1 Introduction
The well-posedness of the classical Naiver–Stokes equations and the corresponding fractional versions has attracted many mathematicians’ attention in recent decades. In this article, we will introduce new logarithmic Q-type spaces to cover many known critical spaces for the classical Naiver–Stokes equations and the fractional versions, which can be defined on the half-space
$\mathbb {R}^{n+1}_{+}= \mathbb {R}^{n}\times (0,\infty ) , n\geq 2$
as follows:

where
$(-\Delta )^{\beta }$
represents the fractional Laplacian defined by the Fourier transform in the space variable:

Here, it is appropriate to point out that when
$\beta =1,$
(1.1) becomes the classical Naiver–Stokes equations. Moreover, (1.1) can be seen as a generalization of the two-dimensional quasi-geostrophic equation, which has continued to attract attention extensively.
Equations (1.1) are invariant under the following dilations:

A function space
$X(\mathbb {R}^{n})$
is critical for (1.1) if it is invariant under the scaling

When
$\beta =1$
, the well-posedness of equations (1.1) has been established in the following critical spaces but not in
$\dot {B}^{-1}_{\infty ,\infty }(\mathbb {R}^n),$

Kato [Reference Kato18] initiated the study of the Naiver–Stokes equations in critical spaces by proving that (1.1) is locally well-posed in
$L^n(\mathbb {R}^n)$
and globally well-posed with the initial data being small in
$L^n(\mathbb {R}^n).$
In 2001, Koch and Tataru [Reference Koch and Tataru19] established the well-posedness of the classical Naiver–Stokes equations in
$BMO^{-1}(\mathbb {R}^{n}),$
which is the largest critical space where the well-posedness has been established.
Recently, Q-type spaces and their equivalent characterizations have been used to study the well-posedness of the Naiver–Stokes equations. In the literature, Q-type spaces were introduced as a new space between
$W^{1,n}(\mathbb {R}^{n})$
and
$BMO(\mathbb {R}^{n})$
. In 1995, Aulaskari, Xiao, and Zhao first introduced a class of Möbius invariant analytic function spaces
$Q_{p}(\mathbb {D})$
for
$p\in (0,1)$
on the unit disc
$\mathbb {D}$
of the complex plane. The class
$Q_{p}(\mathbb {D})$
,
$p\in (0,1)$
, can be seen as subspaces and subsets of
$BMOA$
and
$UBC$
on
$\mathbb {D}$
. Since then, the Q-type spaces
$Q_{p}(\mathbb {D})$
were investigated extensively and various characterizations of
$Q_{p}(\mathbb {D})$
have been established (see [Reference Aulaskari, Stegenga and Xiao1, Reference Aulaskari, Xiao and Zhao3, Reference Xiao40, Reference Xiao42] and the references therein). As a class of analytic function spaces, the boundary of
$Q_{p}(\mathbb {D})$
is
$Q_{p}(\partial \mathbb {D})$
introduced by Nicolau and Xiao in [Reference Nicolau and Xiao28], where
$\partial \mathbb D$
denotes the boundary of
$\mathbb D$
.
Correspondingly, in the setting of Euclidean spaces, the real-variable Q-type spaces
$Q_{\alpha }(\mathbb R^{n})$
were first introduced by Essén, Janson, Peng, and Xiao [Reference Essén, Janson, Peng and Xiao13]. For
$\alpha \in (-\infty ,\infty )$
, the Q-type space
$Q_{\alpha }(\mathbb R^{n})$
is defined as the set of all measurable functions on
$\mathbb {R}^n$
that satisfy

Here and henceforth, the symbol
$\sup _{I}$
denotes the supremum taken over all cubes I with the edge length
$\ell (I)$
and the edges parallel to the coordinate axes in
$\mathbb {R}^n$
. In 2004, applying Hausdorff capacities and the tent spaces, Dafni and Xiao [Reference Dafni and Xiao11] established the Carleson measure characterization and the predual space of
$Q_{\alpha }(\mathbb R^{n})$
denoted by
$HH^{1}_{-\alpha }(\mathbb R^{n})$
. Follow-up studies have shown that the spaces
$Q_{\alpha }(\mathbb R^{n})$
are essentially Campanato–Sobolev spaces (see [Reference Wu and Xie37, Reference Xiao43]). Unlike the case of
$Q_{p}(\mathbb D)$
, the Q-type spaces of meromorphic functions defined via the Green function are not equivalent to the ones generated by the Möbius transformation (see [Reference Aulaskari, Wulan and Zhao2, Reference Wulan38]). To exhibit a more general structure behind this difference, a class of Q-type spaces related to weight functions denoted by
$Q_{K}(\mathbb D)$
was introduced (see [Reference Essén and Wulan14, Reference Essén, Wulan and Xiao15] for the details). Further, Bao, and Wulan in [Reference Bao and Wulan5] introduced the analog of
$Q_{K}(\mathbb D)$
on
$\mathbb R^{n}$
denoted by
$Q_{K}(\mathbb R^{n})$
, which generalized the Q-type space
$Q_{\alpha }(\mathbb R^{n})$
. For further information on Q-type spaces, we refer the reader to [Reference Cui, Li and Lou9, Reference Cui and Yang10, Reference Dafni and Xiao12, Reference Peng and Yang29, Reference Wang and Xiao31, Reference Xiao39, Reference Xiao44, Reference Xiao45, Reference Yang and Yuan47, Reference Yang and Yuan48, Reference Yang, Qian and Li50, Reference Yang, Qian and Li51] and the references therein.
In [Reference Xiao43], by the aid of the Carleson measure characterizations of
$Q_{\alpha }(\mathbb R^{n})$
obtained in [Reference Dafni and Xiao11], Xiao proved that the equations (1.1) with
$\beta =1$
are well-posed with data in
$Q_{\alpha }^{-1}(\mathbb {R}^{n})=\nabla \cdot (Q_{\alpha }(\mathbb {R}^{n}))^{n}$
being small, where
$\alpha =(0,1)$
. Also, it was proved in [Reference Essén, Janson, Peng and Xiao13] that
$Q_\alpha (\mathbb {R}^{n})=BMO(\mathbb {R}^{n})$
for
$\alpha \in (-{n}/{2},0).$
Hence, the result obtained in [Reference Dafni and Xiao11] generalizes Koch and Tataru’s result on the well-posedness of equations (1.1) from
$BMO^{-1}(\mathbb {R}^{n})$
to
$Q_{\alpha }^{-1}(\mathbb {R}^{n})$
. For the endpoint
$\alpha =0$
, due to lack of the corresponding semigroup characterization, the method of [Reference Xiao43] is invalid for the space
$Q_{0}(\mathbb R^{n})$
. In [Reference Xiao and Zhang46], Xiao and Zhang introduced a logarithmic Q-type space
$\mathcal {Q}_{0}$
which is defined as the set of all measurable functions f satisfying

Here, the symbol
$\nabla _{x}$
denotes the gradient operator for the variable
$x\in \mathbb R^{n}$
,

With
$\beta =1$
and the data in
$\mathcal {Q}^{-1}_{0}(\mathbb R^{n}):=\nabla (\mathcal {Q}_{0}(\mathbb R^{n}))^{n}$
being small, the well-posedness of equations (1.1) was obtained in [Reference Xiao and Zhang46].
The well-posedness of the fractional Naiver–Stokes equations has been established in the fractional counterparts of (1.2). Letting the spatial dimension
$n=3$
, in [Reference Lions26], Lions proved the global existence of the classical solutions of (1.1) when
$\beta \geq {5}/{4}$
. For general spatial dimension n, the existence result in [Reference Lions26] was extended to
$\beta \geq {1}/{2}+{n}/{4}$
by Wu [Reference Wu33]. Moreover, for the important case
$\beta <{1}/{2}+{n}/{4}$
, Wu [Reference Wu34, Reference Wu35] established the global existence for (1.1) in the Besov spaces
$\dot {B}^{1+{n}/{p}-2\beta ,q}_{p}(\mathbb {R}^{n})$
for
$1\leq q\leq \infty $
and for either
${1}/{2}<\beta $
and
$p=2$
or
${1}/{2}<\beta \leq 1$
and
$2<p<\infty $
and in
$\dot {B}^{r,\infty }_{2}(\mathbb {R}^{n})$
with
$r>\max \{1,1+{n}/{p}-2\beta \}$
(see also [Reference Wu36] concerning the corresponding regularity). Following the idea of [Reference Koch and Tataru19, Reference Xiao43], for
$\beta \in (\frac {1}{2},1)$
, Li and Zhai introduced a new critical Q-type spaces
$Q_{\alpha }^{\beta }(\mathbb {R}^{n})$
for
$\alpha +\beta -1\geq 0.$
With data in
$Q_{\alpha }^{\beta , -1}(\mathbb {R}^{n})=\nabla \cdot (Q_{\alpha }^{\beta }(\mathbb {R}^{n}))^{n}$
being small, in [Reference Li and Zhai23], Li and Zhai obtained the well-posedness of equations (1.1) with
${\frac {1}{2}<\beta <1.}$
Zhai in [Reference Zhai53] and Zhai-Yu in [Reference Yu and Zhai52] proved the well-posedness in
$BMO^{-(2\beta -1)}(\mathbb {R}^n)$
and
$\dot {B}^{-(2\beta -1)}_{\infty , \infty }(\mathbb {R}^n),$
correspondingly. For more progress on the well-posedness/ill-posedness of fractional Naiver–Stokes equations, we refer the reader to [Reference Bourgain and Pavlovć6–Reference Colombo, De Lellis and Massaccesi8, Reference Jiao, Huang, Li and Liu17, Reference Li, Xiao and Yang21, Reference Li and Yang22, Reference Lin and Yang25, Reference Wang and Xiao32, Reference Yang and Li49] and the references therein. Moreover, function spaces have also been applied to study fractional magneto-hydrodynamic (MHD; see [Reference Wu33]), the 2
$-D$
quasi-geostrophic equation (see [Reference Li and Zhai24]) and other equations (see [Reference Jiao, Huang, Li and Liu17]). Especially in [Reference Li and Zhai24], Li and Zhai proved that the convolution singular integral operators are bounded on
$Q^{\beta }_{\alpha }(\mathbb R^{n})$
. As an application, the well-posedness of the quasi-geostrophic equation with small data in
$Q_{\alpha }^{\beta , -1}(\mathbb {R}^{n})$
was obtained in [Reference Li and Zhai24].
In this article, we aim to introduce a new class of logarithmic Q-type spaces to cover the
$BMO^{-1}(\mathbb {R}^n), Q^{-1}_\alpha (\mathbb {R}^n), \mathcal {Q}^{-1}_{0}(\mathbb {R}^n), BMO^{-(2\beta -1)}(\mathbb {R}^n), Q_{\alpha }^{\beta ,-1}(\mathbb {R}^n)$
and the largest critical space
$\dot {B}^{-(2\beta -1)}_{\infty ,\infty }(\mathbb {R}^n).$
Definition 1.1 For any
$k\in \mathbb N$
,
$1<p<\infty $
and
$n\geq 2$
, let
$l\geq k$
. Suppose that
$\lambda \geq 0$
. The space
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
consists of all
$f\in L_{loc}^p(\mathbb {R}^n)$
which is the set of all locally
$L^{p}$
-integrable functions, such that

In Lemma 3.14, we deduce a Stegenga-type inequality (3.20). This inequality enables us to establish the extension from
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
to the functions on
$\mathbb R^{n+1}_{+}$
belonging to
$\mathscr H_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb R_+^{n+1})$
via fractional heat semigroups
$\{e^{-t(-\Delta )^{\beta }}\}_{t>0}$
(see Theorems 3.15 and 3.16). As an immediate corollary, a Carleson measure characterization of
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
can be deduced (see Corollary 3.18).
For a function
$u(\cdot ,\cdot )$
on
$\mathbb R^{n+1}_{+}$
, the operator
$\nabla _{(x,t)}$
is defined as

For any cube I on
$\mathbb {R}^n$
, the Carleson box
$S(I)$
is defined by

Theorems 3.5 and 3.15 indicate that the following three conditions are equivalent:

when
$p\in [n/(n-1),2n/(n-1))$
,
$0\leq \lambda <2-p+\frac {p}{n}$
and
$(p-1)n-p<l<(p-1)n.$
Here,
$u(x,t):=f\ast K_{t^{2\beta }}^\beta (x), \frac {1}{2}\leq \beta \leq 1$
, and
$K_{\ln }^{(l,k)}(\cdot )$
denotes a piecewise function on
$(0,\infty )$
which is defined in Definition 3.1. For the endpoint cases
$l=(p-1)n-p$
and
$l=(p-1)n$
, we deduce an embedding relation of
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
(see Theorem 3.19).
Based on the extension result obtained in Theorem 3.15, it is also worth mentioning that, when
$\frac {1}{2}\leq \beta \leq 1, k\in \mathbb N$
and
$l\in (n+2\beta -4, n+2\beta -2)$
, there holds

In [Reference Li and Zhai23], under the assumption that
$\beta \in (\frac {1}{2}, 1]$
, the equivalent relation (1.3) was used to obtain the well-posedness of equations (1.1) with data in

To investigate the well-posedness of equations (1.1) with data in Q-type spaces when
$\beta>1$
, we introduce the following logarithmic Q spaces.
Definition 1.2 Let
$l>n+2\beta -4$
,
$\beta>\frac {1}{2}$
and
$k\in \mathbb N$
. The space
$Q^{(l,k)}_\beta (\mathbb {R}^n)$
is defined as the set of all measurable functions satisfying
$\|f\|_{Q^{(l,k)}_\beta }<\infty ,$
where
$\|f\|_{Q^{(l,k)}_\beta }$
is defined as

We will establish some basic properties of
$Q^{(l,k)}_\beta (\mathbb {R}^n)$
and
$Q_{l,k,\beta }^{-1}(\mathbb {R}^n)=\nabla \cdot (Q^{(l,k)}_\beta (\mathbb {R}^n))^{n}$
in Section 4.1. Assume that
$\beta \in (\frac {1}{2},\min \{2,\frac {1}{2}+\frac {n}{4}\})$
,
$n+2\beta -4<l\leq n+4\beta -4$
and
$k\in \mathbb N$
. By the quadratic estimates for integral operators obtained in Section 4.2, we obtain the well-posedness of solutions to equations (1.1) with data in
$Q_{l,k,\beta }^{-1}(\mathbb {R}^n)$
being small (see Theorem 4.12). For
$\beta =1$
,
$l=n$
and
$k=2$
,
$Q^{(n,2)}_{1}(\mathbb {R}^n)$
comes back to the space
$\mathcal {Q}_{0}$
introduced by Xiao and Zhang [Reference Xiao and Zhang46]. Moreover, for
$n-2<l\leq n$
and
$k\in \mathbb N$
, the well-posedness of equations (1.1) with
$\beta =1$
(i.e., the classical Naiver–Stokes equations) holds for all critical spaces
$Q^{(l,k)}_\beta (\mathbb {R}^n).$
For
$\beta =1$
,
$k=0$
and
$l=n-2\alpha $
,
$Q^{(n-2\alpha , 0)}_{1}(\mathbb R^{n})$
coincides with the Q-type spaces
$Q_{\alpha }(\mathbb R^{n})$
which were introduced by Essén, Janson, Peng, and Xiao [Reference Essén, Janson, Peng and Xiao13]. By [Reference Essén, Janson, Peng and Xiao13, Theorem 2.3(iii)], for
$l>n$
,
$Q^{(l, 0)}_{1}(\mathbb R^{n})=BMO(\mathbb R^{n})$
. Hence, for
$k=0, l>n, \beta =1$
, the well-posedness result for equations (1.1) with data in
$Q^{-1}_{l, 0,1}(\mathbb R^{n})$
has been proved by Koch and Tataru in [Reference Koch and Tataru19]. We also mention that for the special case
$\beta =1$
,
$l=n$
and
$k=2$
, with small data in
$Q^{-1}_{n,2,1}(\mathbb R^{n})$
, the well-posedness of equations (1.1) was obtained by Xiao and Zhang in [Reference Xiao and Zhang46]. Theorem 4.12 indicates that even for the special case
$\beta =1$
, there exist a family of logarithmic Q-type spaces
$Q^{-1}_{l,k,1}(\mathbb R^{n})$
with
$k\in \mathbb N$
and
$l\in (n-2, n]$
such that the solutions to equations (1.1) are well-posed for the data in
$Q^{-1}_{l,k,1}(\mathbb R^{n})$
being small.
In [Reference Li and Zhai23], under the assumption that
$\alpha \in (0,1)$
and
$\beta \in (\max \{\alpha ,\frac {1}{2}\},1)$
with
$\alpha +\beta -1>0$
, Li and Zhai obtained the well-posedness of equations (1.1) with data in
$Q_{\alpha }^{\beta , -1}(\mathbb {R}^{n})$
, where
$Q_{\alpha }^{\beta , -1}(\mathbb {R}^{n})=\nabla \cdot (Q_{\alpha }^{\beta }(\mathbb R^{n}))^{n}$
. Here, the Q-type space
$Q^{\beta }_{\alpha }(\mathbb R^{n})$
is defined as the set of all measurable functions f with

where
$\alpha \in (0, 1)$
and
$\beta \in (\frac {1}{2}, 1)$
. It can be seen from (1.4) that for
$\alpha \in (0,1)$
and
$\beta \in (\max \{\alpha ,\frac {1}{2}\},1)$
with
$\alpha +\beta -1>0$
,
$Q^{\beta }_{\alpha }(\mathbb R^{n})= Q_{\ln , \frac {4(1-\beta )}{n}}^{2, 0, n-2(\alpha -\beta +1)}(\mathbb R^{n})$
. If
$\alpha \in (0,1)$
and
$\beta \in (\max \{\alpha ,\frac {1}{2}\},1)$
with
$\alpha +\beta -1>0$
,
$n+2\beta -4<n-2(\alpha -\beta +1)<n+4\beta -4$
. Then, Theorem 4.12 holds for
$\nabla \cdot (Q_{\ln , \frac {4(1-\beta )}{n}}^{2, 0, n-2(\alpha -\beta +1)}(\mathbb R^{n}))^{n}$
and covers [Reference Li and Zhai23, Theorem 4.14.]. Compared with the well-posedness result obtained by Li and Zhai [Reference Li and Zhai23], in Theorem 4.12, the well-posedness of (1.1) holds for
$\beta \in (\frac {1}{2},\min \{2,\frac {1}{2}+\frac {n}{4}\})$
with small data in the Q-type spaces related to logarithmic type functions, while the scope of
$\beta $
is restricted to
$(\frac {1}{2},1)$
in [Reference Li and Zhai23, Theorem 4.14]. By Proposition 4.3,
$Q^{(l,k)}_{\beta }(\mathbb R^{n})$
is a subclass of the Campanato space
$L^{2, 1+\frac {4(1-\beta )}{n}}(\mathbb R^{n})$
, which indicates that the reasonable scope of
$\beta $
should be
$(\frac {1}{2}, 1+\frac {n}{4}]$
(see Remark 4.4). On the other hand, according to Lions [Reference Lions26] and Wu [Reference Wu33] if
$\beta \geq \frac {1}{2}+\frac {n}{4}$
, there exist classical solutions globally to equations (1.1). Hence, in Theorems 4.12 and 4.13, we assume that
$\beta \in (\frac {1}{2}, \min \{2, \frac {1}{2}+\frac {n}{4}\})$
. By Theorem 3.19, if
$l\leq n+2\beta -4$
, the space
$Q^{(l,k)}_{\beta }(\mathbb R^{n})$
is trivial (see Remark 3.21). On the other hand, for
$\frac {1}{2}<\beta <1$
and
$l>n+4\beta -4$
, Theorem 4.7 implies that the space

coincides with the homogeneous Besov spaces
$\dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb R^{n})$
. Hence, the well-posedness of equations (1.1) with data in
$Q^{-1}_{l,k,\beta }(\mathbb R^{n})$
with
$(\beta , l)\in (\frac {1}{2},1)\times (n+4\beta -4,\infty )$
follows from Yu and Zhai [Reference Yu and Zhai52, Theorem 3.1]. In Theorems 4.12 and 4.13, we assume that
$n+2\beta -4<l\leq n+4\beta -4$
, which is different from that in [Reference Yu and Zhai52, Theorem 3.1]. The method of Theorem 4.12 can also be used to investigate the well-posedness of fractional MHD equations (see Theorem 4.13).
Remark 1.3 It is worth emphasizing that, in Theorem 4.12, the endpoint case
$\beta = 1$
holds particular significance. Observe that the embedding
$ Q_{l,k,\beta }^{-1}(\mathbb {R}^n) \hookrightarrow \dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb {R}^n) $
follows from the scaling and translation invariance of
$ Q_{l,k,\beta }^{-1}(\mathbb {R}^n) $
. However, the converse embedding
$ \dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb {R}^n) \hookrightarrow Q_{l,k,\beta }^{-1}(\mathbb {R}^n) $
requires the restriction
$\beta < 1$
(see Theorem 4.7). This indicates that
$ Q_{l,k,1}^{-1}(\mathbb {R}^n) $
is strictly smaller than
$\dot {B}^{-1}_{\infty ,\infty }(\mathbb {R}^n),$
which aligns with the known fact that the classical Navier–Stokes equations are well-posed in
$ Q_{n,2,1}^{-1}(\mathbb {R}^n) $
(obtained by Xiao and Zhang [Reference Xiao and Zhang46]), despite being ill-posed in
$\dot {B}^{-1}_{\infty ,\infty }(\mathbb {R}^n) $
.
Remark 1.4 By Theorems 2.5 and 2.8, the following relations hold:

In Example 2.4, we provide an example such that
$CIS^{p,\lambda }(\mathbb {R}^n)$
and
$Q^{p,k,l}_{\ln ,\lambda }(\mathbb {R}^n)$
are nontrivial for
$p\in (n/(n-1), 2n/(n-1))$
and
$0\leq \lambda \leq 2+\frac {p}{n}-p$
. Meanwhile if
$p+\lambda -1>1+\frac {p}{n},$
the Campanato space
$L^{p, p+\lambda -1}(\mathbb {R}^n)$
only consists of constants. Considering
$\lambda \geq 0$
in the Definition 1.1, we restrict the scope of p to
$ [n/(n-1), 2n/(n-1))$
in the main results of Section 3.
Some notations: Throughout the rest of this article, denote by
$\mathbb N$
the set of all nonnegative integers.
$U\approx V$
indicates that there is a constant
$c>0$
such that
$c^{-1}V\leq U\leq cV$
, whose right inequality is also written as
$U\lesssim V$
. Similarly, one writes
$V\gtrsim U$
for
$V\geq cU$
. On
$\mathbb {R}^n$
, denote by
$f\ast g$
the convolution of functions f and g, i.e.,

For
$1\leq k<\infty $
,
$\mathscr C^k(\mathbb {R}^n)$
denotes the set of all functions
$f:\mathbb {R}^n\rightarrow \mathbb R$
that is k-time continuously differentiable. For
$1\leq p<\infty $
,
$L^p(\mathbb {R}^n)$
denotes the set of all Lebesgue measurable functions
$f: \mathbb {R}^n\rightarrow \mathbb R$
such that
$\int _{\mathbb {R}^n}|f(x)|^pdx<\infty $
. Denote by
$\mathscr S(\mathbb {R}^n)$
the class of Schwartz functions on
$\mathbb {R}^n$
, denote by
$\mathscr S'(\mathbb {R}^n)$
the dual of
$\mathscr S(\mathbb {R}^n)$
. Denote

where
$|\gamma |=\gamma _1+\gamma _2+\cdots +\gamma _n$
and
$x^\gamma =x_1^{\gamma _1}x_2^{\gamma _2}\cdots x_n^{\gamma _n}$
.
2 Basic properties of
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
As we all know, the classical Q-type space
$Q_\alpha (\mathbb {R}^n)$
is invariant under the rotation, the translation and conformal mappings. It can be deduced that
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
has similar properties; the proof is similar to that of [Reference Bao4, Theorem 2.2.1] and we omit the details.
Lemma 2.1 Suppose that
$k\in \mathbb N$
and
$p\in (1,\infty )$
. Let
$\lambda \geq 0$
and
$l\geq k$
.
-
(i)
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$ is rotation invariant;
-
(ii)
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$ is translation invariant;
-
(iii)
$\|f_\delta \|_{Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)}=\|f\|_{Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)}$ , where
$f_{\delta }(x):=\delta ^{(2-\lambda -p)n/p}f(\delta x)$ ,
$\delta>0$ .
Proposition 2.2 Let
$k\in \mathbb N$
,
$1<p<\infty , n\geq 2$
,
$\lambda \geq 0$
and
$l\geq k$
.
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
if and only if

Proof This proposition can be proved by a change of variables:
$x-y\rightarrow y$
in Definition 1.1 and a direct calculation. So we omit the details.
Motivated by [Reference Xiao41], to show that
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
is well defined, we need the following function spaces.
Definition 2.3 Suppose that
$p\in (1,\infty )$
and
$\lambda \geq 0$
. The space
$CIS^{p,\lambda }(\mathbb {R}^n)$
is defined to be the set of all functions
$f\in \mathscr {C}^{1}(\mathbb R^{n})$
with

We can see that the space
$CIS^{p,\lambda }(\mathbb {R}^n)$
is nontrivial.
Example 2.4 Suppose that
$p\in [n/(n-1),2n/(n-1))$
and
$0\leq \lambda \leq 2+\frac {p}{n}-p$
. Define

Then,
$f_0\in CIS^{p,\lambda }(\mathbb {R}^n)$
.
Proof A direct computation gives, for
$i=1,2,\ldots , n$
,

which gives

Then,

Denote by
$x_{0}$
the center of I. We divide the rest of the proof into two cases.
Case 1:
$|x_{0}|\leq 2\ell (I)$
. For
$x\in I$
,
$|x|\leq |x-x_{0}|+|x_{0}|<3\ell (I)$
. Notice that if
$p>n/(n-1)$
and
$\lambda \geq 0$
, then
$n\lambda -n+np-p>0$
. We can get

Case 2:
$|x_{0}|>2\ell (I)$
. For this case, if
$x\in I$
, then
$|x|\geq |x_{0}|-|x-x_{0}|>\ell (I)$
. Since
$\lambda \leq 2+\frac {p}{n}-p$
, i.e.,
$p-n(\lambda -2+p)\geq 0$
, we obtain

Combining with (2.2) and (2.3), we can see that

which completes the proof.
Theorem 2.5 Suppose that
$k\in \mathbb N$
and
$p\in (1,\infty )$
. Let
$\lambda \geq 0$
and
$l\geq k$
.
$CIS^{p,\lambda }(\mathbb {R}^n)\subseteq Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
. Moreover,
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
is nontrivial if and only if
$l>(p-1)n-p$
.
Proof We prove this theorem by following the idea of [Reference Essén, Janson, Peng and Xiao13, Theorem 2.3]. Suppose that
$f\in CIS^{p,\lambda }(\mathbb {R}^n)$
. For a cube I, since

we can deduce that

By Minkowski’s inequality, we have

Let
$u=1/t$
,
$v=e\sqrt {n}u$
and
$w=\ln v$
. We can deduce from
$l>(p-1)n-p$
that

Since

it holds

Therefore,
$CIS^{p,\lambda }(\mathbb {R}^n)\subseteq Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
and
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
is nontrivial.
Conversely, suppose that
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
is nontrivial. If
$l\leq (p-1)n-p$
, take
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)\cap \mathscr C^1(\mathbb {R}^n)$
. Without loss of generality, assume that f is real-valued and not an identical constant. Then, there exists a point satisfying
$\nabla f(x)\neq 0$
, where
$x=(x_1,x_2,\ldots , x_n)$
. According to the Householder reflector [Reference Trefethen and Bau30, p. 71], there is an orthogonal matrix
$A=(a_{ij}), i,j=1,2,\ldots ,n$
, satisfying

which implies that

Let
$A^T$
be the transpose matrix of A and
$\varphi (x)=f(xA^T)$
. We can see that det
$(A^T)\neq 0$
. There exists a point
$y=(y_1,y_2,\ldots ,y_n)$
such that
$yA^T=x$
and

Hence,
$\nabla \varphi (x)=(|\nabla f(x)|,0,\ldots ,0)$
. Notice that
$\varphi \in \mathscr C^1(\mathbb {R}^n)$
, for any
$\varepsilon>0$
, there exists a small cube I centered at
$x_0$
and a constant
$\delta>0$
. If
$|y-x|<\delta $
, it holds

Letting
$\varepsilon =|\nabla f(x)|/3$
, we can get
$\partial \varphi (y)/\partial y_1>2\delta $
and
$\partial \varphi (y)/\partial y_i>\delta ,\ j\geq 2$
. Define

If
$x,y\in I$
and
$x-y\in M$
, it can be deduced from the mean value theorem that
$|\varphi (x)-\varphi (y)|>\delta |x_1-y_1|.$
Therefore, by the fact that
$|z|\approx z_1, z\in M$
, we have

which implies that
$\varphi \notin Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
and it is a contraction. For
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
and
$\varphi \in L^1(\mathbb {R}^n)$
, notice that
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
is conformal invariant. By Minkowski’s inequality, we have
$f\ast \varphi \in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
and

In addition, if
$\varphi $
is a compactly supported smooth function, then
$f\ast \varphi \in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)\cap \mathscr C^1(\mathbb {R}^n)$
is a constant almost everywhere. By [Reference Bao and Wulan5, p. 3], there is a consequence
$\{\varphi _n\}$
with
$\varphi _n\geq 0$
,
$\int \varphi _n=1$
and
$\text {supp}\ \varphi _n$
shrinking to the empty set such that
$f\ast \varphi \rightarrow f$
almost everywhere. So f is constant almost everywhere. This completes the proof of Theorem 2.5.
Let f be a locally integrable function on
$\mathbb R^{n}$
. For any cube
$I\subset \mathbb R^{n}$
, denote by
$f_{I}$
the mean of f on I, i.e.,

Definition 2.6 Let
$p\in (1,\infty )$
and
$\gamma \in (0,1+\frac {p}{n})$
. The Campanato space
$L^{p,\gamma }(\mathbb {R}^{n})$
is defined as the set of all locally integrable functions f such that

Remark 2.7 It follows from a direct computation that
$f\in L^{p,\gamma }(\mathbb R^{n})$
,
$p\in (1,\infty )\ \&\ \gamma \in (0,1+\frac {p}{n})$
, if and only if

Theorem 2.8 Suppose that
$k\in \mathbb N$
and
$p\in (1,\infty )$
. Let
$\lambda \geq 0$
and
$l\geq k$
. Then,
-
(i)
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)\subseteq L^{p,p+\lambda -1}(\mathbb {R}^n)$ ;
-
(ii) if
$l>(p-1)n$ ,
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)=L^{p,p+\lambda -1}(\mathbb {R}^n)$ .
Proof (i) Suppose that
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
. For any cube I and
$x,y\in I$
, if a positive constant
$C_1$
is small enough, there exists another constant
$C_2$
such that

Define
$J(x,z)=\min \bigg \{|x-z|^{l}\left (\ln \left (\frac {e\sqrt {n}\ell (I)}{|x-z|}\right )\right )^k,|y-z|^{l}\left (\ln \left (\frac {e\sqrt {n}\ell (I)}{|y-z|}\right )\right )^k\bigg \}.$
By the fact that
$t^l\left (\ln (\frac {e\sqrt {n}}{t})\right )^k$
is nondecreasing, we have

Therefore,

which implies that
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)\subseteq L^{p,p+\lambda -1}(\mathbb {R}^n)$
.
(ii) Suppose that
$f\in L^{p,p+\lambda -1}(\mathbb {R}^n)$
. For a cube I and any
$y\in \mathbb {R}^n$
such that
$|y|<\sqrt {n}\ell (I)$
, we can deduce that

Since
$l>(p-1)n$
, it holds

Therefore,
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n) = L^{p,p+\lambda -1}(\mathbb {R}^n)$
.
Corollary 2.9 For
$p\in (2n/(n-1),\infty )$
,
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
contains constant functions only.
Remark 2.10 Theorems 2.5 and 2.8 indicate the following inclusion relations: for fixed
$k, l$
with
$l\geq k$
, if
$\lambda \geq 0$
,

which means that for
$\lambda> 2+\frac {p}{n}-p$
,
$Q^{p,k,l}_{\ln ,\lambda }(\mathbb {R}^n)$
is trivial, i.e., only consists of constant functions.
Theorem 2.11 Suppose that
$k\in \mathbb N$
and
$p\in (1,2n/(n-1))$
. Let
$l\geq k$
,
$0<\lambda <2+\frac {p}{n}-p$
and
$(p-1)n-p<l<(p-1)n$
. Define

with
$s^l\left (\ln \left (\frac {e\sqrt {n}}{s}\right )\right )^k>0$
for some
$s>0$
. Then,
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
if and only if

Proof If
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
, by the definition of
$M^{(l,k)}_{\ln }(\cdot )$
, we can deduce that

and

Conversely, if (2.4) holds, for any cube I, split

where

$I_1(s)=\{x\in I:|x-y|\leq s\ell (I)\}, I_2(s)=\{x\in I:s\ell (I)\leq |x-y|\leq \sqrt {n}\ell (I)\}.$
For
$A_1$
, it is easy to see that

For
$A_2$
, we know

Since
$t^l\left (\ln \left (\frac {e\sqrt {n}}{t}\right )\right )^k$
is nondecreasing on
$(0,\sqrt {n})$
for
$l\geq k$
,

which implies that
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
.
3 Extension via fractional heat semigroups
To characterize
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
via fractional heat semigroups, we introduce the following Q-type spaces related to piece-wise functions.
Definition 3.1 For any
$k\in \mathbb N$
,
$1<p<2n/(n-1)$
and
$n\geq 2$
, let
$l\geq k$
and

Suppose that
$\lambda \geq 0$
. The space
$Q_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb {R}^n)$
consists of all measurable functions of
$f\in L_{loc}^p(\mathbb {R}^n)$
such that

Remark 3.2 The function
$K_{\ln }^{(l,k)}(\cdot )$
defined in (3.1) is nondecreasing. Also
$K_{\ln }^{(l,k)}(2t)\approx K_{\ln }^{(l,k)}(t)$
on
$(0,\infty )$
. In fact, when
$0<t\leq \sqrt {n}$
and
$2t>\sqrt {n}$
, it is obvious that
$K_{\ln }^{(l,k)}(2t)\geq K_{\ln }^{(l,k)}(t)$
. On the other hand,

which indicates that
$K_{\ln }^{(l,k)}(2t)\approx K_{\ln }^{(l,k)}(t)$
. When
$0<t\leq 2t\leq \sqrt {n}$
, we have
$K_{\ln }^{(l,k)}(2t)\geq K_{\ln }^{(l,k)}(t)$
. By the inequality:

we get
$K_{\ln }^{(l,k)}(2t)\approx K_{\ln }^{(l,k)}(t).$
For the case
$2t>t>\sqrt {n}$
, it is obvious that
$t^l\approx (2t)^l$
. We can also get
$K_{\ln }^{(l,k)}(2t)\approx K_{\ln }^{(l,k)}(t).$
To further investigate
$Q_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb {R}^n)$
, following the idea of [Reference Bao and Wulan5, Reference Essén, Wulan and Xiao15], we define the following auxiliary function:

Next, we will verify that

and

hold for
$K_{\ln }^{(l,k)}(\cdot )$
.
Proposition 3.3 Suppose that
$k\in \mathbb N$
and
$p\in (1,2n/(n-1))$
. Let
$l\geq k$
,
$n\geq 2$
and
$(p-1)n-p<l<(p-1)n.$
With
$K_{\ln }^{(l,k)}(\cdot )$
and
$\Phi ^{(l,k)}_{\ln }(\cdot )$
defined above, there hold (3.3) and (3.4).
Proof Case 1: If k is an even number, let
$k=2m, m\in \mathbb N$
. Let
$0<t\leq 1$
. First, we will investigate the case of
$0<s<\sqrt {n}.$
When
$0<s<\sqrt {n},$
we get
$0<st<\sqrt {n}$
and

A direct computation gives us

When
$0<s<1$
, then
$\ln s<0$
, which implies that
$h'(t)>0$
,
$h(\cdot )$
is increasing and

When
$1\leq s<\sqrt {n}$
, it holds that
$\ln s>0$
. Then,
$h'(t)<0$
which implies that
$h(\cdot )$
is decreasing and

Next, we investigate the case of
$s\geq \sqrt {n}$
. When
$\sqrt {n}/s<t\leq 1$
, it follows that

It is easy to get
$g'(t)=\frac {2m}{t(1+\frac {1}{2}\ln n-\ln t)^{2m+1}}$
and
$\left (1+\frac {1}{2}\ln n-\ln t\right )^{2m+1}>0,$
which implies that
$g'(t)>0$
and
$g(\cdot )$
is increasing. Then, we get

When
$0<t\leq \sqrt {n}/s$
, it follows that

According to the case of
$0<s<\sqrt {n}$
, we can get

which indicates that
$h(t)$
is decreasing and

Due to (3.2), it holds

By the fact that
$(p-1)n-p<l<(p-1)n$
, we can deduce that

Notice that

Since
$(p-1)n-p<l<(p-1)n$
, it can be deduced that

which gives

Hence, (3.3) holds.
Case 2: If k is odd, let
$k=2m+1, m\in \mathbb N$
. At first, we investigate the case of
$0<s\leq 1$
. By the definition of
$\Phi ^{(l,k)}_{\ln }(\cdot )$
, it follows that
$0<t\leq 1$
and
$0<st\leq 1$
. Then,

A direct computation gives us

Since
$\ln s<0$
,
$(\ln s+\ln t-\frac {1}{2}\ln n-1)^{2m}>0$
and
$t(\ln t-\frac {1}{2}\ln n-1)^{2m+2}>0$
, we can see that
$u'(t)>0$
and
$u(\cdot )$
is increasing, i.e.,

Next, we investigate the case of
$s>1$
. When
$0<t\leq \sqrt {n}/s$
, we can deduce that

Then, we can get

which implies that
$u(\cdot )$
is decreasing and
$u(t)\leq u(0)=1.$
When
$\sqrt {n}/s<t\leq 1$
, we can deduce that

It is easy to get
$v'(t)=\frac {2m+1}{t(1+\frac {1}{2}\ln n-\ln t)^{2m+2}}>0$
and
$v(\cdot )$
is increasing, then

Due to (3.2), it holds

By the fact that
$(p-1)n-p<l<(p-1)n$
, we can deduce that

Notice that

According to (3.5) and (3.6), we can deduce that (3.3) holds. Then, (3.4) follows. This completes the proof of Proposition 3.3.
Definition 3.4 Suppose that
$k\in \mathbb N$
and
$p\in (1,2n/(n-1))$
. Let
$0\leq \lambda <2+\frac {p}{n}-p$
and
$l\geq k$
. Denote

The space
$\mathscr H_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb R_+^{n+1})$
consists of all measurable functions
$u(\cdot ,\cdot )$
such that

For a cube I, denote by
$S(I)$
the Carleson box centered at
$(y,r)$
, where
$r=\text {dist}\{(y,r),\mathbb {R}^n\}=\ell (I)/2$
.
Theorem 3.5 Suppose that
$k\in \mathbb N$
and
$p\in [n/(n-1),2n/(n-1))$
. Let

Then,
$u\in \mathscr H_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb R_+^{n+1})$
if and only if

Proof We can see that (3.7) is equivalent to

First, assume that
$u(\cdot ,\cdot )\in \mathscr H_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb R_+^{n+1})$
. If
$(x,t)\in S(I)$
, then
$|r+t|\leq \frac {3}{2}\ell (I)$
and
$|x-y|\leq \sqrt {n}\ell (I)$
. Furthermore, we have
$(|x-y|^2+|r+t|^2)^{\frac {1}{2}}\lesssim \ell (I)$
and

Therefore, it can be deduced that

Conversely, assume that
$u(\cdot ,\cdot )$
satisfies (3.7). Let
$(y,r)\in \mathbb R^{n+1}_{+}$
and I be a cube parallel to the coordinate centered at y with edge length r. For any integer
$j>0$
, denote by
$I_j$
the cube centred at y and with edge length
$2^j\ell (I)$
. The Carleson box corresponding to
$I_{j}$
is written as
$S(I_j)$
. When
$(x,t)\in S(I)$
,
$(|x-y|^2+|r+t|^2)^{\frac {1}{2}}\geq r$
. When
$(x,t)\in S(I_{j+1})\setminus S(I)$
,
$(|x-y|^2+|r+t|^2)^{\frac {1}{2}}\geq 2^jr$
. Write

where

It follows from (3.2) that

Since
$u(\cdot ,\cdot )$
satisfies (3.7), there holds

which indicates that

Notice that
$K_{\ln }^{(l,k)}(t)\approx K_{\ln }^{(l,k)}(2t)$
. We can deduce from (3.2) that
$\Phi ^{(l,k)}_{\ln }(2^{-j/n})\approx \Phi ^{(l,k)}_{\ln }(2^{-j/n-1})$
. By the fact that
$2^{-j/n-1}<s<2^{-j/n}$
, we have
$2^{-j\lambda }\lesssim s^{n\lambda }$
and
$2^{j/n}<s^{-1}$
. Hence, by (3.3) and
$p\in [n/(n-1),\infty )$
, we have
$n\lambda -1+(p-1)(n-1)\geq 0$
and

The proof of Theorem 3.5 is finished.
Theorem 3.6 Suppose that
$k\in \mathbb N$
and
$p\in (n/(n-1),2n/(n-1))$
. Let
$0\leq \lambda <2+\frac {p}{n}-p$
and
$l\geq k$
. Define

Then,
$f\in Q_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb {R}^n)$
if and only if

Proof At first, let (3.9) hold. For any cube I,

where


with
$I(x):=\{y\in I:\ell (I)\leq |x-y|<\sqrt {n}\ell (I)\}.$
We can estimate
$U_1$
as

It follows from the fact
$\ell (I)\leq |x-y|<\sqrt {n}\ell (I)$
that
$1\leq |x-y|/{\ell (I)}<\sqrt {n}$
and

Hence,

which means
$f\in Q_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb {R}^n)$
.
Now, assume that
$f\in Q_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb {R}^n)$
. Then,

where

and

Since
$\mathcal K(t)=K_{\ln }^{(l,k)}(t), 0<t<1$
and
$|x-y|<\ell (I).$
Then, for
$V_1$
, it holds

For
$V_2$
,

Hence,

which implies (3.9). This completes the proof of Theorem 3.6.
Lemma 3.7 Suppose that
$k\in \mathbb N$
and
$p\in (1,2n/(n-1))$
. Let
$l\geq k$
and

There holds

Proof Define

According to Theorem 3.6, without loss of generality,

By a direct calculation, we know that
$K^*(\cdot )$
is increasing. When
$0< t<1$
, it follows from (3.2) that

Moreover, when
$t\geq 1$
, we can see from the definition of
$\Phi ^{(l,k)}_{\ln }(\cdot )$
that there exists a constant
$C>0$
such that

Hence,

which implies that
$K^*(t)\approx K_{\ln }^{(l,k)}(t)$
on
$(0,\infty )$
. For some sufficiently small
$c>0$
,

Therefore,
$K^*(t)/t^{c+p(n-1)-n}$
is increasing on
$(0,\infty )$
. For
$0<s<\infty $
,

This completes the proof of Lemma 3.7.
Theorem 3.8 Suppose that
$k\in \mathbb N$
and
$p\in (1,2n/(n-1))$
. Let
$l\geq k$
and
$(p-1)n-p<l<(p-1)n$
. There exists
$u\in (0,\infty ]$
such that

Proof According to (3.4) and the definition of
$\Phi ^{(l,k)}_{\ln }(\cdot )$
, it follows that

Hence,

which implies that
$\lim \limits _{t\rightarrow 0}\inf g(t)>0$
, where

Since
$l<(p-1)n$
, we can deduce that
$g(\cdot )$
is decreasing and

Define

By a direct calculation, we can get
$(K^\#(t))'\geq 0$
and
$K_{\ln }^{(l,k)}(t)\lesssim K^\#(t),0<t<\infty $
. If
$0<t<1$
, notice that


We can deduce that

and
$K^\#(t)\approx K_{\ln }^{(l,k)}(t)$
on
$(0,1)$
. If
$1\leq t<\infty $
, without loss of generality, assume that
$K_{\ln }^{(l,k)}(t)=K_{\ln }^{(l,k)}(1)t^{(p-1)(n-1)}$
. It holds

Hence,
$K^\#(t)\approx K_{\ln }^{(l,k)}(t),0<t<\infty $
. Suppose that
$c>0$
is sufficiently small. We can deduce that
$(K^\#(t)t^{(1-p)n+c})'<0$
, which indicates that
$K^\#(t)t^{(1-p)n+c}$
is decreasing. Moreover, for some
$0<u\leq \infty $
, we can get

It follows that

Lemma 3.9 Suppose that
$k\in \mathbb N$
and
$p\in (1,2n/(n-1))$
. Let
$l\geq k$
. The following two statements are true.
-
(i) There holds:
(3.13)$$ \begin{align} \sup\limits_{0<s<\infty}\left(\int_s^\infty K_{\ln}^{(l,k)}(t)t^{(1-p)n-1}dt\right)^{\frac{1}{p}}\left(\int_0^s(K_{\ln}^{(l,k)}(t))^{\frac{1}{1-p}}t^{n-1}dt\right)^{1-\frac{1}{p}}<\infty. \end{align} $$
-
(ii) There exists another nonnegative function
$\mathscr {K}^{(l,k)}_{\ln }(\cdot )$ such that
-
(a)
$\mathscr {K}^{(l,k)}_{\ln }\approx K_{\ln }^{(l,k)}$ on
$[0,\infty );$
-
(b) for some small enough
$c>0, \mathscr {K}^{(l,k)}_{\ln }(t)t^{(1-p)n+c}$ is decreasing;
-
(c)
$\mathscr {K}^{(l,k)}_{\ln }(t)t^{(1-p)n}$ is also decreasing.
Proof (i) Since
$K_{\ln }^{(l,k)}$
satisfies (3.11), if
$u=\infty $
, it is easy to get (3.13) holds. We divide the proof into two cases if
$0<u<\infty $
.
Case 1:
$0<s<u$
. Since (3.11) holds for
$K_{\ln }^{(l,k)}$
,

Notice that
$K_{\ln }^{(l,k)}(t)=K_{\ln }^{(l,k)}(1)t^{(p-1)(n-1)}, t\geq 1$
. It follows that

Let
$s=u/2$
in (3.11). We can deduce that

Therefore,

Case 2:
$u\leq s<\infty $
. When
$s<1$
,

When
$s\geq 1$
, due to (3.11), we can see that

and

which indicates that

Therefore,

Combining (3.14) with (3.15), it holds

(ii) Define

By a direct calculation,
$\mathscr K^{(l,k)}_{\ln }(t)\approx K_{\ln }^{(l,k)}(1)t^{(p-1)(n-1)},$
when
$t\geq 1$
. Since
$K_{\ln }^{(l,k)}(\cdot )$
is increasing, we can deduce

which implies that
$\mathscr K^{(l,k)}_{\ln }(\cdot )\approx K_{\ln }^{(l,k)}(\cdot ), 0\leq t<\infty $
. Suppose that
$c>0$
is sufficiently small. We have

Hence,
$\mathscr K^{(l,k)}_{\ln }(t)t^{(1-p)n+c}$
and
$\mathscr K^{(l,k)}_{\ln }(t)t^{(1-p)n}$
are decreasing. The proof of Lemma 3.9 is completed.
We need the following Hardy-type inequalities to investigate the characterizations of
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
.
Lemma 3.10 [Reference Kufner and Persson20, p. 4, equations (0.10) and (0.12)]
Suppose that
$f\geq 0$
is a measurable function,
$0<m\leq \infty $
and
$\frac {1}{p}+1/q=1$
such that
$1<p,q<\infty $
. Let
$\sigma $
and
$\tau $
be nonnegative functions on
$(0,h)$
. Then,
-
(i)
$$ \begin{align*}\int_0^h\left(\int_0^sf(t)dt\right)^p\sigma(s)ds\lesssim\int_0^h(f(s))^p\tau(s)ds\end{align*} $$
$$ \begin{align*}\sup\limits_{0<s<h}\left(\int_s^h\sigma(t)dt\right)^{\frac{1}{p}}\left(\int_0^s(\tau(t))^{1-q}dt\right)^{\frac{1}{q}}<\infty.\end{align*} $$
-
(ii)
$$ \begin{align*}\int_0^h\left(\int_s^hf(t)dt\right)^p\sigma(s)ds\lesssim\int_0^h(f(s))^p\tau(s)ds\end{align*} $$
$$ \begin{align*}\sup\limits_{0<s<h}\left(\int_0^s\sigma(t)dt\right)^{\frac{1}{p}}\left(\int_s^h(\tau(t))^{1-q}dt\right)^{\frac{1}{q}}<\infty.\end{align*} $$
Suppose that f is a measurable function on
$\mathbb R^{n}$
satisfying

Define the semigroup operator with a dimensional constant c as

where
$\Delta $
denotes the Laplace operator, a second-order differential operator in an n-dimensional Euclidean space. Denote by
$K_{t^{2\beta }}^\beta (\cdot )$
the kernel of the operator
$e^{-t^{2\beta }(-\Delta )^{\beta }}$
, i.e.,

and denote by
$u(\cdot ,\cdot )$
the convolution of f and
$K_{t^{2\beta }}^\beta $
, i.e.,

If
$\phi \in \mathscr S(\mathbb {R}^n)$
, then
$e^{-t^{2\beta }(-\Delta )^{\beta }}\phi (x):=K_{t^{2\beta }}^\beta \ast \phi (x)$
. In [Reference Miao, Yuan and Zhang27, Remark 2.1], by the Fourier transforms, Miao et al. obtained the following regularity estimate of fractional heat kernels
$K_{t^{2\beta }}^\beta (\cdot )$
:

In fact, the estimate for
$|\nabla _{x}K_{t^{2\beta }}^\beta (x)|$
in (3.18) can be improved.
Lemma 3.11 Let
$\beta \in (0,1)$
,
$m\in \mathbb {Z}_{+}$
. There exists a positive constant C such that for
$t>0$
,

Proof Since
$\{e^{-t(-\Delta )^{\beta }}\}_{t>0}$
,
$\beta \in (0,1)$
, is an analytic semigroup, the estimate for
$|t^{m}\partial _{t}^{m}K^{\beta }_{t}(\cdot )|$
can be deduced by that of
$K^{\beta }_{t}(\cdot )$
and the Cauchy integral formula. We omit the details.
Lemma 3.12 Let
$\beta \in (0,1)$
. There exists a constant
$C>0$
such that for all
$x,y\in \mathbb {R}^{n}$
and
$t>0$
,

Proof Denote by
$K_{t}(\cdot )$
the heat kernel of Laplace operator
$(-\Delta )$
. The subordinate formula gives

where
$\eta _{t}^{\beta }(\cdot )$
satisfies (cf. [Reference Grigor’yan16])

A direct computation gives

By (3.19) and the change of variables:
$s=t^{\frac {1}{\beta }}u$
and
$r^2={|x|^2}/{(t^{\frac {1}{\beta }}u)},$
we obtain

On the other hand, we can get

Thus, it is easy to see that

Case 1:
$0\leq t^{\frac {1}{2\beta }}\leq |x|$
. For this case,
${t}/{|x|^{2\beta +n+1}}\leq {1}/{t^{\frac {n+1}{2\beta }}}$
. Moreover, it leads to

Case 2:
$t^{\frac {1}{2\beta }}> |x|$
. It holds

Remark 3.13 Notice that

It can be deduced from Lemmas 3.11 and 3.12 that

which implies

The semigroup characterization of
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
is based on the following Stegenga-type inequality.
Lemma 3.14 For any
$k\in \mathbb N$
and
$p\in (1,\infty )$
, let
$\lambda \geq 0$
,
$l\geq k$
and
$(p-1)n-p<l<(p-1)n$
. If
$f\in L_{loc}^1(\mathbb {R}^n)$
and satisfies (3.16) and
$u(x,t)=f\ast K_{t^{2\beta }}^\beta (x)$
with
$\beta \geq \frac {1}{2},$
then there holds

where I and J are cubes centered at
$x_0$
with
$\ell (J)=3\ell (I)$
.
Proof We prove (3.20) following the idea of [Reference Essén, Janson, Peng and Xiao13, Lemma 3.1]. Without loss of generality, assume that
$x_0=0$
. Let
$\vartheta $
be a function such that

Split
$u(x,t){\kern-1pt}={\kern-1pt}u_1(x,t){\kern-1pt}+{\kern-1pt}u_2(x,t){\kern-1pt}+{\kern-1pt}u_3(x,t)$
, where
$u_1(x,t){\kern-1pt}={\kern-1pt}f_J\ast K_{t^{2\beta }}^\beta (x,t), u_2(x,t)= ((f-f_J)\vartheta )\ast K_{t^{2\beta }}^\beta (x,t), u_3(x,t)=((f-f_J)(1-\vartheta ))\ast K_{t^{2\beta }}^\beta (x,t)$
. Since
$f_J$
is a constant and
$\int _{\mathbb {R}^n}\nabla _{(x,t)}K_{t^{2\beta }}^\beta (x)dx=0$
, we have
$\nabla _{(x,t)} u(x,t)=\nabla _{(x,t)} u_2(x,t)+\nabla _{(x,t)} u_3(x,t)$
. Together with (3.17), we get

It follows from (3.18) that
$|\nabla _{(x,t)}K_{t^{2\beta }}^\beta (x)|\lesssim t^{-(n+1)}$
,
$|\nabla _{(x,t)}K_{t^{2\beta }}^\beta (x)|\lesssim |x|^{-(n+1)}$
and

Write
$y=s\zeta \in \mathbb {R}^n$
, where
$s=|y|$
and
$|\zeta |=1$
. Define

This gives

In Lemma 3.10, let
$\sigma _1(t):=K_{\ln }^{(l,k)}(t)t^{(1-2p)n-1}$
and
$\tau _1(t):=K_{\ln }^{(l,k)}(t)t^{(1-2p)n-1+p}. $
Since

we have

Let
$\sigma _2(t):=K_{\ln }^{(l,k)}(t)t^{(p-1)(1-n)}$
and
$\tau _2(t):=K_{\ln }^{(l,k)}(t)t^{n(1-p)-1+2p}.$
By Lemma 3.7, we have

By (3.22) and (3.21), it can be deduced from Lemma 3.10 that

Next, we deal with

Similarly to the above estimate for
$ u,$
we can get

where
$f_2:=(f-f_J)\vartheta $
and

Since
$0\leq \vartheta \leq 1$
and
$|\vartheta (x)-\vartheta (y)|\lesssim |x-y|/\ell (J), x,y\in \mathbb {R}^n,$
there holds

Thus,

Since
$l>(p-1)(n-1)-1$
, we can get

and

For
$M_2$
, since
$x\notin J$
,
$y\in (3/4)J$
and
$l<(p-1)n$
, we can get

which implies that
$M_2$
can be controlled by

Similarly to
$M_{2}$
, for
$M_{3}$
, we obtain

If
$(x,t)\in S(I)$
and
$y\in \mathbb {R}^n\setminus (2/3)J$
, then

and

which completes the proof of Lemma 3.14.
Now, by the idea of [Reference Essén, Janson, Peng and Xiao13, Theorem 3.2], we give the characterization of
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
via fractional heat semigroups
$\{e^{-t(-\Delta )^{\beta }}\}_{t>0}$
.
Theorem 3.15 For any
$k\in \mathbb N$
and
$p\in [n/(n-1),2n/(n-1))$
, suppose that

Let
$f\in L_{loc}^p(\mathbb {R}^n)$
satisfying (3.16). Denote
$u(x,t)=f\ast K_{t^{2\beta }}^\beta (x)$
with
$\beta \geq \frac {1}{2}.$
There holds

Proof First, let
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
. Let I and J be two cubes centered at the
$x_0$
such that
$\ell (J)=3\ell (I)$
. Without loss of generality, assume that
$x_0=0$
. Applying H
$\ddot {\text {o}}$
lder’s inequality, we can get

and

Notice that

A direct calculation gives

Theorem 2.8 implies
$f\in L^{p,p+\lambda -1}(\mathbb {R}^n)$
and
$\|f\|_{L^{p,p+\lambda -1}(\mathbb {R}^n)}\lesssim \|f\|_{Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)}$
. Lemma 3.14 and Theorem 3.5 indicate that

Conversely, if
$u(\cdot ,\cdot )\in \mathscr H_{K_{\ln }^{(l,k)}}^{p,\lambda }(\mathbb R_+^{n+1})$
, we will prove that

It is easy to see that
$|f(x+y)-f(x)|$
can be controlled by

For
$U_3$
, notice that
$U_3=|u(x,|y|)-u(x,0)|\leq \int _0^{|y|}|\nabla _ru(x,r)|dr.$
Similarly to Lemma 3.14, we can apply Lemma 3.10 and Minkowski’s inequality to deduce that

and

According to the fact,

for any y with
$|y|<\ell (I),$
we can get

For
$U_2$
, it is easy to see that

It follows from Minkowski’s inequality that

Therefore, it yields

Hence, the inequality (3.23) holds, and this completes the proof of Theorem 3.15.
Following the procedure of Theorem 3.15, we can deduce the following result.
Theorem 3.16 For any
$k\in \mathbb N$
and
$p\in [n/(n-1),2n/(n-1))$
, suppose that

Let
$f\in L_{loc}^p(\mathbb {R}^n)$
satisfying (3.16). Denote
$u(x,t)=f\ast K_{t^{2\beta }}^\beta (x), (x,t)\in \mathbb R_+^{n+1}.$
The following statements are equivalent:

By the aid of Theorems 3.15 and 3.16, we can characterize
$Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$
via the following Carleson-type measures related to logarithmic functions.
Definition 3.17 Let
$(k,l,\lambda )\in \mathbb {N}\times (0,\infty ) \times (0,\infty )$
. A positive Borel measure
$\mu (\cdot ,\cdot )$
on
$\mathbb R^{n+1}_{+}$
is said to be a
$(k,l,\lambda )$
-logarithmic Carleson measure if

The following Carleson measure characterizations can be deduced from Theorems 3.15 and 3.16 immediately.
Corollary 3.18 For any
$k\in \mathbb N$
and
$p\in [n/(n-1),2n/(n-1))$
, suppose that

-
(i)
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$ if and only if
$$ \begin{align*}d\mu(x,t):=|\nabla_{(x,t)}e^{-t^{2\beta}(-\Delta)^{\beta}}f(x)|^{p}t^{-(p-1)(n-1)}dxdt\end{align*} $$
$(k,l,\lambda )$ -logarithmic Carleson measure.
-
(ii)
$f\in Q_{\ln ,\lambda }^{p,k,l}(\mathbb {R}^n)$ if and only if
$$ \begin{align*}d\nu(x,t):=|\nabla_{x}e^{-t^{2\beta}(-\Delta)^{\beta}}f(x)|^{p}t^{-(p-1)(n-1)}dxdt\end{align*} $$
$(k,l,\lambda )$ -logarithmic Carleson measure.
Notice that for
$k\geq 0$
, the term

which, together with Theorem 3.16, for
$0\leq \lambda <2-p+\frac {p}{n}$
,
$l\geq k$
and
$(p-1)n-p<l<(p-1)n$
, indicates the following embedding relation:

Below we will prove (3.24) still holds for the endpoint cases
$l= (p-1)n-p$
and
$l=(p-1)n$
.
Theorem 3.19 Suppose that f is a measurable function such that (3.16) and
$k\in \mathbb N$
and
$k\geq 2$
. Let
$\beta \geq \frac {1}{2}$
,
$\lambda \geq 0$
and

If

then

Proof Assume that
$\sigma (x,t)= e^{-t^{2\beta }(-\Delta )^\beta }f(x)$
and
$\sigma (x,0)=f(x)$
. Write

where

According to the symmetry of
$D_i, i=1,2,$
we only need to verify

Letting
$z=x+\frac {ty}{2}, s=\frac {t|y|}{2}$
, we can deduce that

Applying Minkowski’s inequality and the change of variables, we obtain

Therefore, we can see that

When
$l=(p-1)n$
, since
$1<p<k$
, it follows that
$\lim \limits _{u\rightarrow \infty }(\ln s)u^{p-k-1}=0$

and

When
$l\leq (p-1)n-p$
, since
$1<p<k+1$
, it follows that

and

Therefore, for
$l=(p-1)n$
and
$l\leq (p-1)n-p$
,

By Lemma 3.10, we can get

which completes the proof of Theorem 3.19.
From Theorem 3.19, we can deduce the following result.
Corollary 3.20 Let
$\beta \geq \frac {1}{2}, l\leq (p-1)n-p$
,
$\lambda \geq 0$
and
$k\in \mathbb N$
with
$1<p<k+1$
. Suppose that f is a measurable function such that (3.16) holds. If f satisfies (3.25), then f is a constant almost everywhere.
Proof Similarly to the proof of Theorem 2.5, without loss of generality, we assume that f satisfies (3.25) and
$f\in \mathscr C^1(\mathbb {R}^n)$
. By Theorem 3.19, f satisfies (3.26). If f is real-valued and not an identical constant, there exists a point
$x_0$
such that
$\nabla f(x_0)\neq 0$
. According to Definition 1.1 and Lemma 2.1, we can deduce from (3.26) that f is rotation invariant. Thus, we can assume that
$\nabla f(x_0)$
is directed along the positive axis of
$x_1$
. There exists
$\theta>0$
and a small cube I about
$x_0$
such that

Define
$V=\{x: |x_2|+\cdots +|x_n|<x_1<\ell (I)/2\}$
. If
$x,y\in I$
with
$x-y\in V$
, we have
$f(x)-f(y)>\theta (x_1-y_1)$
and

which is a contradiction, and f is a constant almost everywhere.
Remark 3.21 Corollary 3.20 will be used in Section 4 to study the critical logarithmic Q-type spaces
$Q^{(l,k)}_\beta (\mathbb {R}^n)$
(see Definition 1.2) for the well-posedness of equations (1.1). Specially, for
$p=2$
and
$k\geq 2$
, we can deduce from Theorem 3.19 that

Based on Corollary 3.20, for
$l+2-2\beta \leq n-2$
, any function f satisfying

is a constant function. Hence, in the definition of
$Q^{(l,k)}_\beta (\mathbb {R}^n),$
the index l should satisfy
$l>n+2\beta -4$
.
4 Well-posedness of fractional Naiver–Stokes/MHD equations
In this section, with data in the logarithmic Q spaces, we study the well-posedness of the fractional Naiver–Stokes equations (1.1) and the fractional MHD equations with
$\beta>\frac {1}{2}$
. As a matter of convenience, we first state some basic properties of homogeneous Besov spaces. Let
$\mathscr D(\mathbb {R}^n)$
be the space of polynomials. The dual of
$\mathscr {S}_{0}(\mathbb R^{n})$
is

Suppose that
$\varphi $
is a function on
$\mathbb {R}^{n}$
satisfying
$\text { supp }\widehat {\varphi }\subset \{\xi \in \mathbb {R}^{n}: |\xi |\leq 1\}$
and
$\widehat {\varphi }(\xi )=1$
for
$\{\xi \in \mathbb {R}^{n}:\ |\xi |\leq {1}/{2}\}$
. Let

be the Littlewood–Paley functions. For
$1\leq p, q\leq \infty $
and
$s\in \mathbb {R}^n$
, the homogeneous Besov space
$\dot {B}_{p,q}^s(\mathbb {R}^n)$
is defined as the set of all
$f\in \mathscr S_0'(\mathbb {R}^n)$
such that

It is well known that for
$1\leq p<\infty $
,
$0<q<\infty $
and
$-\infty <s<\infty $
,
$(\dot {B}_{p,q}^s(\mathbb {R}^n))'=\dot {B}_{p',q'}^{-s}(\mathbb {R}^n)$
, where
$\frac {1}{p}+\frac {1}{p}'=1$
and
$1/q+1/q'=1$
. The following fractional semigroup characterization of
$\dot {B}_{p,q}^s(\mathbb {R}^n)$
is due to Miao et al. [Reference Miao, Yuan and Zhang27].
Proposition 4.1 [Reference Miao, Yuan and Zhang27, Proposition 2.1] Let
$1\leq p,q\leq \infty $
,
$s<0$
and assume that
$n\in \mathbb N$
and
$0<\beta <\infty $
. Then,
$f\in \dot {B}_{p,q}^s(\mathbb {R}^n)$
if and only if

4.1 Critical Q-type spaces
By the change of variable, we can get the following equivalent norm of
$\|\cdot \|_{Q^{(l,k)}_\beta (\mathbb {R}^n)}.$
Proposition 4.2 Let
$l>n+2\beta -4$
,
$\beta>\frac {1}{2}$
and
$k\in \mathbb N$
.
$f\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
if and only if

Proposition 4.3 Let f be a measurable function satisfying (3.16),
$k\in \mathbb N$
with

and

If
$f\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
, then

Proof Let
$f\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
and
$F(x,t):=e^{-t^{2\beta }(-\Delta )^\beta }f(x)$
. By the triangle inequality, we get

Now, we prove

It is easy to see that

On the other hand, for
$x\in I$
and
$z=x+te_{y}$
,

which gives

For
$0<s<\ell (I)$
,
$\ln \left ({e\ell (I)}/{s}\right )=\ln e+\ln \left ({\ell (I)}/{s}\right )\geq 1$
. We have

Now we turn to estimate
$B_2$
. It holds

which implies

By the change of variables:
$s=u\ell (I)$
and
$t=\ell (I)v$
, we can obtain

Let

Letting
$e/t=u$
and
$u=e^w$
, we get

On the other hand,

Then, when
$\beta =1$
and
$k\geq 2$
, we get

It is easy to see that

When
$\beta \neq 1$
and
$k\geq 1$
,

A direct computation gives

which yields

Applying Lemma 3.10 and the change of variable
$t\ell (I)=s$
, we obtain

Now we estimate
$B_3$
. Since
$|y|<\ell (I)$
, then
$x\in I$
implies that
$x+y\in I+y$
. Hence,

and

This completes the proof of Proposition 4.3.
Remark 4.4 Proposition 4.3 enables us to identify the scope of
$\beta $
such that the space
$Q^{(l,k)}_\beta (\mathbb {R}^n)$
is nontrivial. If
$f\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
, it can be deduced from Proposition 4.3 that f satisfies

Similarly to [Reference Li and Zhai23, Theorem 3.2], we can prove that

i.e.,
$f\in L^{2, 1+\frac {4(1-\beta )}{n}}(\mathbb R^{n})$
. This means that if f is not a constant a.e., then
$0\leq 1+\frac {4(1-\beta )}{n}\leq 1+\frac {n}{4}$
, equivalently,
$\frac {1}{2}\leq \beta < 1+\frac {n}{4}$
.
Definition 4.5 Let
$l>n+2\beta -4$
,
$\beta>\frac {1}{2}$
and
$k\in \mathbb N$
. The symbol
$Q_{l,k,\beta }^{-1}(\mathbb {R}^n)$
denotes the divergence space of
$Q^{(l,k)}_\beta (\mathbb {R}^n)$
, i.e.,

In other words,
$f\in Q_{l,k,\beta }^{-1}(\mathbb {R}^n)$
if and only if there are
$f_1,\ldots ,f_n\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
such that

with
$\|f\|_{Q_{l,k,\beta }^{-1}(\mathbb {R}^n)}=\sum \limits _{j=1}^n\|f_j\|_{Q^{(l,k)}_\beta (\mathbb {R}^n)}$
.
Theorem 4.6 Suppose that
$k\in \mathbb N$
,
$n+2\beta -4<l\leq n+4\beta -4$
,
$\beta>\frac {1}{2}$
. Let f be a tempered distribution on
$\mathbb {R}^n$
. Then,
$f\in Q_{l,k,\beta }^{-1}(\mathbb {R}^n)$
if and only if

Proof Suppose that
$f\in Q_{l,k,\beta }^{-1}(\mathbb {R}^n)$
. It follows that there exist
$f_1,\ldots ,f_n\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
such that
$f(x)=\sum \limits _{j=1}^n\partial _{x_j}f_j(x)$
. Hence,

Conversely, suppose that (4.1) holds. Set
$f_{j,k}(x):=\partial _j\partial _k(-\Delta )^{-1}f(x)$
. If
$\|f\|_{Q_{l,k,\beta }^{-1},*}<\infty $
,
$f_k=-\partial _k(-\Delta )^{-1}f\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
. In fact, we can see that

which implies that
$f=\sum \limits _{k=1}^n\partial _kf_k\in Q_{l,k,\beta }^{-1}(\mathbb {R}^n)$
.
If
$\|f_{j,k}\|_{Q_{l,k,\beta }^{-1},*}<\infty $
with
$f_{j,k}=\partial _j\partial _k(-\Delta )^{-1}f$
, we have

which implies

that is,
$f_k\in Q^{(l,k)}_\beta (\mathbb {R}^n)$
.
Now, we prove
$\|f_{j,k}\|_{Q_{l,k,\beta }^{-1},*}<\infty $
. Take a function
$\phi \in C_0^\infty (\mathbb {R}^n)$
such that

Then,

where
$g_r(x,t)=\phi _r*\partial _j\partial _k(-\Delta )^{-1}e^{-t(-\Delta )^\beta }f(x)$
. Since
$(\dot {B}_{1,1}^{2\beta -1})^*=\dot {B}_{\infty ,\infty }^{1-2\beta },$
we can get

The above estimate yields

Since
$l>n+2\beta -4$
, by the changes of variables
$u=t^{\frac {1}{\beta }}$
,
$er^2/u=v$
and
$v=e^w$
, we can get

and

Now we estimate
$f_r$
. Take
$\varphi \in C_0^\infty (\mathbb {R}^n)$
with

where

So

We first estimate
$M_2$
. Write

Let
$R_j,$
for
$j=1,2,\ldots ,n,$
denote Riesz transforms. By the
$L^2$
- boundedness of
$R_j$
, we can get

For
$M_{2,2}$
, notice that

It follows from the Plancherel formula that

thereby

Now, we estimate
$M_1$
. Since

we can see that

where
$M_{1,k}$
is defined as

Notice that
$\left (\ln \left (\frac {er}{t^{\frac {1}{2\beta }}}\right )\right )^k=\left (\ln \left (\frac {e10^{k+1}r}{10^{k+1}t^{\frac {1}{2\beta }}}\right )\right )^k\lesssim \left (\ln \left (\frac {e10^{k+1}r}{t^{\frac {1}{2\beta }}}\right )\right )^k, $
we obtain

Since
$\beta>\frac {1}{2}$
and
$l\leq n+4\beta -4$
,
$M_1\lesssim \sum \limits _{k=1}^\infty 10^{k(l-n-6\beta +5)}\|f\|_{Q_{l,k,\beta }^{-1},*}^2. $
This completes the proof of Theorem 4.6.
Theorem 4.7 Let
$k\in \mathbb N$
,
$l>n+4\beta -4$
and
$\beta \in (\frac {1}{2},1)$
.

Proof It is easy to see that
$Q^{-1}_{l,k,\beta }(\mathbb R^{n})$
is invariant under the dilation:
$f_{\delta }(x)=\delta ^{2\beta -1}f(\delta x)$
. By [Reference Li and Zhai23, Proposition 4.6],
$Q^{-1}_{l,k,\beta }(\mathbb R^{n})\subseteq \dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb R^{n})$
. Now, we assume that
$f\in \dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb R^{n})$
. Then, for any
$(x,t)\in \mathbb R^{n+1}_{+}$
, it follows from Proposition 4.1 that

which, together with Theorem 4.6, gives

Letting
$s=t^{\frac {1}{2\beta }}$
and
$er/s=v$
, we can get

Since
$\beta \in (\frac {1}{2},1)$
and
$l>n+4\beta -4$
, the integral
$\int ^{\infty }_{e}(\ln v)^{k}v^{n-l+6\beta -7}dv<\infty ,$
which implies that
$\|f\|_{Q^{-1}_{l,k,\beta }}\lesssim \|f\|_{\dot {B}^{1-2\beta }_{\infty ,\infty }}$
, and
$\dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb R^{n})\subseteq Q^{-1}_{l,k,\beta }(\mathbb R^{n})$
. This completes the proof of Theorem 4.7.
4.2 Quadratic estimates for an integral operator
Lemma 4.8 Given
$n\in \mathbb N_+$
and
$k\in \mathbb N$
. Let
$0<l\leq n+4\beta -4$
,
$\beta>\frac {1}{2}$
and

Then,

Proof By Plancherel’s formula and Cauchy–Schwarz’s inequality, we obtain

Noting that

for
$0<s<t$
, we can deduce from
$0<l\leq n+4\beta -4$
that

Therefore,

This completes the proof of Lemma 4.8.
We need the following estimate established by Wang and Xiao [Reference Wang and Xiao32].
Proposition 4.9 [Reference Wang and Xiao32, Lemma 2.1]
For
$\beta>\frac {1}{2}$
, if
$s\in (0,1)$
and
$K_{\beta , s}(\cdot )$
is the kernel of
$(-\Delta )^{1-\beta }(e^{-(-\Delta )^\beta }-e^{-s(-\Delta )^\beta }),$
then

Lemma 4.10 For
$k\in \mathbb N$
and
$\beta>\frac {1}{2}$
. Suppose that
$0<l\leq n+4\beta -4$
and
$N(x,t)$
defined on
$(0,1)\times \mathbb {R}^n$
. Let
$A(\beta ,N)$
be the quantity:

Then,

Proof Let
$L(s):=\frac {\left (\ln \left (\frac {e}{s^{\frac {1}{2\beta }}}\right )\right )^k}{s^{\frac {n-l}{2\beta }+2-\frac {2}{\beta }}}, I=\int _0^1\left \|(-\Delta )^{\frac {1}{2}}e^{-t(-\Delta )^\beta /2}\int _0^tN(\cdot ,s)ds\right \|_{L^2(\mathbb {R}^n)}^2L(t)dt.$
We can see that,

By Proposition 4.9, we have

Notice that
$0<l\leq n+4\beta -4$
. A direct computation gives for
$0<h<r^{2\beta }$
,

thereby

This completes the proof of Lemma 4.10.
4.3 Well-posedness of fractional Naiver–Stokes equations
In this section, we prove that the equations (1.1) is well-posed with data
$f\in Q^{-1}_{l,k,\beta }(\mathbb R^{n})$
being small. By Theorem 4.7,
$Q^{-1}_{l,k,\beta }(\mathbb R^{n})=\dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb R^{n})$
for
$\beta \in (\frac {1}{2},1)$
,
$k\in \mathbb N$
and
$l>n+4\beta -4$
. In [Reference Yu and Zhai52], Yu and Zhai proved that, with data
$f\in \dot {B}^{1-2\beta }_{\infty ,\infty }(\mathbb R^{n})$
being small, the well-posedness of equations (1.1) holds (see [Reference Yu and Zhai52, Theorem 3.1]). Moreover, for
$l\leq n+2\beta -4$
, Remark 3.21 indicates that if
$f\in Q^{(l,k)}_{\beta }(\mathbb R^{n})$
then f is a constant. Equivalently, the space
$Q_{l,k,\beta }^{-1}(\mathbb {R}^n)=\{0\}$
. Hence, in Theorem 4.12, the scope of l is restricted to
$(n+2\beta -4, n+4\beta -4]$
. It is worth mentioning that the scope of l in Theorem 4.12 is different from that of [Reference Yu and Zhai52, Theorem 3.1].
Definition 4.11 Let
$l>n+2\beta -4$
,
$\beta>\frac {1}{2}$
and
$k\in \mathbb N$
. A function
$v(\cdot ,\cdot )$
on
$\mathbb R_+^{n+1}$
belongs to the space
$X_{T}^{l,k,\beta }$
provided
$\|v\|_{X_{T}^{l,k,\beta }}<\infty $
, where

Theorem 4.12 Suppose that
$n\geq 2$
and
$k\in \mathbb N$
. Let
$\frac {1}{2}<\beta <\min \{2, \frac {1}{2}+\frac {n}{4}\}$
and
$n+2\beta -4<l\leq n+4\beta -4$
. There is
$\epsilon _0>0$
such that if the incompressible initial data a satisfies

then equations (1.1) have a unique mild solution

with
$\|u\|_{(X_T^{l,k,\beta })^n}:=\sum \limits _{j=1}^n\|u_{j}\|_{X_T^{l,k,\beta }}<\infty ,$
where

Proof First, by Proposition 4.1,

By Picard’s contraction principle, it is sufficient to verify the bilinear operator

is bounded from
$(X_{T}^{l,k,\beta })^n\times (X_{T}^{l,k,\beta })^n$
to
$(X_{T}^{l,k,\beta })^n$
.
Part I:
$L^\infty $
- bounded. The aim is to prove

If
$\frac {t}{2}\leq s\leq t$
, then

If
$0<s<\frac {t}{2}$
,

This gives

Obviously,

For
$I_3$
, by H
$\ddot {\text {o}}$
lder’s inequality, we have

A direct computation gives

Similarly,
$I_{3,2}\lesssim t^{\frac {n}{4\beta }+\frac {1}{2\beta }-\frac {1}{2}}\|v\|_{(X_{T}^{l,k,\beta })^n}$
, thereby

Part II:
$L^2$
-bounded. We want to establish that

Define the characteristic function on the ball

We divide
$B(u,v)$
into three parts:


and
For
$B_1$
, we get
$e^{-t(-\Delta )^\beta }\mathbb {P}\nabla \cdot f(x)=\int _{\mathbb {R}^n}\nabla K_{j,k,t}^\beta (x-y)f(y)dy.$
Since

we have

Notice that
$|z-x|\geq 10r$
,
$0<s<r^{2\beta }$
and
$|y-x|<r$
, then
$|x-z|\lesssim |y-z|$
. This, together with H
$\ddot {\text {o}}$
lder’s inequality, gives

where

We can see that

The function
$g(h):=\left (\ln \left (\frac {e2^{j+1}r}{h^{\frac {1}{2\beta }}}\right )\right )^{-k}$
is increasing for
$0<h<r^{2\beta }$
, that is,

The above estimate implies

Similarly, we obtain
$I_2\lesssim r^{\frac {1}{2}-\beta }\|v\|_{(X_T^{l,k,\beta })^n}$
, which gives
$|B_1(u,v)|\lesssim r^{1-2\beta }\|u\|_{(X_T^{l,k,\beta })^n}\|v\|_{(X_T^{l,k,\beta })^n}.$
Hence,

Since
$l>n+2\beta -4$
, we have
$\int _0^\infty e^{w(-l+n-4+2\beta )}dw\approx 1$
and

By the change of variables:
$t^{\frac {1}{2\beta }}=u$
,
$v=r/u$
and
$w=\ln v$
, respectively, it holds

Hence,

Now, we deal with the term
$B_2(u,v)$
. Recall that

Obviously, the operator
$(-\Delta )^{\frac {1}{2}}\mathbb {P}\nabla $
is bounded on
$L^2(\mathbb {R}^n)$
. By Lemma 4.8, we obtain

The convolution operator
$t^{\frac {1}{2\beta }-1}(-\Delta )^{\frac {1}{2}-\beta }(I-e^{-t(-\Delta )^\beta })$
is represented as

When
$\beta>\frac {1}{2}$
,
$\lim \limits _{|\zeta |\rightarrow 0}|\zeta |^{1-2\beta }(1-e^{-|\zeta |^{2\beta }})=\lim \limits _{|\zeta |\rightarrow \infty }|\zeta |^{1-2\beta }(1-e^{-|\zeta |^{2\beta }})=0, $
i.e.,
$\sup \limits _{|\zeta |\in (0,\infty )}|\zeta |^{1-2\beta }(1-e^{-|\zeta |^{2\beta }})<\infty .$
So
$t^{\frac {1}{2\beta }-1}(-\Delta )^{\frac {1}{2}-\beta }(I-e^{-t(-\Delta )^\beta }), \beta>\frac {1}{2},$
are bounded on
$L^2(\mathbb {R}^n)$
. Let
. The Cauchy–Schwarz’s inequality implies

Now, for
$B_3(u,v),$
by the change of variables
$t=r^{2\beta }\tau $
and
$s=r^{2\beta }\theta $
, we can get

By the Fourier transform,

On the other hand,
$\widehat {f(r\cdot )}(\eta )=r^{-n}\widehat {f}(\frac {\eta }{r})$
. Denote

Then, we can get

Therefore, by Lemma 4.10, we get

where
$A(\beta ,M(r\cdot , r^\beta \cdot ))$
is defined as

By the change of variables
$r^{2\beta }s=t$
and
$ry=x$
, we can obtain

On the other hand, by the change of variables
$r^{2\beta }s=t$
and
$ry=z$
, we can deduce that
$A(\beta ,M(r\cdot , r^\beta \cdot ))$
can be controlled by

which gives

4.4 Well-posedness of fractional MHD equations
The method of Theorem 4.12 can also be used to investigate the well-posedness of fractional MHD equations in
$\mathbb {R}^{n+1}_{+}$
,
$n\geq 2$
:

with
$\nabla \cdot u=\nabla \cdot b=0$
and
$\beta \in (\frac {1}{2},\min \{2,\frac {1}{2}+\frac {n}{4}\}).$
Following the procedure of Theorem 4.12, we can obtain the following result.
Theorem 4.13 Let
$n\geq 2$
,
$\frac {1}{2}<\beta <\min \{2, \frac {1}{2}+\frac {n}{4}\}$
,
$n+2\beta -4<l\leq n+4\beta -4$
. The fractional MHD equation (4.2) has a unique small global mild solution in
$(X_T^{l,k,\beta })^n$
for an initial data
$(u_0,b_0)$
with
$\nabla \cdot u_0=\nabla \cdot b_0=0$
and small
$\|(u_0,b_0)\|_{(Q_{l,k,\beta }^{-1}(\mathbb {R}^n))^n}$
.
Proof The solution
$(u,b)$
to equation (4.2) can be written as


with
$ B(u,v)=\int _{0}^{t}e^{-(t-s)(-\Delta )^{\beta }}P\nabla \cdot (u\otimes v)(s)ds. $
We rewrite the solution
$(u,b)$
as

Since
$B(u,v)$
is bounded from
$(X_T^{l,k,\beta })^n(\mathbb {R}^n)\times (X_T^{l,k,\beta })^n(\mathbb {R}^n)$
to
$(X_T^{l,k,\beta })^n(\mathbb {R}^n),$
we have


Similarly, we get
$\|F_{2}(u,b)\|_{(X_{T}^{l,k,\beta })^n}\leq C_1\|b_{0} |_{(Q_{l,k,\beta }^{-1}(\mathbb {R}^n))^n}+C_2\|(u,b)\|^2_{(X_{T}^{l,k,\beta })^n} $
and

Thus, we have


Let
$R\leq 2C\|(u_0,b_{0} )|_{(Q_{l,k,\beta }^{-1}(\mathbb {R}^n))^n}$
be small enough. So F is a contraction mapping on

into itself. Thus, a fixed point
$(u,b)$
of the operator F exists.