}\sup _{j\in \mathbb N, u\in \mathbb R^n}2^{jr}|\psi _j*f(u)\\\le & {} C|x-y|^r \sup _{j\in \mathbb N, u\in \mathbb R^n}2^{jr}|\psi _j*f(u)|. \end{aligned}$$

Thus, we complete the proof. \(\square \)

Theorem 3.2

\(f\in bmo(\mathbb R^n)\) if and only if \(f\in cmo(\mathbb R^n).\) Moreover, \(\Vert f\Vert _{bmo(\mathbb R^n)}\approx \Vert f\Vert _{cmo(\mathbb R^n)}.\)

Note that, for convenience, here we denote \(cmo(\mathbb {R}^{n})=cmo^{1}(\mathbb {R}^{n})\) .

Proof

Given any \(f\in cmo(\mathbb {R}^{n})\), we now prove

$$\begin{aligned} \Vert f\Vert _{bmo(\mathbb {R}^{n})}\lesssim \Vert f\Vert _{cmo(\mathbb {R}^{n})}. \end{aligned}$$

To do this, define a linear functional on \(h^{1}(\mathbb {R}^{n})\) by

$$\begin{aligned} \mathcal {L}_{f}(g)=\langle f, g\rangle \end{aligned}$$

for \(g \in h^{1}(\mathbb {R}^{n}).\) By the above the duality argument, namely \((h^{1}(\mathbb {R}^{n}))^{*}=cmo(\mathbb {R}^{n}),\) we have

$$\begin{aligned} |\mathcal {L}_{f}(g)|\le \Vert f\Vert _{cmo(\mathbb {R}^{n})}\Vert g\Vert _{h^{1}(\mathbb {R}^{n})}. \end{aligned}$$

Now fix a \(Q\in \bar{\Pi }\), and let \(L^{2}_{Q}\) denote the space of all square integrable functions supported in Q. If \(\ell (Q)\ge 1\), it is easy to see that each \(g\in L^{2}_{Q}\) is a multiple of an (1, 2) atom of \(h^{1}(\mathbb {R}^{n})\) with \(\Vert g\Vert _{h^{1}(\mathbb {R}^{n})}\le C|Q|^{\frac{1}{2}}\Vert g\Vert _{L^{2}(\mathbb {R}^{n})}\). Hence, \(\mathcal {L}_{f}\) is a linear functional on \(L^{2}_{Q}\) with norm at most \(C|Q|^{\frac{1}{2}}\Vert f\Vert _{cmo(\mathbb {R}^{n})}\). Then, the Riesz representation theorem for Hilbert spaces \(L^{2}_{Q}\) tells us that there exists \(F^{Q}\in L^{2}_{Q}\) such that

$$\begin{aligned} \mathcal {L}_{f}(g)=\langle f, g\rangle =\int _{Q}F^{Q}(x)g(x)\mathrm{d}x, \forall g\in L^{2}_{Q} \end{aligned}$$

with \(\Vert F^{Q}\Vert _{L^{2}_{Q}}=\Vert \mathcal {L}_{f}\Vert \le C|Q|^{\frac{1}{2}}\Vert f\Vert _{cmo(\mathbb {R}^{n})},\) which yields that f must be a square integrable function on Q and \(f=F^{Q}\) on Q. Therefore, for any cubes Q satisfying \(\ell (Q)\ge 1\),

$$\begin{aligned} \left\{ \frac{1}{|Q|}\int _{Q}|f(x)|^{2}\mathrm{d}x\right\} ^{1/2} =\left\{ \frac{1}{|Q|}\int _{Q}|F^{Q}(x)|^{2}\mathrm{d}x\right\} ^{1/2}\le C\Vert f\Vert _{cmo(\mathbb {R}^{n})}, \end{aligned}$$

which implies that

$$\begin{aligned} \sup _{|Q|\ge 1}\frac{1}{|Q|}\int _{Q}|f(x)|\mathrm{d}x\le C\Vert f\Vert _{cmo(\mathbb {R}^{n})}. \end{aligned}$$

On the other hand, if \(\ell (Q)< 1\), let \(L^{2}_{Q,0}=\{f\in L_Q^2, \int f=0\}\). Recall that a function a(x) supported in a cube Q is said to be a (p, 2) atom of \(h^{p}(\mathbb {R}^{n}), 0<p\le 1,\) if it satisfies a size condition \(\Vert a\Vert _{L^{2}(\mathbb {R}^{n})}\le |Q|^{\frac{1}{2}-\frac{1}{p}}\) and a cancellation condition \(\int x^{\alpha }a(x)\mathrm{d}x=0, \alpha \in \mathbb {N}^{n}, |\alpha |\le n(\frac{1}{p}-1)\), when \(\ell (Q)<1\) [14]. Then, each \(g\in L^{2}_{Q,0}\) is a multiple of an (1, 2) atom of \(h^{1}(\mathbb {R}^{n})\) with \(\Vert g\Vert _{h^{1}(\mathbb {R}^{n})}\le C|Q|^{\frac{1}{2}}\Vert g\Vert _{L^{2}(\mathbb {R}^{n})}\), and \(\mathcal {L}_{f}\) is a linear functional on \(L^{2}_{Q,0}\) with norm at most \(C|Q|^{\frac{1}{2}}\Vert f\Vert _{cmo(\mathbb {R}^{n})}\). By the Riesz representation theorem, there exists some \(F^{Q}\in L^{2}_{Q,0}\) such that

$$\begin{aligned} \mathcal {L}_{f}(g)=\langle f, g\rangle =\int _{Q}F^{Q}(x)g(x)\mathrm{d}x, \forall g\in L^{2}_{Q,0} \end{aligned}$$

with \(\Vert F^{Q}\Vert _{L^{2}_{Q}}=\Vert \mathcal {L}_{f}\Vert \le C|Q|^{\frac{1}{2}}\Vert f\Vert _{cmo(\mathbb {R}^{n})},\) which yields that f must be a square integrable function on Q and \(f=F^{Q}+c_{Q}\) on Q for some constant \(c_{Q}\) since \(\int g(x)\mathrm{d}x=0\). Therefore, for any cube Q satisfying \(\ell (Q)< 1\),

$$\begin{aligned} \left\{ \frac{1}{|Q|}\int _{Q}|f(x)-c_{Q}|^{2}\mathrm{d}x\right\} ^{1/2} =\left\{ \frac{1}{|Q|}\int _{Q}|F^{Q}(x)|^{2}\mathrm{d}x\right\} ^{1/2}\le C\Vert f\Vert _{cmo(\mathbb {R}^{n})}, \end{aligned}$$

which implies that

$$\begin{aligned} \sup _{|Q|<1}\big \{\frac{1}{|Q|}\int _{Q}|f(x)-c_{Q}|^{2}\mathrm{d}x\big \}^{1/2}\le C\Vert f\Vert _{cmo(\mathbb {R}^{n})}. \end{aligned}$$

Hence

$$\begin{aligned} \sup _{|Q|<1}\big \{\frac{1}{|Q|}\int _{Q}|f(x)-f_{Q}|^{2}\mathrm{d}x\big \}^{1/2}\le C\Vert f\Vert _{cmo(\mathbb {R}^{n})}. \end{aligned}$$

The estimates above imply that \(\Vert f\Vert _{bmo(\mathbb {R}^{n})}\lesssim \Vert f\Vert _{cmo(\mathbb {R}^{n})}.\)

We now prove

$$\begin{aligned} \Vert f\Vert _{cmo(\mathbb {R}^{n})}\lesssim \Vert f\Vert _{bmo(\mathbb {R}^{n})}. \end{aligned}$$

By the definition of \(bmo(\mathbb {R}^{n})\), it is easy to see that,

$$\begin{aligned} \sup _{Q}\frac{1}{|Q|}\int _{Q}|f(x)-f_{Q}|\mathrm{d}x\le 2\Vert f\Vert _{bmo(\mathbb {R}^{n})}, \end{aligned}$$

where \(\sup \) is taken over all cubes with sides parallel to the coordinate axes in \(\mathbb {R}^{n}\).

For any \(P\in \bar{\Pi }\), let \(P^{*}=3\sqrt{n}P\), the cube with the same center of P and side length \(3\sqrt{n}\ell (P)\). Split f as \( f=f_{1}+f_{2}+f_{3}\) with \(f_{1}=(f-f_{P^{*}})\chi _{P^{*}}\), \(f_{2}=(f-f_{P^{*}})\chi _{(P^{*})^{c}}\) and \(f_{3}=f_{P^{*}}\). It is easy to see that

$$\begin{aligned}&\frac{1}{|P|} \int _{P}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}(|\psi _{j}*f(x_{Q})|^2 \chi _{Q}(x))\\&\quad \lesssim \sum _{i=1}^{3}\int _{P}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}(|\psi _{j}*f_{i}(x_{Q})|^2 \chi _{Q}(x)). \end{aligned}$$

For \(f_{1}\), by Littlewood–Paley theory,

$$\begin{aligned}&\frac{1}{|P|} \int _{P}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}(|\psi _{j}*f_{1}(x_{Q})|^2 \chi _{Q}(x)) \mathrm{d}x\\&\quad =\frac{1}{|P|}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f_{1}(x_{Q})|^2\\&\quad \le \frac{1}{|P|}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j}}|Q||\psi _{j}*f_{1}(x_{Q})|^2\\&\quad \lesssim \frac{1}{|P|}\Vert f_{1}\Vert _{L^{2}(\mathbb {R}^{n})}\lesssim \frac{1}{|P|}\int _{P^{*}}|f(y)-f_{P^{*}}|^{2}\mathrm{d}y\lesssim \Vert f\Vert ^{2}_{bmo(\mathbb {R}^{n})}. \end{aligned}$$

For \(f_{2}\), first of all, it is easy to see that \(|\psi _{j}*f_{2}(x_{Q})| =|\int \psi _{j}(x_{Q}-y)f_{2}(y)dy|\) is dominated by

$$\begin{aligned} \int _{(P^{*})^{c}}\frac{2^{-j}}{(2^{-j}+|x_{Q}-y|)^{n+1}}|f_{2}(y)|\mathrm{d}y \approx \int _{(P^{*})^{c}}\frac{2^{-j}}{(2^{-j}+|x_{P}-y|)^{n+1}}|f_{2}(y)|\mathrm{d}y, \end{aligned}$$

where \(x_{P}\) is the center of P. Hence

$$\begin{aligned}&|\psi _{j}*f_{2}(x_{Q})|\\&\quad \lesssim 2^{-j}\sum _{k=1}^{+\infty }\int _{2^{k-1}\ell (P)\le |x_{P}-y|<2^{k}\ell (P)} \frac{1}{(2^{-j}+|x_{P}-y|)^{n+1}}|f_{2}(y)|\mathrm{d}y\\&\quad \lesssim 2^{-j}\sum _{k=1}^{+\infty }\frac{1}{(2^{k}\ell (P))^{n+1}}\int _{|x_{P}-y|<2^{k}\ell (P)}|f_{2}(y)|\mathrm{d}y\\&\quad \lesssim 2^{-j}\sum _{k=1}^{+\infty }\frac{1}{(2^{k}\ell (P))^{n+1}}\left\{ \int _{|x_{P}-y|<2^{k}\ell (P)}|f(y)\right. \\&\left. \qquad -f_{2^{k}P^{*}}|\mathrm{d}y+(2^{k}\ell (P))^{n} |f_{2^{k}P^{*}}- f_{P^{*}}| \right\} , \end{aligned}$$

combining with classical result about \(BMO(\mathbb {R}^{n})\) functions, gives

$$\begin{aligned} |\psi _{j}*f_{2}(x_{Q})|\lesssim 2^{-j}\ell (P)^{-1}\Vert f\Vert _{bmo(\mathbb {R}^{n})}. \end{aligned}$$

Note that there exists \(j_{0}\in \mathbb {Z}\) such that \(2^{-j_{0}}<\ell (P)\le 2^{-j_{0}+1}\). Then,

$$\begin{aligned}&\frac{1}{|P|}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f_{2}(x_{Q})|^2\\&\qquad \lesssim \frac{1}{|P|}\sum \limits _{j=j_{0}}^{\infty }\sum \limits _{Q\in \Pi _{j},Q\subseteq P}|Q|2^{-2(j-j_{0})}\Vert f\Vert ^{2}_{bmo(\mathbb {R}^{n})}\lesssim \Vert f\Vert ^{2}_{bmo(\mathbb {R}^{n})}. \end{aligned}$$

For the constant item \(f_{3}\), note that \( \int \psi _{0}=1\) and \( \int \psi _{j}=0, j\ge 1\). Hence if \(\ell (P)<1\),

$$\begin{aligned} \frac{1}{|P|}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f_{3}(x_{Q})|^2=\frac{1}{|P|}\sum \limits _{j\ge 1} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f_{3}(x_{Q})|^2=0. \end{aligned}$$

Otherwise, if \(\ell (P)\ge 1\)

$$\begin{aligned} \frac{1}{|P|}\sum \limits _{j\in \mathbb {N}} \sum \limits _{Q\in \Pi _{j},Q\subseteq P}|Q||\psi _{j}*f_{3}(x_{Q})|^2\le |f_{3}|^{2}\le (S_{2}(f))^{2}\le \Vert f\Vert ^{2}_{bmo(\mathbb {R}^{n})}, \end{aligned}$$

since \(f_{3}=f_{P^{*}}\) with \(\ell (P^{*})>1\).

Thus, we complete the proof. \(\square \)

We want to point out that the results of this section can also be seen in [8]. For completeness, we give their proofs.

4 Applications

In this section, we discuss the boundedness of inhomogeneous para-product operators on \(h^{p}(\mathbb {R}^{n})\). Firstly, we recall some about non-convolution singular integral operators.

A locally integral function \(\mathcal {K}(x,y)\) defined away from the diagonal \(x=y\) in \(\mathbb {R}^{n}\times \mathbb {R}^{n}\) is called a Calderón–Zygmund kernel with regularity exponent \(\varepsilon >0\) if there exists a constant \(C >0\) such that

$$\begin{aligned} |\mathcal {K}(x,y)|\le C\frac{1}{|x-y|^{n}},\quad \ \hbox {for}\quad \ x\ne y, \end{aligned}$$
(4.1)

and

$$\begin{aligned} |\mathcal {K}(x,y)-\mathcal {K}(x,y')|\le C \frac{|y-y'|^{\varepsilon }}{|x-y|^{n+\varepsilon }}, \end{aligned}$$
(4.2)

whenever \(|y-y'|\le \frac{1}{2}|x-y|\), and

$$\begin{aligned} |\mathcal {K}(x,y)-\mathcal {K}(x',y)|\le C\frac{|x-x'|^{\varepsilon }}{|x-y|^{n+\varepsilon }}, \end{aligned}$$
(4.3)

whenever \(|x-x'|\le \frac{1}{2}|x-y|\). The operator T is said to be non-convolution Calderón–Zygmund singular integral if T is a continuous linear operator from \( \mathcal {D}(\mathbb {R}^{n})\) to \( \mathcal {D}'(\mathbb {R}^{n})\) defined by

$$\begin{aligned} \langle Tf,g\rangle =\int \mathcal {K}(x,y)f(y)g(x)\mathrm{d}x\mathrm{d}y \end{aligned}$$

for all \(f,g\in \mathcal {D}(\mathbb {R}^{n})\) with disjoint supports, where \(\mathcal {K}\) is a Calderón–Zygmund kernel. The fundamental result for the third generation Calderón–Zygmund singular integrals is the Theorem T1 contained in [6]. See [7, 16, 17, 20,21,22, 30] for other versions of the T1 theorems for Hardy, Besov and Triebel–Lizorkin spaces. If the Calderón–Zygmund kernel \(\mathcal {K}\) satisfies a restrictve size condition, namely, the condition (4.1) is replaced by

$$\begin{aligned} |\mathcal {K}(x,y)|\le C\min \{\frac{1}{|x-y|^{n}},\frac{1}{|x-y|^{n+\delta }}\},\ \hbox {for some }\delta >0 \ \hbox {and}\ \ x\ne y, \end{aligned}$$
(4.4)

we then obtain an inhomogeneous Calderón–Zygmund kernel associate with regularity exponent \(\varepsilon , \delta > 0\), and inhomogeneous Calderón–Zygmund singular integral associate with regularity exponent \(\varepsilon , \delta > 0\), respectively. It is well known that each pseudo-differential operator \(T_{\sigma }f(x)=\int \sigma (x,\xi )e^{2\pi ix\xi }\hat{f}(\xi )d\xi \) with \(\sigma \in S_{1,0}^{0}\) is an inhomogeneous Calderón–Zygmund singular integral. For the boundedness of operators on local Hardy spaces, we refer the readers to the work in [14, 26]

Using atomic decomposition, one has the following result.

Theorem 4.1

Suppose that T is an inhomogeneous Calderón–Zygmund singular integral associate with regularity exponent \(\varepsilon , \delta > 0\). Then, if T is bounded on \(L^{2}(\mathbb {R}^{n})\), T is bounded from \(h^{p}(\mathbb {R}^{n})\) to \(L^{p}(\mathbb {R}^{n})\) if \(\max \{\frac{n}{n+\varepsilon },\frac{n}{n+\delta }\}<p\le 1\).

The proof of Theorem 4.1 is standard, and we refer the readers to [15] for non-convolution Calderón–Zygmund singular integral on \(H^{p}(\mathbb {R}^{n})\), to [12] for Journé’s type of multi-parameter singular integral operators on multi-parameter Hardy spaces \(H^{p}(\mathbb {R}^{n_{1}} \times \mathbb {R}^{n_{2}})\), to [10] for inhomogeneous Journé’s type of multi-parameter singular integral operators on multi-parameter Hardy spaces \(h^{p}(\mathbb {R}^{n_{1}} \times \mathbb {R}^{n_{2}})\), so we omit its proof. We want to remark that it is enough to prove Theorem 4.1 under conditions (4.4) and (4.2).

Now we begin to define inhomogeneous para-product operators. In this section, we always suppose that \(\psi _{0},\psi , \varphi \) belong to \(\mathcal {D}(\mathbb {R}^{n})\) supported on unit ball centered at origin satisfying that \(\psi \) has vanishing moments up to some proper order N and \(\int \psi _{0}(x)\mathrm{d}x=\int \varphi (x)\mathrm{d}x=1\). Set \(\psi _{j}(x)=2^{jn} \psi (2^jx)\) for \(j\in \mathbb {N}, j\ge 1\), and \(\varphi _{j}(x)=2^{jn} \varphi (2^jx)\) for \(j\in \mathbb {N}\).

By Corollary 1.3, \(bmo(\mathbb {R}^{n})\supseteq cmo^{p}(\mathbb {R}^{n})\) for all \(0<p\le 1\). Fixed any \(b\in bmo(\mathbb {R}^{n})\), inhomogeneous para-product operators are defined as following:

$$\begin{aligned} \pi _{b}(f)(x)=\sum _{j\in N}\sum _{Q\in \Pi _{j}}|Q|\psi _{j}*b(x_{Q})\cdot \psi _{j}(x-x_{Q})\varphi _{j}*f (x_{Q}), \end{aligned}$$

and its adjoint operator

$$\begin{aligned} \pi _{b}^{*}(f)(x)= \sum \limits _{k\in N} \sum \limits _{Q\in \Pi _{k}}|Q|\varphi _{k}(x-x_{Q})\psi _{k}*b(x_{Q})\cdot \psi _{k}*f(x_{Q}). \end{aligned}$$

One can check that \(\pi _{b},\pi _{b}^{*}\) are inhomogeneous Calderón–Zygmund singular integrals associate with regularity exponent \(\varepsilon =\delta =1\) and bounded on \(L^{2}(\mathbb {R}^{n})\). Hence \(\pi _{b},\pi _{b}^{*}\) are bounded from \(h^{p}(\mathbb {R}^{n})\) to \(L^{p}(\mathbb {R}^{n})\) if \(\frac{n}{n+1}<p\le 1\). Now we give the main result of this section.

Theorem 4.2

Suppose that \(b\in Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})\). Then, \(\pi _{b},\pi _{b}^{*}\) are bounded on \(h^{p}(\mathbb {R}^{n})\) if \(\frac{n}{n+1}<p\le 1\).

To prove Theorem 4.2, we will follow the approach in [23, 24] by reducing the \(h^{p}(\mathbb {R}^{n})\) boundedness of \(\pi _{b}\) and \(\pi _{b}^{*}\) to \(h^{p}(\mathbb {R}^{n})\rightarrow L^{p}(\mathbb {R}^{n})\) boundedness.

Since the proofs of the \(h^{p}(\mathbb {R}^{n})\) boundedness of \(\pi _{b}^{*}\) is similar as \(\pi _{b}'s\) but more difficult, we only give the proof of \(\pi _{b}^{*}\). Obviously, \(L^{2}(\mathbb {R}^{n})\cap h^{p}(\mathbb {R}^{n})\) is dense in \(h^{p}(\mathbb {R}^{n})\) since \(\mathcal {S}(\mathbb {R}^{n})\) is dense in \(h^{p}(\mathbb {R}^{n})\). For \(f\in L^{2}(\mathbb {R}^{n})\cap h^{p}(\mathbb {R}^{n})\), using (1.3), one has

$$\begin{aligned} \Vert \pi _{b}^{*}(f)\Vert _{h^{p}(\mathbb R^{n})}=\left\| \left\{ \sum \limits _{k\in \mathbb {N }}|\psi _{k}*\pi _{b}^{*}(f)(x)|^{2}\right\} ^{\frac{1}{2}}\right\| _{L^{p}(\mathbb R^{n})}. \end{aligned}$$

It is easy to see that \(\{\sum \limits _{k\in \mathbb {N }}|\psi _{k}*\pi _{b}^{*}(f)(x)|^{2}\}^{\frac{1}{2}}\) is bounded on \(L^{2}(\mathbb {R}^{n})\), since \(b\in Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})\subseteq bmo(\mathbb {R}^{n})\). Moreover,

$$\begin{aligned}&\psi _{k}*\pi _{b}^{*}(f)(x)=\int \psi _{k}(x-z)\pi _{b}^{*}(f)(z)\mathrm{d}z\\&\quad =\int \psi _{k}(x-z)\sum \limits _{k'\in \mathbb {N}} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\varphi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'})\psi _{k'}*f(x_{Q'})\mathrm{d}z\\&\quad =\iint \sum \limits _{k'\in \mathbb {N}} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\psi _{k}(x-z)\varphi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)f(y)\mathrm{d}z\mathrm{d}y. \end{aligned}$$

Set

$$\begin{aligned} S_{k}(x,y)=\int \sum \limits _{k'\in \mathbb {N}} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\psi _{k}(x-z)\varphi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)\mathrm{d}z. \end{aligned}$$

Then, the \(h^{p}(\mathbb {R}^{n})\) boundedness of \(\pi _{b}^{*}\) can be reduced to \(h^{p}(\mathbb {R}^{n})\rightarrow L^{p}(\mathbb {R}^{n})\) boundedness of vector value inhomogeneous Calderón–Zygmund singular integral with kernel \(\{S_{k}(x,y)\}\). We will finish the proof of Theorem 4.2 by proving that \(\{S_{k}(x,y)\}\) satisfies vector value inhomogeneous Calderón–Zygmund kernel conditions, namely,

Lemma 4.3

Let \(S_{k}(x,y)=\int \sum \limits _{k'\in \mathbb {N}} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\psi _{k}(x-z)\varphi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)dz,\) and suppose that \(b\in Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})\). Then for every \(\varepsilon \in (0,1)\),

  1. (i)

    \(\{\sum \limits _{k\in \mathbb {N}}|S_{k}(x,y)|^{2}\}^{1/2}\le C\min \{\frac{1}{|x-y|^{n}},\frac{1}{|x-y|^{n+1}}\}\) if \(|x-y|>0;\)

  2. (ii)

    \(\{\sum \limits _{k\in \mathbb {N}}|S_{k}(x,y')-S_{k}(x,y)|^{2}\}^{1/2}\le C \frac{|y-y'|^{\varepsilon }}{|x-y|^{n+\varepsilon }}\), if \(|y-y'|\le \frac{1}{2}|x-y|.\)

Proof

Split \(S_{k}(x,y)\) as following

$$\begin{aligned}S_{k}(x,y)= & {} \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\int \psi _{k}(x-z)\varphi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)\mathrm{d}z\\&+ \sum \limits _{k'<k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\int \psi _{k}(x-z)\varphi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)\mathrm{d}z\\&=S^{1}_{k}(x,y)+S^{2}_{k}(x,y). \end{aligned}$$

We first prove size condition (i). For \(S^{1}_{k}\), one has

$$\begin{aligned} S^{1}_{k}= & {} \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\int [\psi _{k}(x-z-x_{Q'})\\&-\psi _{k}(x-x_{Q'})]\varphi _{k'}(z)\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)\mathrm{d}z\\&+\sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\psi _{k}(x-x_{Q'})\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)=A+B. \end{aligned}$$

We first show that

$$\begin{aligned} |B|\lesssim \frac{2^{-k}}{(2^{-k}+|x-y|)^{n+1}}. \end{aligned}$$
(4.5)

If \(|x-y|\le 2\cdot 2^{-k}\),

$$\begin{aligned} |B|\le & {} \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'||\psi _{k}(x-x_{Q'})||\psi _{k'}*b(x_{Q'})||\psi _{k'}(x_{Q'}-y)|\nonumber \\\le & {} \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\frac{2^{-k\varepsilon '}}{(2^{-k}+|x-x_{Q'}|)^{n+\varepsilon '}}|\psi _{k'}*b(x_{Q'})||\psi _{k'}(x_{Q'}-y)|\nonumber \\\le & {} \frac{2^{-k}}{2^{-k(n+1)}}\sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'||\psi _{k'}*b(x_{Q'})||\psi _{k'}(x_{Q'}-y)|\nonumber \\\lesssim & {} \frac{2^{-k}}{(2^{-k}+|x-y|)^{n+1}}\sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'||\psi _{k'}*b(x_{Q'})||\psi _{k'}(x_{Q'}-y)|. \end{aligned}$$
(4.6)

When \(|x-y|> 2\cdot 2^{-k}\). Note that \(|x_{Q'}-y|\le 2^{-k'}\le 2^{-k}\). Hence, \(|x-x_{Q'}|\ge |x-y|-|y-x_{Q'}|\ge \frac{1}{2}|x-y|,\) which also yields (4.6). Thus, (4.5) is obtained if we can prove

$$\begin{aligned} \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'||\psi _{k'}*b(x_{Q'})||\psi _{k'}(x_{Q'}-y)|\le C. \end{aligned}$$

Indeed, using \(b\in Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})\), one has

$$\begin{aligned}&\sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'||\psi _{k'}*b(x_{Q'})||\psi _{k'}(x_{Q'}-y)|\\&\qquad \le \Vert b\Vert _{Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})}\sum \limits _{k'\ge k}\sum \limits _{Q'\in \Pi _{k'}} |Q'|2^{k'(1-\frac{n+1}{n})}|\psi _{k'}(x_{Q'}-y)|\\&\qquad \lesssim \sum \limits _{k'\ge 0}2^{k'(1-\frac{n+1}{n})}\sum \limits _{Q'\in \Pi _{k'}}\int _{Q'}\frac{2^{-k'}}{(2^{-k'}+|x_{Q'}-y|)^{n_1}}dz\\&\qquad \lesssim \int _{Q'}\frac{2^{-k'}}{(2^{-k'}+|z-y|)^{n_1}}dz\le C, \end{aligned}$$

since \(2^{-k'}+|x_{Q'}-y|\approx 2^{-k'}+|z-y|\) for any \(z\in Q'\).

For A, firstly, by the fact that \(\psi _{k'}*b(x_{Q'})\) is bounded uniformly, one has

$$\begin{aligned} |A|\lesssim & {} \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\int |\psi _{k}(x-z-x_{Q'})-\psi _{k}(x-x_{Q'})||\varphi _{k'}(z)||\psi _{k'}(x_{Q'}-y)|\mathrm{d}z. \end{aligned}$$

Hence, by the following classical result, for any \(\varepsilon \in (0,1)\),

$$\begin{aligned}&|\psi _{k}(x-z-x_{Q'})-\psi _{k}(x-x_{Q'})|\\&\qquad \lesssim \left( \frac{|z|}{2^{-k}}\right) ^{\varepsilon }\left( \frac{2^{-k}}{(2^{-k}+|x-z-x_{Q'}|)^{n+1}}+\frac{2^{-k}}{(2^{-k}+|x-x_{Q'}|)^{n+1}}\right) , \end{aligned}$$

and using the fact that the support of \(\psi \) is unit ball at origin, A is dominated by

$$\begin{aligned}&\sum \limits _{k'\ge k}\sum \limits _{Q'\in \Pi _{k'}}|Q'|\\&\qquad \int \left( \frac{2^{-k'}}{2^{-k}}\right) ^{\varepsilon } \left( \frac{2^{-k}}{(2^{-k}+|x-z-x_{Q'}|)^{n+1}}+\frac{2^{-k}}{(2^{-k}+|x-x_{Q'}|)^{n+1}}\right) \\&\qquad \frac{2^{-k'}}{(2^{-k'}+|z|)^{n+1}}|\psi _{k'}(x_{Q'}-y)|\mathrm{d}z\\&\quad =\sum \limits _{k'\ge k}\sum \limits _{Q'\in \Pi _{k'}}|Q'|2^{-(k'-k)\varepsilon }|\psi _{k'}(x_{Q'}-y)|\\&\qquad \int \frac{2^{-k}}{(2^{-k}+|x-z-x_{Q'}|)^{n+1}} \frac{2^{-k'}}{(2^{-k'}+|z|)^{n+1}}\mathrm{d}z\\&\qquad +\sum \limits _{k'\ge k}\sum \limits _{Q'\in \Pi _{k'}} |Q'|2^{-(k'-k)\varepsilon }\frac{2^{-k}}{(2^{-k}+|x-x_{Q'}|)^{n+1}}| \psi _{k'}(x_{Q'}-y)|\\&\qquad \int \frac{2^{-k'}}{(2^{-k'}+|z|)^{n+1}}\mathrm{d}z\\&\quad \lesssim \sum \limits _{k'\ge k}\sum \limits _{Q'\in \Pi _{k'}}|Q'|2^{-(k'-k)\varepsilon }|\psi _{k'}(x_{Q'}-y)| \frac{2^{-k}}{(2^{-k}+|x-x_{Q'}|)^{n+1}}\\&\quad \lesssim \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|2^{-(k'-k)\varepsilon } \frac{2^{-k'}}{(2^{-k'}+|y-x_{Q'}|)^{n+1}}\frac{2^{-k}}{(2^{-k}+|x-x_{Q'}|)^{n+1}}\\&\quad =\sum \limits _{k'\ge k}\sum \limits _{Q'\in \Pi _{k'}} 2^{-(k'-k)\varepsilon '}\int _{Q'}\frac{2^{-k'(1-\varepsilon ')}}{(2^{-k'}+|y-x_{Q'}|)^{n+1-\varepsilon '}}\frac{2^{-k(1-\varepsilon ')}}{(2^{-k}+|x-x_{Q'}|)^{n+1-\varepsilon '}}\mathrm{d}z\\&\thickapprox \sum \limits _{k'\ge k} \sum \limits _{Q'\in \Pi _{k'}}2^{-(k'-k)\varepsilon '}\int _{Q'}\frac{2^{-k'(1-\varepsilon ')}}{(2^{-k'}+|y-z|)^{n+1-\varepsilon '}}\frac{2^{-k(1-\varepsilon ')}}{(2^{-k}+|x-z|)^{n+1-\varepsilon '}}\mathrm{d}z, \end{aligned}$$

since \(2^{-k'}+|y-x_{Q'}|\thickapprox 2^{-k'}+|y-z|,2^{-k}+|x-x_{Q'}|\thickapprox 2^{-k}+|x-z|\) if \(z\in Q'\). Then

$$\begin{aligned} |A|\lesssim & {} \sum \limits _{k'\ge k} 2^{-(k'-k)\varepsilon '}\frac{2^{-k}}{(2^{-k}+|x-y|)^{n+1}}\lesssim \frac{2^{-k}}{(2^{-k}+|x-y|)^{n+1}}. \end{aligned}$$
(4.7)

By estimates (4.5) and (4.7), one has

$$\begin{aligned} |S^{1}_{k}(x,y)| \lesssim \frac{2^{-k}}{(2^{-k}+|x-y|)^{n+1}}. \end{aligned}$$
(4.8)

For \(S^{2}_{k}\), using the cancelation condition of \(\psi \), one has

$$\begin{aligned}&S^{2}_{k}(x,y)\\&\quad =\sum \limits _{k'<k} \sum \limits _{Q'\in \Pi _{k'}}|Q'|\int \psi _{k}(x-z)[\varphi _{k'}(z-x_{Q'})\\&\qquad -\varphi _{k'}(x-x_{Q'})]\psi _{k'}*b(x_{Q'})\psi _{k'}(x_{Q'}-y)\mathrm{d}z. \end{aligned}$$

Then with a similar process to estimate A, one can see that (4.8) also holds for \(S^{2}_{k}\). Thus, we can complete the proof of size condition (i).

To prove (ii), we only estimate the smoothness of \(S^{1}_{k}(x,y)\) with the second variable y since the proof to obtain the smoothness of \(S^{2}_{k}(x,y)\) is similar and easier. Set

$$\begin{aligned}&S_{k}^{1}(x,y)-S_{k}^{1}(x,y')\\&\quad =\int \sum _{k'\ge k}\sum _{Q'\in \Pi _{k'}}|Q'|\psi _{k}(x-z)\psi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'}) [\psi _{k'}(x_{Q'}-y)]\\&\qquad -\psi _{k'}(x_{Q'}-y')]\mathrm{d}z\\&\quad =\int \sum _{k'\ge k}\sum _{Q'\in \Pi _{k'}}|Q'|[\psi _{k}(x-z)-\psi _{k}(x-x_{Q'})]\varphi _{k'}(z-x_{Q'})\psi _{k'}*b(x_{Q'})\\&\qquad \cdot [\psi _{k'}(x_{Q'}-y)-\psi _{k'}(x_{Q'}-y')]dz\\&\qquad +\sum _{k'\ge k}\sum _{Q'\in \Pi _{k'}}|Q'| \psi _{k}(x-x_{Q'})\psi _{k'}*b(x_{Q'})\left[ \psi _{k'}(x_{Q'}-y)-\psi _{k'}(x_{Q'}-y')\right] \\&\quad =A^{1}_{k}(x,y)+A^{2}_{k}(x,y). \end{aligned}$$

For any \(\varepsilon \in (0,1)\), using the support condition, it is easy to have

$$\begin{aligned}&|A^{1}_{k}(x,y)|\\&\quad \lesssim \int _{| z-x_{Q'}|\le 2^{-k'}}\sum _{k'\ge k}\sum _{Q'\in \Pi _{k'}}\mid Q'\mid \left( \frac{| z-x_{Q'}|}{2^{-k}}\right) ^{\varepsilon }\left[ \frac{2^{-k}}{(2^{-k}+\mid x-z\mid )^{n+1}}\right. \\&\left. \qquad +\frac{2^{-k}}{(2^{-k}+\mid x-x_{Q'}\mid )^{n+1}}\right] \frac{2^{-k'}}{(2^{-k'}+\mid z-x_{Q'}\mid )^{n+1}}\mid \left[ \psi _{k'}(x_{Q'}-y)\right. \\&\left. \qquad -\psi _{k'}(x_{Q'}-y')\right] \mid \mathrm{d}z, \end{aligned}$$

since \(\mid \psi _{k'}*b(x_{Q'})\mid \le C\), uniformly for \(k'\) and \(x_{Q'}\). Then,

$$\begin{aligned} |A^{1}_{k}(x,y)|\lesssim & {} \sum _{k'\ge k}2^{(k-k')\varepsilon }\sum _{Q'\in \Pi _{k'}}|Q'|\frac{2^{-k}}{(2^{-k}+| x-x_{Q'}|)^{n+1}}\left( \frac{| y-y'|}{2^{-k'}}\right) ^{\varepsilon '}\\&\cdot \left[ \frac{2^{-k'}}{(2^{-k'}+|x_{Q'}-y|)^{n+1}}+\frac{2^{-k'}}{(2^{-k'}+| x_{Q'}-y'|)^{n+1}}\right] \\\lesssim & {} \sum _{k'\ge k}2^{(k-k')\varepsilon }\int \frac{2^{-k}}{(2^{-k}+| x-u|)^{n+1}}\left( \frac{| y-y'|}{2^{-k'}}\right) ^{\varepsilon '}\\&\cdot \left[ \frac{2^{-k'}}{(2^{-k'}+| u-y|)^{n+1}}+\frac{2^{-k'}}{(2^{-k'}+| u-y'|)^{n+1}}\right] \mathrm{d}u\\\lesssim & {} \sum _{k'\ge k}2^{(k-k')\varepsilon }2^{(k'-k)\varepsilon '}(\frac{| y-y'|}{2^{-k}})^{\varepsilon '}\left[ \frac{2^{-k}}{(2^{-k}+| x-y|)^{n+1}}\right. \\&\left. +\frac{2^{-k}}{(2^{-k}+| x-y'|)^{n+1}}\right] \\\lesssim & {} \left( \frac{| y-y'|}{2^{-k}}\right) ^{\varepsilon '}\left[ \frac{2^{-k}}{(2^{-k}+| x-y|)^{n+1}}+\frac{2^{-k}}{(2^{-k}+| x-y'|)^{n+1}}\right] \end{aligned}$$

provided \(\varepsilon >\varepsilon '\). Hence, for any \(\varepsilon '\in (0,1)\), when \(|y-y'|\le \frac{1}{2}|y-x|\),

$$\begin{aligned} \sum _{k\in \mathbb {N}}|A^{1}_{k}(x,y)|\lesssim & {} \sum _{k\in \mathbb {N}}\left( \frac{\mid y-y'\mid }{2^{-k}}\right) ^{\varepsilon '}\left[ \frac{2^{-k}}{(2^{-k}+\mid x-y\mid )^{n+1}}+\frac{2^{-k}}{(2^{-k}+| x-y'|)^{n+1}}\right] \\\lesssim & {} \frac{|y-y'|^{\varepsilon '}}{|x-y|^{n+\varepsilon '}}. \end{aligned}$$

At last, for any \(\varepsilon \in (0,1)\),

$$\begin{aligned} |A^{2}_{k}(x,y)|&\lesssim \sum _{k'\ge k}\sum _{Q'\in \Pi _{k'}}|Q'|\frac{2^{-k}}{(2^{-k}+|x-x_{Q'}|)^{n+1}} |\psi _{k'}*b(x_{Q'})|\left( \frac{| y-y'|}{2^{-k'}}\right) ^{\varepsilon }\\&\qquad \cdot \left[ \frac{2^{-k'}}{(2^{-k'}+| x_{Q'}-y|)^{n+1}}+\frac{2^{-k'}}{\left( 2^{-k'}+| x_{Q'}-y'|\right) ^{n+1}}\right] \\&\quad \lesssim \sum _{k'> k}\sum _{Q'\in \Pi _{k'}}\int _{Q'}\frac{2^{-k}}{(2^{-k}+|x-u|)^{n+1}} |\psi _{k'}*b(x_{Q'})|(\frac{| y-y'|}{2^{-k'}})^{\varepsilon }\\&\qquad \cdot \left[ \frac{2^{-k'}}{(2^{-k'}+| u-y|)^{n+1}}+\frac{2^{-k'}}{(2^{-k'}+| u-y'|)^{n+1}}\right] du\\&\quad \lesssim \Vert b\Vert _{Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})}\sum _{k'> k}2^{-k'}\sum _{Q'\in \Pi _{k'}}\int _{Q'}\frac{2^{-k}}{(2^{-k}+|x-u|)^{n+1}} \left( \frac{| y-y'|}{2^{-k'}}\right) ^{\varepsilon }\\&\qquad \cdot \left[ \frac{2^{-k'}}{(2^{-k'}+| u-y|)^{n+1}}+\frac{2^{-k'}}{(2^{-k'}+| u-y'|)^{n+1}}\right] \mathrm{d}u\\&\quad \lesssim \Vert b\Vert _{Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})}\sum _{k'> k}2^{-k'}(\frac{| y-y'|}{2^{-k'}})^{\varepsilon }\frac{2^{-k}}{(2^{-k}+| x-y|)^{n+1}}\\&\quad \lesssim \Vert b\Vert _{Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})}\frac{2^{-k}}{(2^{-k}+| x-y|)^{n+1}}|y-y'|^{\varepsilon }\sum _{k'> k}2^{-k'(1-\varepsilon )}\\&\quad \lesssim \Vert b\Vert _{Lip^{\frac{n}{n+1}}(\mathbb {R}^{n})}\frac{2^{-k}}{(2^{-k}+| x-y|)^{n+1}}|y-y'|^{\varepsilon }. \end{aligned}$$

Thus, for any \(\varepsilon \in (0,1)\),

$$\begin{aligned} \sum _{k\in \mathbb {N}}|A^{2}_{k}(x,y)|\lesssim & {} \frac{|y-y'|^{\varepsilon }}{|x-y|^{n+\varepsilon }}. \end{aligned}$$

Then, we complete the proof. \(\square \)