1 Introduction

The three-dimensional incompressible liquid crystals system are the following coupled equations

$$\begin{aligned} \left\{ \begin{array}{l} {u_t-\Delta u+u\cdot \nabla u +\nabla p=-\mathrm {div}(\nabla d \odot \nabla d),}\\ {\mathrm {div}\ u=0,}\\ {d_t+u\cdot \nabla d-\Delta d=-f(d),}\\ \end{array} \right. \end{aligned}$$
(1.1)

in the domain \(Q_T\equiv \mathbb {R}^3\times (0, T)\). Here, the unknowns \(u=(u_1, u_2, u_3)\) is the velocity field, p is the scalar pressure and \(d=(d_1, d_2, d_3)\) is the optical molecule direction after penalization, and \(f(d)=\frac{1}{\sigma ^2}(|d|^2-1)d\), \(\nabla d \odot \nabla d\) is a symmetric tensor with its component \((\nabla d \odot \nabla d)_{ij}\) is given by \(\partial _i d \cdot \partial _j d\). And the initial conditions are

$$\begin{aligned} u(x,0)=u_0(x), \quad \mathrm {div}(u_0)=0, \quad d(x,0)=d_0(x), \end{aligned}$$
(1.2)

with \(|d_0|=1.\)

System (1.1) is the simplified system of the original Ericksen–Leslie system of variable length for the flow of liquid crystals, that is the Ginzburg–Landau energy \(\int _{\Omega } (\frac{1}{2} |\nabla d |^2 +\frac{(1-|d|^2)^2}{4\sigma ^2})\). For this system, Lin and Liu [19,20,21] first proved a global existence of weak solutions under \(L^2\) data and regularity result of the suitable weak solution under the C-K-N condition. Other results to liquid crystals equations refer to [8,9,10,11, 15, 16, 18, 22, 23, 28].

Let us now recall the notion of a suitable weak solution of liquid crystals equations.

Definition 1.1

([20]) a triple (udp) is called a suitable weak solution of (1.1) in \(\mathbb {R}^3\times (0,T)\) if the following conditions hold:

  1. (1).

    the weak solution (ud) satisfies system (1.1) in the distribution sense;

  2. (2).

    the solution (ud) satisfy the energy inequality, i.e.,

    $$\begin{aligned}&\Vert u\Vert _{L^{2,\infty }(\mathbb {R}^3\times (0,T))}^2+\Vert \nabla d\Vert _{L^{2,\infty }(\mathbb {R}^3\times (0,T))}^2\\&\quad +\Vert \nabla ^2d\Vert _{L^2(\mathbb {R}^3\times (0,T)))}^2 +\Vert \nabla u\Vert _{L^2(\mathbb {R}^3\times (0,T)))}^2\le c_0; \end{aligned}$$
  3. (3).

    the press \(p\in L^{\frac{3}{2}}_{loc}(\mathbb {R}^3\times (0,T))\);

  4. (4).

    the triple (udp) satisfy the modified generalized local energy inequality, for a.e. \(t\in (0,T)\) and for all \(\phi \in C_{0}^{\infty }(\mathbb {R}^3\times (0,T))\) with \(\phi \ge 0,\)

    $$\begin{aligned}&\int _{\mathbb {R}^3\times \{t\}}(|u|^2+|\nabla d|^2)\phi \mathrm{d}x +2\int _{0}^t\int _{\mathbb {R}^3} (|\nabla u|^2+|\nabla ^2 d|^2)\phi \\&\quad \le \int _{0}^t\int _{\mathbb {R}^3} (|u|^2+|\nabla d|^2)(\phi _t+\Delta \phi )\\&\qquad +\int _0^t\int _{\mathbb {R}^3} (|u|^2+|\nabla d|^2+2p)(u\cdot \nabla \phi )\\&\qquad +2\int _0^t\int _{\mathbb {R}^3} ((u\cdot \nabla ) d\odot \nabla d)\cdot \nabla \phi -2\int _0^t\int _{\mathbb {R}^3}\nabla _x f(d)\cdot \nabla \mathrm{d} \phi . \end{aligned}$$

Denote for \(z=(x,t)\in \mathbb {R}^{3}\times \mathbb {R}_{+}\) the standard notations

$$\begin{aligned} B_{r}(x)=\{y\in \mathbb {R}^{3}:|y-x|<r\}, \quad Q_{r}(z)=B_{r}(x)\times (t-r^{2},t), \end{aligned}$$

to be the Euclidean ball and parabolic cylinder. For \(z=(0,0)\), we simply write them as \(B_{r}\) and \(Q_{r}\).

Now, we mention some relevant results on regularity. For 3D Navier–Stokes system, the study of partial regularity was originated by Scheffer in a series of papers [31, 32] and [33]. The notion of suitable weak solutions was introduced in a celebrated paper [1] by Caffarelli, Kohn and Nirenberg. It is proved that, for any suitable weak solution (up),  there is an open subset in which the velocity field u is Hölder continuous, and the complement of it has zero 1-D Hausdorff measure. Latter Lin [17] gave a simpler proof for the CKN theorem, Lin’s method was used extensively by Seregin’s papers for example in [3, 12]. These results are based on some regularity criterions when certain dimensionless quantities are small. For liquid crystal system (1.1), Lin and Liu[20] proved the following theorem.

Theorem(A). Let the triple (udp) be a suitable weak solution to system (1.1). There exist a small constant \(\epsilon _{0}>0\), such that if

$$\begin{aligned} \int _{Q_{1}}|u|^{3}+|\nabla d|^{3}+|p|^{\frac{3}{2}}<\epsilon _{0}, \end{aligned}$$

then u and d are smooth in \(\overline{Q_{\frac{1}{2}}}\). In particular, for any \(z_0\) if

$$\begin{aligned} \frac{1}{r^2}\int _{Q_r(z_0)}|u|^3+|\nabla d|^3<\epsilon _0, \quad \text {for all}\,\, 0<r\le 1, \end{aligned}$$

then \(z_0\) is a regular point.

We can drop the pressure p by Wolf’ method of local suitable weak solutions, the proof to see Sect. 5.

Proposition 1.2

Let the triple (udp) be a suitable weak solution to system (1.1). There exist a small constant \(\varepsilon _{0}>0\), such that if

$$\begin{aligned} \int _{Q_{1}}|u|^{3}+|\nabla d|^{3}<\varepsilon _{0}, \end{aligned}$$

then u and d are smooth in \(\overline{Q_{\frac{1}{2}}}\).

There is another type of regularity criterion called the Ladyzhenskaya–Prodi–Serrin condition (to see [13, 30, 36]). For system (1.1), [24] showed that if

$$\begin{aligned} ||u||_{L^{p,q}(Q)}=\left(\int ^{0}_{-1}\left(\int _{B}|u|^{p}\right)^{\frac{q}{p}}\right)^{\frac{1}{q}}<\infty , \end{aligned}$$

where \(\frac{3}{p}+\frac{2}{q}=1\) and \((p,q)\ne (3,\infty )\), then the weak solution (ud) is regular in \(Q=B\times (-1,0)\), Serrin’s method [36] and Struwe’s method [38] dealing with Navier–Stokes equations are applied. For the Borderline case where \((p,q)=(3,\infty )\), [28] showed that the weak solution (ud) is smooth in \(\mathbb {R}^{3}\times (0,T]\) when \(u\in L^{\infty }(0,T;L^{3}(\mathbb {R}^{3}))\)(also to see [25]). The \((3,\infty )\) case requires a technique utilizing the backward uniqueness of heat operator and unique continuation through spatial boundary, which was used to deal with the Navier–Stokes equations in [3]. For general Lorenz space \(L^{3,q}(\mathbb {R}^3), 3<q<\infty \) case, we prove in [26] if \(u\in L^\infty (0,T; L^{3,q}(\mathbb {R}^3))\) then (ud) is smooth. For N-S equations we refer to [29]. For \(L^{3,\infty }(\mathbb {R}^3)\) case, Choe, Wolf and Yang [2] prove that if \(u\in L^\infty (0,T; L^{3,\infty }(\mathbb {R}^3))\) and an additional local small condition

$$\begin{aligned} \frac{1}{r^3}|\{x\in B_r(x_0)|\,|u(\cdot ,t_0)|>\frac{\epsilon }{r}\}|\le \epsilon , \end{aligned}$$
(1.3)

then weak solution u of Navier–Stokes equations is smooth on \(Q_{\epsilon r}(z_0)\). Also to see Seregin [35].

In this paper, we shall established the regularity of weak solutions of liquid crystals equations in Lorentz space \(L^{3,\infty }\), the technicality is different from Navier–Stokes equations, the main technique is to deal with the new norms. Our main results can be stated as following.

Theorem 1.3

Assume (udp) is a weak solution of (1.1) with \(u\in L^\infty (0,T; L^{3,\infty }(\mathbb {R}^n))\). There exists a positive constant \(\varepsilon \) small with following property. If \(z_0=(x_0,t_0)\in Q_T\) and \(R>0\) such that \(Q_R(z_0)\subset Q_T\) and (dp) satisfy

$$\begin{aligned} C(\nabla d,R, z_0)+D(p, R, z_0)\le N, \end{aligned}$$

for some \(0< r\le R/2,\)

$$\begin{aligned} r^{-3}\left| \left\{ x\in B_r(x_0): |u(x,t_0)|>\varepsilon r^{-1}\right\} \right| \le \varepsilon . \end{aligned}$$

Then (ud) is smooth in \(Q_{\varepsilon r}(z_0).\) Here \(C(\nabla d,R,z_0)\) and \(D(p, R,z_0)\) are dimensionless quantities in Sect. 2.

Theorem 1.4

Let (ud) be a suitable weak solution to the liquid crystal equations in \(Q_T\) with \(u\in L^\infty (0,T; L^{3,\infty }(\mathbb {R}^n))\). Then there exist at most finite number \(\mathcal {N}\) of singular points at any singular time t.

2 Notations and preliminaries

Let us recall the scaling property of (1.1). Denote

$$\begin{aligned} u_{\lambda }(x,t)=\lambda u(\lambda x,\lambda ^{2}t),\quad p_{\lambda }(x,t)=\lambda ^{2} p(\lambda x,\lambda ^{2}t),\quad d_{\lambda }(x,t)=d(\lambda x,\lambda ^{2}t). \end{aligned}$$

If (upd) is a solution in \(\mathbb {R}^{3}\times (0,T)\), then obviously \((u_{\lambda },p_{\lambda },d_{\lambda })\) is a solution to the following equations

$$\begin{aligned} {\left\{ \begin{array}{ll} u_{\lambda t}-\Delta u_{\lambda }+u_{\lambda }\cdot \nabla u_{\lambda }+\nabla p_{\lambda }=-\nabla \cdot (\nabla d_{\lambda }\odot \nabla d_{\lambda }) &{}\mathrm{in}\quad \mathbb {R}^{3}\times (0,\lambda ^{2}T), \\ \nabla \cdot u_{\lambda }=0 &{}\mathrm{in}\quad \mathbb {R}^{3}\times (0,\lambda ^{2}T), \\ d_{\lambda t}-\Delta d_{\lambda }+u_{\lambda }\cdot \nabla d_{\lambda }=-{\lambda }^{2}f(d_{\lambda }) &{}\mathrm{in}\quad \mathbb {R}^{3}\times (0,\lambda ^{2}T). \end{array}\right. } \end{aligned}$$

Thus, the scaling dimension of corresponding quantities are \(\dim u=-1\), \(\dim p=-2\), and \(\dim d=0\) (we assign x with dimension 1 and t with 2). There are some useful dimensionless quantities and we list them here, let \(z_0=(x_0,t_0),\)

$$\begin{aligned} A(u,r,z_0)=\, & {} \sup _{t_0-r^{2}<t<t_0}r^{-1}\int _{B_{r}(x_0)}|u|^{2}\mathrm{d}x, \quad E(u,r, z_0)=r^{-1}\int _{Q_{r}(z_0)}|\nabla u|^{2}\mathrm{d}x\mathrm{d}t,\\ C(u,r, z_0)=\, & {} r^{-\frac{16}{7}}\int _{t_0-r^2}^{t_0}\Vert u\Vert ^4_{L^{\frac{14}{5}}(B_r(x_0))}\mathrm{d}t, \quad K(u,r, z_0)=r^{-3}\int _{Q_{r}(z_0)}|u|^{2}\mathrm{d}x\mathrm{d}t,\\ C_1(u,r, z_0)=\, & {} r^{-2}\int _{Q_{r}(z_0)}| u|^{3}\mathrm{d}x\mathrm{d}t,\quad D_1(p, r, z_0)=r^{-2}\int _{Q_{r}(z_0)}|p|^{\frac{3}{2}}\mathrm{d}x\mathrm{d}t,\\ D(p,z_0,r)=\, & {} r^{-\frac{16}{7}}\int _{t_0-r^2}^{t_0}\Vert p\Vert ^{2}_{L^{\frac{7}{5}}(B_r(x_0))}\mathrm{d}t. \end{aligned}$$

Similarly, we denote these notations for \(\nabla d\):

$$\begin{aligned} A(\nabla d,r, z_0)=\, & {} \sup _{t_0-r^{2}<t<t_0}r^{-1}\int _{B_{r}(x_0)}|\nabla d|^{2}\mathrm{d}x, \quad E(\nabla d,r, z_0)=r^{-1}\int _{Q_{r}(z_0)}|\nabla ^{2} d|^{2}\mathrm{d}x\mathrm{d}t,\\ C(\nabla d,r, z_0)=\, & {} r^{-\frac{16}{7}}\int _{t_0-r^2}^{t_0}\Vert \nabla d\Vert ^{4}_{L^{\frac{14}{5}}(B_r(x_0))}\mathrm{d}t, \quad K(\nabla d,r,z_0)=r^{-3}\int _{Q_{r}(z_0)}|\nabla d|^{2}\mathrm{d}x\mathrm{d}t,\\ C_1(\nabla d,r, z_0)=\, & {} r^{-2}\int _{Q_{r}(z_0)}| \nabla d|^{3}\mathrm{d}x\mathrm{d}t. \end{aligned}$$

For simplicity when \(z_0=(0,0)\), we write \(A(u,r)=A(u,r,(0,0))\) , and write \(A(r)\equiv A(u,\nabla d,r)=A(u,r)+A(\nabla d,r)\), and the meaning of E(r), C(r), K(r) are alike.

Next, we write down several facts about Lorentz spaces. We say a locally integrable function \(f\in L^{p,q}(\Omega )\), if the quasi-norm below is bounded

$$\begin{aligned} ||f||_{L^{p,q}(\Omega )}&=\left( p\int _{0}^{\infty }\alpha ^{q}d_{f,\Omega }(\alpha )^{\frac{q}{p}}\frac{d\alpha }{\alpha }\right) ^{\frac{1}{q}}, \quad q<\infty ,\nonumber \\ ||f||_{L^{p,\infty }(\Omega )}&=\sup _{\alpha >0}\alpha d_{f,\Omega }(\alpha )^{\frac{1}{p}}, \quad q=\infty , \end{aligned}$$
(2.1)

where

$$\begin{aligned} d_{f,\Omega }(\alpha )=|\{x\in \Omega :|f(x)|>\alpha \}|. \end{aligned}$$

A basic fact for such spaces is

$$\begin{aligned} L^{p,q_{1}}\subset L^{p,p}=L^{p}\subset L^{p,q_{2}}\subset L^{p,\infty }=L^{p}_{w}, \end{aligned}$$
(2.2)

where \(0<q_{1}<p<q_{2}<\infty \), and \(L^{p}_{w}\) is the weak-\(L^{p}\) space. If \(|\Omega |\) is finite then \(L^{p,q}(\Omega )\subset L^r(\Omega )\) for all \(0<q\le \infty \) and \(0<r<p\),

$$\begin{aligned} \Vert g\Vert _{L^r(\Omega )}\le |\Omega |^{\frac{1}{r}-\frac{1}{p}}\Vert g\Vert _{L^{p,q}(\Omega )}. \end{aligned}$$
(2.3)

Lemma 2.1

Let (udp) be a weak solution to the liquid crystal Eq. (1.1) in \(Q=\Omega \times (a,b)\). Let \(z_0=(x_0,t_0)\) and let \(\rho >0\) be such that \(Q_\rho (z_0)\subset Q.\) For every \(r<(0,\frac{\rho }{4}]\), we have

$$\begin{aligned} D_1(p,r,z_0)\le &\, {} c\left( \frac{\rho }{r}\right) ^\frac{3}{2}\left[ A(u,\rho , z_0)^\frac{3}{4}E(u,\rho ,z_0)^\frac{3}{4}+A(\nabla d, \rho , z_0)^\frac{3}{4}E(\nabla d, \rho , z_0)^\frac{3}{4}\right] \nonumber \\&\,+c\frac{r}{\rho }D(p,z_0,\rho )^\frac{3}{4}, \end{aligned}$$
(2.4)
$$\begin{aligned} D(p, r, z_0)\le\, & {} c\,\left[ \left( \frac{r}{\rho }\right) ^2D(p,\rho , z_0)+\left( \frac{\rho }{r}\right) ^{\frac{16}{7}} C(u,\nabla d, \rho , z_0)\right] . \end{aligned}$$
(2.5)

Proof

Let \(z_0=(0,0)\), we decompose p so that

$$\begin{aligned} p=p_1+p_2, \end{aligned}$$

where \(p_1\) satisfies in \(B_\rho \) for a.e. \(t\in [-\rho ^{2},0]\), in the weak sense,

$$\begin{aligned} {\left\{ \begin{array}{ll} \Delta p_1=-\mathrm {div\,div}\,(u\otimes u-[u\otimes u]_{B_\rho }))-\mathrm {div\,div}\,(\nabla d \otimes \nabla d -[\nabla d\otimes \nabla d]_{B_\rho })\\ p_1|_{\partial B_\rho }=0. \end{array}\right. } \end{aligned}$$
(2.6)

And \(p_2\) is a harmonic function in \(B_\rho \), i.e.,

$$\begin{aligned} \Delta p_2=0. \end{aligned}$$

Regarding \(p_1\), by theory of Laplace operator and Calderón–Zygmund theorem, we have

$$\begin{aligned}&\left( \int _{B_\rho }|p_1|^{\frac{3}{2}}dx\right) ^{\frac{2}{3}}\\&\quad \le c\left( \int _{B_\rho }|u\otimes u-[u\otimes u]_{B_\rho }|^{\frac{3}{2}}+|\nabla d\otimes \nabla d-[\nabla d\otimes \nabla d]_{B_\rho }|^{\frac{3}{2}})dx\right) ^{\frac{2}{3}}\\&\quad \le c\int _{B_\rho }|\nabla u||u|+|\nabla ^2 d||\nabla d|\\&\quad \le c\left[ \Vert \nabla u\Vert _{L^2(B_\rho )}\Vert u\Vert _{L^2(B_\rho )}+\Vert \nabla ^2 d\Vert _{L^2(B_\rho )}\Vert \nabla d\Vert _{L^2(B_\rho )}\right] ,\\&\int _{Q_\rho }|p_1|^{\frac{3}{2}}\le c\rho ^2\left[ A^{\frac{3}{4}}(u, \rho )E^{\frac{3}{4}}(u,\rho )+ A^{\frac{3}{4}}(\nabla d, \rho )E^{\frac{3}{4}}(\nabla d,\rho )\right] . \end{aligned}$$

For \(x\in B_{\frac{\rho }{2}}\),

$$\begin{aligned} |p_2(x,t)|\le c -\int _{B_\rho }|p_2|\le c \left( -\int _{B_\rho }|p_2|^l\right) ^{\frac{1}{l}}, \quad l>1, \end{aligned}$$

i.e., for \(r\le \frac{\rho }{2},\)

$$\begin{aligned} \frac{1}{r^2}\int _{Q_r}|p_2|^{\frac{3}{2}}\le\, & {} \frac{c\, r}{\rho ^{\frac{45}{14}}}\int _{-\rho ^2}^0\Vert p_2\Vert _{L^{\frac{7}{5}}(B_\rho )}^{3/2}.\\ D_1(p, r)\le &\,\, {} \frac{1}{r^2}\int _{Q_\rho }|p_1|^{\frac{3}{2}}+\frac{1}{r^2}\int _{Q_r}|p_2|^{\frac{3}{2}}\\\le\, & {} c(\frac{\rho }{r})^2\left[ A^{\frac{3}{4}}(u, \rho )E^{\frac{3}{4}}(u,\rho )+ A^{\frac{3}{4}}(B, \rho )E^{\frac{3}{4}}(\nabla d,\rho )\right] \\&+\,\frac{c\, r}{\rho ^{\frac{45}{14}}}\int _{-\rho ^2}^0\left( \Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}^{3/2}+ \Vert p_1\Vert _{L^{\frac{7}{5}}(B_\rho )}^{3/2}\right) . \end{aligned}$$

Now,

$$\begin{aligned}&\frac{r}{\rho ^{\frac{45}{14}}}\int _{-\rho ^2}^0\Vert p_1\Vert _{L^{\frac{7}{5}}(B_\rho )}^{3/2}\le \frac{r}{\rho }D_1(p_1, \rho ),\\&\frac{ r}{\rho ^{\frac{45}{14}}}\int _{-\rho ^2}^0\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}^{3/2}\le \frac{r}{\rho }D(p,\rho )^{\frac{3}{4}}, \end{aligned}$$

so that

$$\begin{aligned} D_1(p, r)\le c\left(\frac{\rho }{r}\right)^2\left[ A^{\frac{3}{4}}(u, \rho )E^{\frac{3}{4}}(u,\rho )+ A^{\frac{3}{4}}(\nabla d, \rho )E^{\frac{3}{4}}(\nabla d,\rho )\right] +c\frac{r}{\rho }D(p,\rho )^{\frac{3}{4}}. \end{aligned}$$

On the other hand

$$\begin{aligned} \int _{B_r}|p_1|^{\frac{7}{5}}\le c\int _{B_r}|u|^{\frac{14}{5}}+|\nabla d|^{\frac{14}{5}}, \end{aligned}$$

we have

$$\begin{aligned} D(p_1, r)\le c\,\left( \frac{\rho }{r}\right) ^{\frac{16}{7}} C(u, \nabla d, \rho ). \end{aligned}$$

For \(x\in B_{\frac{\rho }{2}}\),

$$\begin{aligned} |p_2(x,t)|\le c\left( -\int _{B_\rho }|p_2|^{\frac{7}{5}}dx\right) ^{\frac{5}{7}}, \end{aligned}$$

we have

$$\begin{aligned} D(p_2,r)=\, & {} r^{-\frac{16}{7}}\int _{-r^2}^0\Vert p_2\Vert _{L^{\frac{7}{5}}(B_r)}^2\le c\left( \frac{r}{\rho }\right) ^2D(p_2,\rho )\\\le &\, {} c\left( \frac{r}{\rho }\right) ^2\left[ D(p,\rho )+D(p_1,\rho )\right] , \end{aligned}$$

therefore,

$$\begin{aligned} D(p, r)\le\, & {} c\left[ D(p_1, r)+D(p_2,r)\right] \\\le\, & {} c\,\left( \frac{\rho }{r}\right) ^{\frac{16}{7}} C(u,\nabla d, \rho )+ c\,D(p_2,r)\\\le \,& {} c\,\left[ \left( \frac{r}{\rho }\right) ^2D(p, \rho )+\left( \frac{\rho }{r}\right) ^{\frac{16}{7}} C(u,\nabla d, \rho )\right] . \end{aligned}$$

\(\square \)

We use following analysis Lemma 2.2 which can be found in ([6],  Lemma 6.1) to prove local estimate Lemma 2.3.

Lemma 2.2

Let I(s) be a bounded nonnegative function in the interval \([R_1,R_2].\) Suppose that for any s\(\rho \in [R_1,R_2]\) and \(s<\rho \), the following yields

$$\begin{aligned} I(s)\le \left[ a_1(\rho -s)^{-\alpha }+a_2(\rho -s)^{-\beta }+a_3(\rho -s)^{-\gamma }+ a_4\right] +\theta I(\rho ), \end{aligned}$$

with \(\alpha>\beta>\gamma >0\), \( a_i>0 , i=1,2,3, 4 \) and \(\theta \in [0,1)\). Then,

$$\begin{aligned} I(R_1)\le c(\alpha ,\beta ,\gamma )[a_1(R_2-R_1)^{-\alpha }+a_2(R_2-R_1)^{-\beta }+a_3(R_2-R_1)^{-\gamma }+a_4]. \end{aligned}$$

Lemma 2.3

Let (udp) be a suitable weak solution to the liquid crystal Eq. (1.1) in \(Q=\Omega \times (a,b).\) Assume that \(z_0=(x_0,t_0)\) and \(1\ge r>0\) with \(Q_r(z_0)\subset Q\). Then the following holds:

$$\begin{aligned}&A(u,\nabla d, r/2, z_0)+E(u,\nabla d, r/2, z_0)\nonumber \\&\quad \le c\left[ C(u, \nabla d, r, z_0)^{\frac{1}{2}}+C(u, r, z_0)^{\frac{1}{2}}C( \nabla d, r, z_0)^{\frac{1}{2}}+C(u,r,z_0)\right. \nonumber \\&\qquad +\left. D(p, r, z_0)^\frac{7}{10}C(u, r, z_0)^\frac{3}{20}\right] . \end{aligned}$$
(2.7)

Proof

Let \(r/2\le s<\rho \le r<1,\) and \(Q_r\subset Q_1\equiv Q.\) Choosing test function \(\phi (x,t)=\eta _1(x)\eta _2(t)\) with \(\eta _1\in C^\infty _0(B_\rho (x_0)),\) \(0\le \eta _1\le 1\) in \(\mathbb {R}^3,\) \(\eta _1\equiv 1\) on \(B_s(x_0),\) and \( |\nabla ^\alpha \eta _1|\le \frac{C}{(\rho -s)^{|\alpha |}},\) for all multi-index \(\alpha ,\) with \(|\alpha |\le 3.\) And \(\eta _2\in C^\infty _0(t_0-\rho ^2,t_0+\rho ^2),\) \(0\le \eta _2\le 1\) in \(\mathbb {R},\) \(\eta _2(t)\equiv 1\) for \(t\in [t_0-s^2,t_0+s^2],\) with \( |\eta _2'(t)|\le \frac{C}{\rho ^2-s^2}\le \frac{C}{r(\rho -s)}. \) From the local energy inequality we have

$$\begin{aligned}&\int _{\Omega }(|u|^2+|\nabla d|^2)\phi \mathrm{d}x+2\int _{a}^{t}\int _{\Omega }(|\nabla u|^2+|\nabla ^2 d|^2)\phi \mathrm{d}x\mathrm{d}s\nonumber \\&\quad \le c\int _{t_0-\rho ^2}^{t_0}\Vert |u|^2+|\nabla d|^2\Vert _{W^{-1,2}(B_\rho (x_0))}\Vert \nabla (\phi _t+\Delta \phi )\Vert _{L^2(B_\rho (x_0))}\mathrm{d}t \nonumber \\&\qquad +c\int _{t_0-\rho ^2}^{t_0}\Vert |u|^2\Vert _{W^{-1,2}(B_\rho (x_0))}\Vert \nabla (u\cdot \nabla \phi )\Vert _{L^2(B_\rho (x_0))}\mathrm{d}t\nonumber \\&\qquad +c\int _{t_0-\rho ^2}^{t_0}\Vert |u||\nabla d|\Vert _{W^{-1,2}(B_\rho (x_0))}\Vert |\nabla ^2d||\nabla \phi |+|\nabla d||\nabla ^2\phi |\Vert _{L^2(B_\rho (x_0))}\mathrm{d}t\nonumber \\&\qquad +c\int _{t_0-\rho ^2}^{t_0}\int _{B_\rho }\,|p\,u\cdot \nabla \phi | \mathrm{d}x\mathrm{d}t + c\int _{t_0-\rho ^2}^{t_0}\int _{B_\rho }\,|\nabla d|^2\phi \nonumber \\&\quad =J_1+J_2+J_3+J_4+J_5. \end{aligned}$$
(2.8)

Here, we rewrite the term

$$\begin{aligned} \int _{t_0-\rho ^2}^{t_0}\int _{B_\rho }|\nabla d|^2 \, u\cdot \nabla \phi \mathrm{d}x\mathrm{d}t= \int _{t_0-\rho ^2}^{t_0}\int _{B_\rho }(u\otimes \nabla d):(\nabla \phi \otimes \nabla d) \mathrm{d}x\mathrm{d}t. \end{aligned}$$

Denote

$$\begin{aligned} I(s)= & {} \sup _{t_0-s^2\le t\le t_0}\int _{B_s(x_0)}|\nabla d|^2\mathrm{d}x+\sup _{t_0-s^2\le t\le t_0}\int _{B_s(x_0)}|u|^2\mathrm{d}x\\&+\int _{t_0-s^2}^{t_0}\int _{B_s(x_0)}|\nabla ^2 d|^2\mathrm{d}x\mathrm{d}t+\int _{t_0-s^2}^{t_0}\int _{B_s(x_0)}|\nabla u|^2\mathrm{d}x\mathrm{d}t\\=\, & {} sA(\nabla d,s,z_0)+sA(u,s,z_0)+sE(\nabla d,s,z_0)+sE(u,s,z_0)\\=\, & {} I_1(\nabla d,s)+I_1(u,s)+I_2(\nabla d,s)+I_2(u,s), \end{aligned}$$

and

$$\begin{aligned} I(u,s)=I_1(u,s)+I_2(u,s), \quad I(\nabla d,s)=I_1(\nabla d,s)+I_2(\nabla d,s). \end{aligned}$$

Estimate \(J_1,\) \(J_2,\) \(J_3, J_4\) and \(J_5\), respectively, as the following:

$$\begin{aligned} J_1\le\, & {} \frac{c\rho ^{3/2}}{(\rho -s)^3}\int _{t_0-\rho ^2}^{t_0}\Vert |u|^2+|\nabla d|^2\Vert _{W^{-1,2}(B_\rho )}\mathrm{d}t\\\le\, & {} \frac{c\rho ^{\frac{5}{2}}}{(\rho -s)^3}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert |u|^2+|\nabla d|^2\Vert _{W^{-1,2}(B_\rho )}^2\mathrm{d}t\right] ^{\frac{1}{2}}, \end{aligned}$$

by Young’s inequality, we get

$$\begin{aligned} J_2\le \,& {} c\int _{t_0-\rho ^2}^{t_0}\left[ \Vert |u|^2\Vert _{W^{-1,2}(B_\rho )}\left( \frac{\Vert \nabla u\Vert _{L^2(B_\rho )}}{\rho -s}+\frac{\Vert u\Vert _{L^2(B_\rho )}}{(\rho -s)^2}\right) \right] \mathrm{d}t\\\le\, & {} \frac{c}{\rho -s}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert |u|^2\Vert ^2_{W^{-1,2}(B_\rho )}\mathrm{d}t\right] ^\frac{1}{2}I_2(v,\rho )^{\frac{1}{2}}\\&+\frac{c\rho }{(\rho -s)^2}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert |u|^2\Vert ^2_{W^{-1,2}(B_\rho )}\mathrm{d}t\right] ^\frac{1}{2}I_1(u,\rho )^{\frac{1}{2}}\\\le\, & {} \frac{1}{4}I(u,\rho )+\left[ \frac{c}{(\rho -s)^2}+\frac{c\rho ^2}{(\rho -s)^4}\right] \int _{t_0-\rho ^2}^{t_0}\Vert |u|^2\Vert ^2_{W^{-1,2}(B_\rho )}\mathrm{d}t, \end{aligned}$$

similarly, we have

$$\begin{aligned} J_3\le\, & {} c\int _{t_0-\rho ^2}^{t_0}\left[ \Vert |u||\nabla d|\Vert _{W^{-1,2}(B_\rho )}\left( \frac{\Vert \nabla ^2 d\Vert _{L^2(B_\rho )}}{\rho -s}+\frac{\Vert \nabla d\Vert _{L^2(B_\rho )}}{(\rho -s)^2}\right) \right] \mathrm{d}t\\\le\, & {} \frac{1}{4}I(\nabla d,\rho )+\left[ \frac{c}{(\rho -s)^2}+\frac{c\rho ^2}{(\rho -s)^4}\right] \int _{t_0-\rho ^2}^{t_0}\Vert |u\nabla d|\Vert ^2_{W^{-1,2}(B_\rho )}\mathrm{d}t. \end{aligned}$$

For the term \(J_4\), using Hölder’s inequality and Sobolev inequality, we have

$$\begin{aligned} J_4\le &\, {} c\int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho (x_0))}\Vert u\nabla \phi \Vert _{L^{\frac{7}{2}}(B_\rho (x_0))}\\\le\, & {} c\int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}\Vert \nabla (u\nabla \phi )\Vert _{L^2(B_\rho )}^{\frac{4}{7}}\Vert u\nabla \phi \Vert _{L^{\frac{9}{4}}(B_\rho )}^{\frac{3}{7}}\\\le\, & {} c\int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}\left[ \Vert \nabla u\nabla \phi )\Vert _{L^2(B_\rho )}+\Vert u\nabla ^2\phi \Vert _{L^2(B_\rho )}\right] ^{\frac{4}{7}}\Vert u\nabla \phi \Vert _{L^{\frac{9}{4}}(B_\rho )}^{\frac{3}{7}} \\\le\, & {} \frac{c}{\rho -s}\Vert \nabla u\Vert _{L^2(Q_\rho (z_0))}^{\frac{4}{7}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}^{\frac{7}{5}}\Vert u\Vert _{L^{\frac{9}{4}}(B_\rho )}^{\frac{3}{5}}\right] ^{\frac{5}{7}}\\&+\frac{c}{(\rho -s)^{\frac{11}{7}}}\sup _t\Vert u\Vert _{L^2(B_\rho )}^{\frac{4}{7}}\int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )} \Vert u\Vert _{L^{\frac{9}{4}}(B_\rho )}^{\frac{3}{7}}\\\le\, & {} \frac{1}{4}I(u,\rho )+ \frac{c}{(\rho -s)^{\frac{7}{5}}}\int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}^{\frac{7}{5}}\Vert u\Vert _{L^{\frac{9}{4}}(B_\rho )}^{\frac{3}{5}}\\&+ \frac{c}{(\rho -s)^{\frac{11}{5}}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )} \Vert u\Vert _{L^{\frac{9}{4}}(B_\rho )}^{\frac{3}{7}}\right] ^{\frac{7}{5}}\\\le\, & {} \frac{1}{4}I(u,\rho )+\frac{c\rho ^{\frac{16}{35}}}{(\rho -s)^{\frac{7}{5}}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}^{2}\right] ^{\frac{7}{10}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert u\Vert _{L^{\frac{14}{5}}(B_\rho )}^{4}\right] ^{\frac{3}{20}}\\&+ \frac{c\rho ^{\frac{44}{35}}}{(\rho -s)^{\frac{11}{5}}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_\rho )}^{2}\right] ^{\frac{7}{10}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert u\Vert _{L^{\frac{14}{5}}(B_\rho )}^{4}\right] ^{\frac{3}{20}}.\\ J_5\le\, & {} c\,\rho ^{\frac{13}{7}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert \nabla d\Vert _{L^{\frac{14}{5}}(B_\rho )}^4\right] ^{1/2}. \end{aligned}$$

From (2.8), using estimates above with respect to \(J_1, J_2, J_3, J_4\) and \(J_5\) we get

$$\begin{aligned} I(s)\le\, & {} \frac{1}{2}I(\rho )+\frac{c r^\frac{5}{2}}{(\rho -s)^3}\left[ \int _{t_0-r^2}^{t_0}\Vert |u|^2+|\nabla d|^2\Vert _{W^{-1,2}(B_r)}^2\right] ^{\frac{1}{2}}\nonumber \\&+\,\left[ \frac{c}{(\rho -s)^2}+\frac{c r^2}{(\rho -s)^4}\right] \int _{t_0-r^2}^{t_0}\left[ \Vert |u\,\nabla d|\Vert ^2_{W^{-1,2}(B_r)}+\Vert |u|^2\Vert ^2_{W^{-1,2}(B_r)}\right] \nonumber \\&+\,\frac{c r^{\frac{16}{35}}}{(\rho -s)^{\frac{7}{5}}}\left[ \int _{t_0-r^2}^{t_0} \Vert p\Vert _{L^{\frac{7}{5}}(B_r)}^{2}\right] ^{\frac{7}{10}}\left[ \int _{t_0-r^2}^{t_0}\Vert u\Vert _{L^{\frac{14}{5}}(B_r)}^{4}\right] ^{\frac{3}{20}}\nonumber \\&+\, \frac{c r^{\frac{44}{35}}}{(\rho -s)^{\frac{11}{5}}}\left[ \int _{t_0-r^2}^{t_0}\Vert p\Vert _{L^{\frac{7}{5}}(B_r)}^{2}\right] ^{\frac{7}{10}}\left[ \int _{t_0-r^2}^{t_0}\Vert u\Vert _{L^{\frac{14}{5}}(B_r)}^{4}\right] ^{\frac{3}{20}}\nonumber \\&+\,c\,\rho ^{\frac{13}{7}}\left[ \int _{t_0-\rho ^2}^{t_0}\Vert \nabla d\Vert _{L^{\frac{14}{5}}(B_\rho )}^4\right] ^{1/2}. \end{aligned}$$
(2.9)

By Lemma 2.2, we have

$$\begin{aligned} I(r/2)\le\, & {} cr^{-\frac{1}{2}}\left[ \int _{t_0-r^2}^{t_0}\Vert |u|^2+|\nabla d|^2\Vert _{W^{-1,2}(B_r)}^2\right] ^{\frac{1}{2}}\\&+cr^{-2}\int _{t_0-r^2}^{t_0}\left( \Vert |u\,\nabla d|\Vert ^2_{W^{-1,2}(B_r)}+\Vert |u|^2\Vert ^2_{W^{-1,2}(B_r)}\right) \mathrm{d}t\\&+\,c r^{-\frac{33}{35}}\left[ \int _{t_0-r^2}^{t_0} \Vert p\Vert _{L^{\frac{7}{5}}(B_r)}^{2}\right] ^{\frac{7}{10}}\left[ \int _{t_0-r^2}^{t_0}\Vert u\Vert _{L^{\frac{14}{5}}(B_r)}^{4}\right] ^{\frac{3}{20}}\\&+\, c\,r^{\frac{13}{7}}\left[ \int _{t_0-r^2}^{t_0}\Vert \nabla d\Vert _{L^{\frac{14}{5}}(B_r)}^4\right] ^{1/2}. \end{aligned}$$

Finally, we get

$$\begin{aligned}&A(u,\nabla d,r/2,z_0)+E(u,\nabla d,r/2,z_0)\nonumber \\&\quad \le c r^{-\frac{3}{2}}\left[ \int _{t_0-r^2}^{t_0}\Vert |u|^2+|\nabla d|^2\Vert _{W^{-1,2}(B_r)}^2\right] ^{1/2}\nonumber \\&\qquad + c r^{-3}\int _{t_0-r^2}^{t_0}\left( \Vert |u\,\nabla d|\Vert ^2_{W^{-1,2}(B_r)}+\Vert |u|^2\Vert ^2_{W^{-1,2}(B_r)}\right) \nonumber \\&\qquad + cr^{-\frac{68}{35}}\left[ \int _{t_0-r^2}^{t_0} \Vert p\Vert _{L^{\frac{7}{5}}(B_r)}^{2}\right] ^{\frac{7}{10}}\left[ \int _{t_0-r^2}^{t_0}\Vert u\Vert _{L^{\frac{14}{5}}(B_r)}^{4}\right] ^{\frac{3}{20}}\nonumber \\&\qquad + c\,r^{\frac{6}{7}}\left[ \int _{t_0-r^2}^{t_0}\Vert \nabla d\Vert _{L^{\frac{14}{5}}(B_r)}^4\right] ^{1/2}. \end{aligned}$$
(2.10)

For \(f\in L^{7/5}(B_r(x_0))\) and \(\varphi \in C^\infty _0(B_r(x_0))\), we have

$$\begin{aligned} \left| \int _{B_r(x_0)}\varphi f(x)dx\right|&\le c\int _{B_r(x_0)}\left[ \int _{B_r(x_0)}\frac{|\nabla \varphi (y)|}{|x-y|^2}\mathrm{d}y\right] |f(x)|\mathrm{d}x\\&= c\int _{B_r(x_0)}|\nabla \varphi (y)|\left[ \int _{B_r(x_0)}\frac{|f(x)|}{|x-y|^2}\mathrm{d}x\right] \mathrm{d}y\\&\le c\Vert \nabla \varphi \Vert _{L^2(B_r(x_0))}\Vert \mathbf {I}_1(\chi _{B_r(x_0)}|f|)\Vert _{L^2(B_r(x_0))}. \end{aligned}$$

where \(\mathbf {I}_1\) is the first order Riesz’s potential defined by

$$\begin{aligned} \mathbf {I}_1(\mu )(x)=c\int _{\mathbb {R}^3}\frac{d\mu (y)}{|x-y|^2},\quad x\in \mathbb {R}^3. \end{aligned}$$

By using Hardy–Littlewood–Sobolev inequality, we have

$$\begin{aligned} \Vert f\Vert _{W^{-1,2}(B_r(x_0))}\le & {} \Vert \mathbf {I}_1(\chi _{B_r(x_0)}|f|)\Vert _{L^2(B_r(x_0))}\nonumber \\\le & {} c\Vert f\Vert _{L^{\frac{6}{5}}(B_r(x_0))}\le cr^{\frac{5}{14}}\Vert f\Vert _{L^{\frac{7}{5}}(B_r(x_0))}. \end{aligned}$$
(2.11)

Applying (2.11) with \(f=|y|^2+|\nabla d|^2,\) and \(f= |u||\nabla d|,\) we obtain

$$\begin{aligned}&r^{-3}\int _{t_0-r^2}^{t_0}\Vert |u|^2+|\nabla d|^2\Vert _{W^{-1,2}(B_r)}^2\\&\quad \le r^{\frac{-16}{7}}\int _{t_0-r^2}^{t_0}\left( \Vert u\Vert _{L^{\frac{14}{5}}(B_r)}^4 +\Vert \nabla d\Vert _{L^{\frac{14}{5}}(B_r)}^4\right) \\&\quad = C(u,\nabla d, r, z_0),\\&r^{-3}\int _{t_0-r^2}^{t_0}\left( \Vert |u\,\nabla d|\Vert ^2_{W^{-1,2}(B_r)}+\Vert |u|^2\Vert ^2_{W^{-1,2}(B_r)}\right) \\&\quad \le \left[ C(u,r, z_0)^{\frac{1}{2}} C(\nabla d, r, z_0)^{\frac{1}{2}}+C(u, r, z_0)\right] . \end{aligned}$$

\(\square \)

We need the bounded estimates for \(C(\nabla d,r)\) and D(pr) with the help of the bounded of C(ur).

Lemma 2.4

Suppose that (udp) is a suitable weak solution in \(Q_1(z_0)=B_1(x_0)\times (t_0-1,t_0)\). Let

$$\begin{aligned} C(u,r, z_0)\le M\quad \text {for any }0<r\le 1, \end{aligned}$$

for \(M>0\). Then for every \(0<r<1/4,\) we have the following estimates:

$$\begin{aligned}&A(u,\nabla d, r/2, z_0)+E(u,\nabla d,r/2, z_0)+C(\nabla d,r/2, z_0)+D(p,r/2,z_0)\\&\quad \le c(M,C(\nabla d, 1/2, z_0 ), D(p,1/2, z_0));\\&C_1(u,\nabla d, r, z_0)+D_1(p, r, z_0)\le c(M, D(p,1/2, z_0), C(\nabla d, 1/2, z_0)). \end{aligned}$$

Proof

Without loss of generality, we consider \(z_0=(0,0).\) It is easy to see that

$$\begin{aligned} C(\nabla d,r/2)\le c\left[ E(\nabla d,r)^\frac{5}{7}C(\nabla d,r)^\frac{2}{7}+C(\nabla d,r)^\frac{9}{14}\right] . \end{aligned}$$
(2.12)

Combining with (2.7) and (2.12) we obtain

$$\begin{aligned} C(\nabla d,r/2)\le \,& {} c(M)\,\left[ 1+ C(\nabla d,r)^{\frac{1}{2}}+D(p,r)^{\frac{7}{10}}\right] ^{\frac{5}{7}}C(\nabla d,r)^{\frac{2}{7}} \nonumber \\&+c\,C(\nabla d,r)^{\frac{9}{14}}. \end{aligned}$$
(2.13)

Let \(r=\theta \rho \) with \(\theta \le \frac{1}{4}.\) From (2.5) (2.13), we have

$$\begin{aligned}&C(\nabla d,r)+D(p,r)^\frac{5}{6}\nonumber \\&\quad \le c(M)\theta ^{-\frac{32}{49}}\,\left[ 1+\theta ^{-\frac{8}{7}} C(\nabla d,\rho )^{\frac{1}{2}}+\theta ^{-\frac{8}{5}}D(p,\rho )^{\frac{7}{10}}\right] ^{\frac{5}{7}}C(\nabla d,\rho )^{\frac{2}{7}}\nonumber \\&\qquad +c\theta ^{-\frac{72}{49}}\,C(\nabla d,\rho )^{\frac{9}{14}} +c\theta ^\frac{5}{3}D(p,\rho )^\frac{5}{6}+c\theta ^{-\frac{40}{21}}C(\nabla d,\rho )^\frac{5}{6}+c(M,\theta ) \end{aligned}$$

Using Young’s inequality, we have

$$\begin{aligned} C(\nabla d,r)+D(p,r)^\frac{5}{6}\le \eta C(\nabla d,\rho )+(\eta +c\theta ^\frac{5}{3})D(p,\rho )^\frac{5}{6}+c(\theta ,\eta ,M) \end{aligned}$$

Set \(F(r)=C(\nabla d,r)+D(p,r)^\frac{5}{6}.\) Choose \(\eta >0\) and \(\theta >0\) small enough, we have

$$\begin{aligned} F(r)\le \frac{1}{2}F(\rho )+c. \end{aligned}$$

By the standard iterating argument

$$\begin{aligned} C(\nabla d,r)+D(p,r)^\frac{5}{6}\le c\left( M, D(p,1/2), C(\nabla d,1/2)\right) , \quad r\in (0,\frac{1}{4}]. \end{aligned}$$

So that for \(0<r\le 1/4,\)

$$\begin{aligned} A(u,\nabla d,r)+E(u,\nabla d,r)+C(\nabla d,r)+D(p,r)\le c(M,C(\nabla d,1/2),D(p,1/2)). \end{aligned}$$

The estimates of \(C_1(u,\nabla d,r)\) and \(D_1(p,r)\) are immediate results. \(\square \)

3 Proof of Theorem 1.3

We shall prove the following Proposition 3.1, Theorem 1.3 is an immediate result.

Proposition 3.1

Let (udp) a weak solution of (1.1) with

$$\begin{aligned} \Vert u\Vert _{L^\infty (0, T; L^{3,\infty }(\mathbb {R}^3))}\le M, \end{aligned}$$

for \(z_0=(x_0,t_0)\) and \(R>0\) such that \(Q_R(z_0)\subset Q_T\), (dp) satisfy

$$\begin{aligned} C(\nabla d,R,z_0)+D(p, R, z_0)\le N. \end{aligned}$$

there exists a positive number \(\varepsilon (M, N)<\frac{1}{4}\) such that if for some \(0<r\le R/2\),

$$\begin{aligned} r^{-3}\left| \left\{ x\in B_r(x_0): |u(x,t_0)|>\varepsilon r^{-1}\right\} \right| \le \varepsilon , \end{aligned}$$
(3.1)

then there exists \(\rho \in [2r\varepsilon , r]\) such that

$$\begin{aligned} \frac{1}{\rho ^2}\int _{Q_\rho (z_0)}|u|^3+|\nabla d|^3<\varepsilon _0 \end{aligned}$$
(3.2)

where \(\varepsilon _0\) is the same number in Proposition 1.2.

Proof

Let (udp) be a weak solution of (1.1) with \(u\in L^\infty (0,T; L^{3,\infty }(\mathbb {R}^3)\). We assume

$$\begin{aligned} \sup _{0<r<T}\Vert u(t)\Vert _{L^{3,\infty }(\mathbb {R}^3)}\le M. \end{aligned}$$
(3.3)

Note that \(|d|\le 1\), we have (see [28])

$$\begin{aligned} ||\nabla d||_{L^{4}_{x}}\le c||\nabla d||^{\frac{1}{2}}_{\dot{B}^{-1}_{\infty ,\infty }}||\nabla d||^{\frac{1}{2}}_{\dot{H}^{1}}\le c\Vert d\Vert _{L^\infty }^{1/2}||\nabla d||^{\frac{1}{2}}_{\dot{H}^{1}}. \end{aligned}$$

Also since the real interpolation \(L^{4}=[L^{6,\infty },L^{3,\infty }]_{\frac{1}{2},4}\) holds, then

$$\begin{aligned} ||u||_{L^{4}_{x}}\le c||u||^{\frac{1}{2}}_{L^{6,\infty }_{x}}||u||^{\frac{1}{2}}_{L^{3,\infty }_{x}}\le c||\nabla u||^{\frac{1}{2}}_{L^{2}_{x}}M^{\frac{1}{2}}. \end{aligned}$$

From energy inequality and estimates above we get

$$\begin{aligned} \Vert (u, \nabla d)\Vert _{L^4(Q_T)}\le c(M, c_0), \end{aligned}$$
(3.4)

which yields (udp) is a local suitable weak solution of (1.1) and \(u\in C([0,T]; L^2(\mathbb {R}^3))\).

We use a contradiction argument for \(z_0=(0,0)\) and \(R=1\). Fixed \(N, M>0\) if the assertion of the proposition were false, then there would exist \(\epsilon _k \downarrow 0\), and suitable weak solutions \((u_k,d_k, p_k)\) of (1.1) and \(r_k\le 1/2\) such that

$$\begin{aligned}&\Vert u_k\Vert _{L^\infty (-1,0; L^{3,\infty }(\mathbb {R}^3))}\le M; \end{aligned}$$
(3.5)
$$\begin{aligned}&C(\nabla d_k, 1)+D(p_k, 1)\le N;\end{aligned}$$
(3.6)
$$\begin{aligned}&r_k^{-3}\left| \left\{ x\in B_{r_k}(0): |u_k(x,0)|>\epsilon _k r_k^{-1}\right\} \right| \le \epsilon _k; \end{aligned}$$
(3.7)

and for all \(\rho \in [2r_k\epsilon _k, r_k]\),

$$\begin{aligned} \frac{1}{\rho ^2}\int _{Q_\rho (0)}|u_k|^3+|\nabla d_k|^3>\varepsilon _0/2. \end{aligned}$$
(3.8)

Since for \(0<r\le 1\)

$$\begin{aligned} C(u_k, r)= & {} r^{-16/7}\int _{-r^2}^0|B_{r}|^{2/21}\Vert u_k\Vert _{L^{3,\infty }(B_{r})}^4\\\le & {} c\, \Vert u_k\Vert _{L^\infty (-1,0; L^{3,\infty }(B_1))}^4\le M^4, \end{aligned}$$

combining the estimate and (3.6) with Lemma 2.4, we get, for \(0<r\le 1/2\),

$$\begin{aligned} A(u_k,\nabla d_k, r)+ E(u_k,\nabla d_k, r)+ C(\nabla d_k, r) + D(p_k, r)\le c(M,N). \end{aligned}$$

Similarly, for any \(z_0\in Q_{1/2}\) and \(0<r\le 1/2\), we have

$$\begin{aligned} A(u_k,\nabla d_k, r, z_0)+ E(u_k,\nabla d_k, r, z_0)+ C(\nabla d_k, r, z_0) + D(p_k, r, z_0)\le c(M,N). \end{aligned}$$

Define, for \((x,t)\in Q_{r_k^{-1}}\),

$$\begin{aligned} {\left\{ \begin{array}{ll} U_{k}(x,t) = r_{k}u_k(r_{k}x,r_{k}^{2}t), \\ D_{k}(x,t) = d_k(r_{k}x,r_{k}^{2}t), \\ P_{k}(x,t) = r_{k}^{2}p_k(r_{k}x,r_{k}^{2}t). \end{array}\right. } \end{aligned}$$
(3.9)

Obviously, \((U_k, D_k, P_k)\) are weak solutions to system (1.1) with the right side of (1.1)\(_{3}\) replaced by \(r_{k}^{2}f(D_{k})\) in \(Q_{r_k^{-1}}\). Now, for \(a>0,\) and \(a r_k\le 1/2,\)

$$\begin{aligned} \Vert U_k\Vert _{L^{\infty }(-r_k^{-1},0; L^{3,\infty }(B_{r_k^{-1}}))}=\, & {} \Vert u_k\Vert _{L^\infty (-1, 0; L^{3,\infty }(B_1))}\le M,\\ C(\nabla D_k, a)+D(P_k, a)=\, & {} C(\nabla d_k, ar_k)+D(p_k, ar_k)\le N,\\ C(U_k, a)= \,& {} C(u_k, ar_k)\\\le\, & {} (ar_k)^{-16/7}\int _{-(ar_k)^2}^0|B_{ar_k}|^{2/21}\Vert u_k\Vert _{L^{3,\infty }(B_{ar_k})}^4\\\le\, & {} c\, \Vert u_k\Vert _{L^\infty (-1,0; L^{3,\infty }(B_1))}^4\le M^4. \end{aligned}$$

We have by Lemma 2.4 again

$$\begin{aligned}&A(U_k,\nabla D_k, a)+E(U_k,\nabla D_k,a)+D(P_k,a)\nonumber \\&\quad +C_1(U_k,\nabla D_k, a)+D_1(P_k,a)\le c(M,N). \end{aligned}$$
(3.10)

So that

$$\begin{aligned} \Vert U_k\Vert _{L^4(Q_a)}^4\le\, & {} c\,\int _{-a^2}^0\Vert U_k\Vert _{L^{3,\infty }(B_a)}^2\Vert U_k\Vert _{L^6(B_a)}^{2}\nonumber \\\le\, & {} c\,\int _{-a^2}^0\Vert U_k\Vert _{L^{3,\infty }(B_{a})}^{2}\Vert \nabla U_k\Vert _{L^2(B_{a})}^{2}+c\,\int _{-a^2}^0|B_a|\Vert U_k\Vert _{L^{3,\infty }(B_{a})}^{2}\Vert U_k\Vert _{L^2(B_{a})}^2\nonumber \\\le\, & {} c\,a M^2\left( A(U_k, a)+E(U_k, a)\right) ,\\ \Vert \nabla D_k\Vert _{L^4(Q_a)}^4\le \,& {} a\, c(M,N)\left( A(\nabla D_k, a)+E(\nabla D_k, a)\right) . \end{aligned}$$

Thus, the \(L^p\) estimate holds for \((U_k, D_k, P_k)\) in \(Q_a\), for any \(a>0\),

$$\begin{aligned}&\int _{Q_{a}}|U_k|^4+ |\nabla D_k|^4+|\partial _{t}U_k|^{\frac{4}{3}}+ |\partial _{t}\nabla D_k|^{\frac{4}{3}}+|\nabla ^{2}U_k|^{\frac{4}{3}}+|\nabla ^{2}\nabla D_k|^{\frac{4}{3}} +|\nabla P_k|^{\frac{4}{3}}\nonumber \\&\quad \le c_2(a,M, N). \end{aligned}$$
(3.11)

By Aubin–Lion’s lemma, there exists a triplet (veq) such that

$$\begin{aligned} {\left\{ \begin{array}{ll} U_k\rightarrow v, \quad &{}\text {in }L^3(Q_a),\\ \nabla D_k\rightarrow \nabla e, \quad &{}\text {in }L^3(Q_a),\\ P_k\rightharpoonup q, \quad &{}\text {in }L^\frac{3}{2}(Q_a), \end{array}\right. } \end{aligned}$$
(3.12)

and

$$\begin{aligned} U_k\rightarrow v,\quad \nabla D_k\rightarrow \nabla e \quad \text {in }C([-a^2,0]; L^{\frac{4}{3}}(B_a)). \end{aligned}$$

Using estimates above, the limit function (veq) satisfy, in the sense of suitable weak solutions on \(\mathbb {R}^3\times (-\infty ,0)\),

$$\begin{aligned} {\left\{ \begin{array}{ll} v_{t}-\Delta v+v\cdot \nabla v+\nabla q=-\nabla \cdot (\nabla e\odot \nabla e), \\ \nabla \cdot v=0, \\ e_{t}-\Delta e+v\cdot \nabla e=0. \end{array}\right. } \end{aligned}$$
(3.13)

From (3.7) and (3.8), we get

$$\begin{aligned} \left| \left\{ x\in B(0): |U_k(x,0)|>\epsilon _k\right\} \right| <\epsilon _k, \end{aligned}$$
(3.14)

and for \(\rho \in [2\epsilon _k,1]\)

$$\begin{aligned} \frac{1}{\rho ^2}\int _{Q_{\rho }}|U_{k}|^{3}+|\nabla D_{k}|^{3}>\varepsilon _0/2. \end{aligned}$$
(3.15)

Taking limit we get

$$\begin{aligned} v(\cdot , 0)=0, \quad \text {in}\, B_1(0), \end{aligned}$$
(3.16)

and for \(\rho \in (0, 1]\)

$$\begin{aligned} \frac{1}{\rho ^2}\int _{Q_{\rho }}|v|^{3}+|\nabla e|^{3}\ge \varepsilon _0/2. \end{aligned}$$
(3.17)

The crucial point here is a reduction to backward uniqueness for the heat operator with lower order terms as [3]. Set

$$\begin{aligned} v_k=\rho _kv(\rho _kx,\rho _k^2t), \quad e_k=e(\rho _kx,\rho _k^2t),\quad q_k=\rho _k^2q(\rho _kx,\rho _k^2t). \end{aligned}$$

Then \((v_k, e_k, q_k)\) satisfy (3.13), and similar to (3.10) for any \(a>0\)

$$\begin{aligned} A(v_k,\nabla e_k, a)+ & {} E(v_k,\nabla e_k,a)+D(q_k,a)\nonumber \\+ & {} C_1(v_k,\nabla e_k, a)+D_1(q_k,a)\le c(M,N). \end{aligned}$$
(3.18)

As before there exists a triplet \((\widetilde{v},\widetilde{e},\widetilde{q})\) is suitable weak solution of (3.13) such that

$$\begin{aligned} {\left\{ \begin{array}{ll} v_k\rightarrow \widetilde{v}, \quad &{}\text {in }L^3(Q_a),\\ \nabla e_k\rightarrow \nabla \widetilde{e}, \quad &{}\text {in }L^3(Q_a),\\ q_k\rightharpoonup \widetilde{q}, \quad &{}\text {in }L^\frac{3}{2}(Q_a), \end{array}\right. } \end{aligned}$$
(3.19)

and

$$\begin{aligned} v_k\rightarrow \widetilde{v},\quad \nabla e_k\rightarrow \nabla \widetilde{e} \quad \text {in }C([-a^2,0]; L^{\frac{4}{3}}(B_a)). \end{aligned}$$

From (3.16) we get

$$\begin{aligned} \widetilde{v}(\cdot , 0)=0, \quad \text {in}\,\mathbb {R}^3, \end{aligned}$$
(3.20)

from (3.17) and take \(\rho =\rho _k\)

$$\begin{aligned} \int _{Q_{1}}|\widetilde{v}|^{3}+|\nabla \widetilde{e}|^{3}\ge \varepsilon _0/2. \end{aligned}$$
(3.21)

On the other hand, for fixed \(z_0\) and \(0<R\le \frac{1}{2r_k}\), as (3.18)

$$\begin{aligned}&A(v_k,\nabla e_k, R, z_0)+E(v_k,\nabla e_k,R, z_0)+D(q_k,R, z_0)\nonumber \\&\quad +C_1(v_k,\nabla e_k, R, z_0)+D_1(q_k,R, z_0)\le c(M,N). \end{aligned}$$
(3.22)

By the Fubini theorem, we have,

$$\begin{aligned}&\left| \left\{ (x,t)\in \mathbb {R}^3\times (-T,0): \, |\widetilde{v}(x,t)|>\gamma \right\} \right| \\&\quad =\int _{-T}^0d_{\widetilde{v}(t)}(\gamma )\mathrm{d}t\le \gamma ^{-3}M^3T. \end{aligned}$$

Hence, for any \(\eta >0,\) there exists a \(B_R\) such that

$$\begin{aligned} \left| \left\{ (x,t)\in (\mathbb {R}^3\backslash B_R)\times (-T,0): \, |\widetilde{v}(x,t)|>\gamma \right\} \right| <\eta . \end{aligned}$$

Let \(Q_1(z_0)\subset (\mathbb {R}^3\backslash B_R)\times (-T,0]\), by (3.22), we have \(A(\widetilde{v},\nabla \widetilde{e},\theta ,z_0)+E(\widetilde{v},\nabla \widetilde{e},\theta ,z_0)\le C(M,N)\) for any \(0<\theta \le 1.\) Thus, by the interpolation inequality we have

$$\begin{aligned} \theta ^{-5/3}\int _{Q_\theta (z_0)}|\widetilde{v}|^{10/3}+|\nabla \widetilde{e}|^{10/3}\mathrm{d}x\mathrm{d}t\le c(M,N). \end{aligned}$$

Thus

$$\begin{aligned} C_1(\widetilde{v},1, z_0)&\le \gamma ^3|Q_1(z_0)|+\mathop {\iint }\limits _{Q_1(z_0)\cap \{|\widetilde{v}|>\gamma \}}|\widetilde{v}|^3dxdt\\&\le c\gamma ^3+\Vert \widetilde{v}\Vert _{L^{10/3}(Q_1(z_0))}\left| Q_1(z_0)\cap \left\{ |\widetilde{v}|>\gamma \right\} \right| ^{1/10}\\&\le c(\gamma ^3+\eta ^\frac{1}{10}). \end{aligned}$$

For any \(\epsilon >0\) we choose \(\gamma \) and \(\eta \) such that \(C_1(\widetilde{v},1,z_0)<\epsilon \).

It is easy to see that, by [28] (to see Lemma 3.2),

$$\begin{aligned} K(\nabla \widetilde{e},\theta , z_{0})\le \,& {} c\theta ^{-3}C_1(\widetilde{v},1,z_0)^{\frac{2}{3}}\left[ A(\nabla \widetilde{e}, 1,z_0) +E(\nabla \widetilde{e}, 1,z_0)\right] +\theta ^{2} K(\nabla \widetilde{e}, 1,z_0))\nonumber \\\le\, & {} c(M,N) (\theta ^{-3}\epsilon ^{2/3}+\theta ^2) \end{aligned}$$
(3.23)

Utilizing (3.23) and Hölder’s inequality we have

$$\begin{aligned} C_1(\nabla \widetilde{e},\theta , z_{0})\equiv \,& {} \theta ^{-2}\int _{Q_{\theta }(z_{0})}|\nabla \widetilde{e}|^{3}\nonumber \\\le \,& {} \left[ \theta ^{-\frac{5}{3}}\int _{Q_{\theta }(z_{0})}|\nabla \widetilde{e}|^{\frac{10}{3}}\right] ^{\frac{3}{4}}\left[ \theta ^{-3}\int _{Q_{\theta }(z_{0})}|\nabla \widetilde{e}|^{2}\right] ^{\frac{1}{4}}\nonumber \\\le &\, {} c(M, N) K(\nabla \widetilde{e},\theta ;z_{0})^{\frac{1}{4}}\nonumber \\\le\, & {} c(M, N)(\theta ^{-3/4}\epsilon ^{1/6}+\theta ^{1/2}). \end{aligned}$$
(3.24)

Thus

$$\begin{aligned} C_1(\widetilde{v},\nabla \widetilde{e}, \theta , z_0)\le & {} \theta ^{-2}\epsilon + c(M, N)(\theta ^{-3/4}\epsilon ^{1/6}+\theta ^{1/2}). \end{aligned}$$

First we take \(\theta \) such that \(c(M,N)\theta ^{1/2}\le \varepsilon _0/2,\) then take \(\epsilon \) such that \(\theta ^{-2}\epsilon +c(M, N)\theta ^{-3/4}\epsilon ^{1/6}\le \varepsilon _0/2\), i.e.,

$$\begin{aligned} C_1(\widetilde{v},\nabla \widetilde{e}, \theta , z_0)\le \varepsilon _0, \end{aligned}$$

which implies that \(z_0\) is a regular point by Proposition 1.2, therefor \((\widetilde{v},\widetilde{e},\widetilde{q})\) are smooth and their derivatives are bounded in \((\mathbb {R}^{3}\setminus B_{2R})\times (-T/2,0)\). Next, we show

$$\begin{aligned} \nabla \widetilde{e}(\cdot ,0)=0. \end{aligned}$$
(3.25)

For any B(y) and \(\phi \in C_0^\infty (B(y)),\) since e is Hölder continuous (to see following Lemma 3.2), we have

$$\begin{aligned}&\left| \int _{B(y)}\nabla \widetilde{e}(x,0)\phi dx\right| \\&\quad \le \int _{B(y)}|\nabla \widetilde{e}(x,0)-\nabla e_{k}(x,0)|\mathrm{d}x+\left| \int _{B(y)}\nabla e_{k}\phi \right| \\&\quad \le c\,\Vert \nabla \widetilde{e}(x,0)-\nabla e_{k}(x,0)\Vert _{L^{4/3}(B(y))}+c\,r_{k}^{-3}\int _{B_{r_{k}}(r_{k}y)}|e(x,0)-e(0,0)|\mathrm{d}x\\&\quad \le o(1)+c\, r_{k}^{-3}r_{k}^{3}r_{k}^{\alpha }\\&\quad \le o(1). \end{aligned}$$

The backward uniqueness theorem of parabolic equations [3], we conclude

$$\begin{aligned} \widetilde{e}(x,t)=0,\quad \text {in }\mathbb {R}^3\backslash \bar{B}_{2R}(0)\times (-T/2,0]. \end{aligned}$$

Using unique continuation theorem of parabolic equation in the bounded domain again [3], we conclude that

$$\begin{aligned} \widetilde{e}(x,t)=0\quad \text {in }\mathbb {R}^3\times (-T/2,0). \end{aligned}$$

Thus, \(\widetilde{v}\) satisfies Navier–Stokes equations in \(\mathbb {R}^3\times (-T/2,0)\)

$$\begin{aligned} {\left\{ \begin{array}{ll} &{}\partial _t\widetilde{v}-\Delta \widetilde{v}+\widetilde{v}\cdot \nabla \widetilde{v}+\nabla \widetilde{q}=0,\\ &{}\mathrm {div} \widetilde{v}=0.\end{array}\right. } \end{aligned}$$

Using (3.20) and backward uniqueness of heat operator again [3], we get

$$\begin{aligned} \widetilde{v}(x,t)=0\quad \text {in }\mathbb {R}^3\times (-T/2,0), \end{aligned}$$

which is a contradiction with (3.21).\(\square \)

We need following lemma.

Lemma 3.2

For \(v\in L^\infty (0, T; L^{3,\infty }(\mathbb {R}^3))\), if e satisfies in \(\mathbb {R}^3\times (0,T)\)

$$\begin{aligned} \partial _t e-\Delta e=-v\cdot \nabla e. \end{aligned}$$

Then e is Hölder continuous.

Proof

From [34], if \(v\in L^\infty (0, T; BMO^{-1}(\mathbb {R}^3))\), then e is Hölder continuous. We only prove the following inclusion relationship for \(3<p<\infty \),

$$\begin{aligned} L^{3}_{w}(\mathbb {R}^{3})\subset \dot{B}^{-1+\frac{3}{p}}_{p,\infty }(\mathbb {R}^{3})\subset BMO^{-1}(\mathbb {R}^{3}), \end{aligned}$$

where \(L^{3}_{w}(\mathbb {R}^{3})=L^{3,\infty }(\mathbb {R}^{3})\) is the weak Lebesgue space, and \(\dot{B}^{-1+\frac{3}{p}}_{p,\infty }(\mathbb {R}^{3})\) is a homogeneous Besov space.

The first inclusion is obtained through Sobolev embedding and real interpolation. To be specific, we write weak space \(L^{3}_{w}\) a real interpolation

$$\begin{aligned} L^{3}_{w}(\mathbb {R}^{3})=(L^{2}(\mathbb {R}^{3}),L^{p}(\mathbb {R}^{3}))_{\theta ,\infty }, \end{aligned}$$

where \(\theta =\frac{p}{3(p-2)}\). Notice we have the following two embedding relations

$$\begin{aligned} L^{2}(\mathbb {R}^{3})=\, & {} F^{0}_{2,2}(\mathbb {R}^{3})=B^{0}_{2,2}(\mathbb {R}^{3})\subset B^{\frac{3}{p}-\frac{3}{2}}_{p,2}(\mathbb {R}^{3}),\\ L^{p}(\mathbb {R}^{3})=\, & {} F^{0}_{p,2}(\mathbb {R}^{3})\subset B^{0}_{p,p}(\mathbb {R}^{3}), \end{aligned}$$

where we have used Littlewood–Paley Theorem to characterize \(L^{r}\) by the Triebel–Lizorkin space \(F^{0}_{r,2}\) (for \(1<r<\infty \)). Thus the identity map is bounded:

$$\begin{aligned}&id:L^{2}(\mathbb {R}^{3})\rightarrow B^{\frac{3}{p}-\frac{3}{2}}_{p,2}(\mathbb {R}^{3}),\\&id:L^{p}(\mathbb {R}^{3})\rightarrow B^{0}_{p,p}(\mathbb {R}^{3}). \end{aligned}$$

By real interpolation, and \((1-\theta )\cdot (\frac{3}{p}-\frac{3}{2})+\theta \cdot 0=-1+\frac{3}{p}\), this leads to

$$\begin{aligned} B^{-1+\frac{3}{p}}_{p,\infty }(\mathbb {R}^{3})=(B^{\frac{3}{p}-\frac{3}{2}}_{p,2}(\mathbb {R}^{3}),B^{0}_{p,p}(\mathbb {R}^{3}))_{\theta ,\infty }, \end{aligned}$$

thus the identity map

$$\begin{aligned} id:L^{3}_{w}(\mathbb {R}^{3})\rightarrow B^{-1+\frac{3}{p}}_{p,\infty }(\mathbb {R}^{3})\subset \dot{B}^{-1+\frac{3}{p}}_{p,\infty }(\mathbb {R}^{3}) \end{aligned}$$

is also bounded. The second inclusion is obtained by using the heat kernel characterization of the corresponding spaces, denote \(s=-1+\frac{3}{p}\),

$$\begin{aligned}&||f||_{\dot{B}^{s}_{p,\infty }(\mathbb {R}^{3})}=\sup _{t>0}||t^{-\frac{s}{2}}e^{t\Delta }f||_{L^{p}(\mathbb {R}^{3})},\\&||f||_{BMO^{-1}(\mathbb {R}^{3})}=\sup _{x\in \mathbb {R}^{3},R>0}\left[\frac{1}{|B_{R}(x)|}\int _{B_{R}(x)}\int _{0}^{R^{2}}|e^{t\Delta }f|^{2}\mathrm{d}y\mathrm{d}t\right]^{\frac{1}{2}}. \end{aligned}$$

Direct calculation yields:

$$\begin{aligned} ||f||_{BMO^{-1}(\mathbb {R}^{3})}&= \sup _{x\in \mathbb {R}^{3},R>0}\left[\frac{1}{|B_{R}(x)|}\int _{0}^{R^{2}}\int _{B_{R}(x)}|e^{t\Delta }f|^{2}\mathrm{d}y\mathrm{d}\right]^{\frac{1}{2}}\\&\le \sup _{x\in \mathbb {R}^{3},R>0}\left[|B_{R}(x)|^{-\frac{2}{p}}\int _{0}^{R^{2}}(\int _{B_{R}(x)}|e^{t\Delta }f|^{p}dy)^{\frac{2}{p}}\mathrm{d}t\right]^{\frac{1}{2}}\\&\le \sup _{x\in \mathbb {R}^{3},R>0}\left[|B_{R}(x)|^{-\frac{2}{p}}\int _{0}^{R^{2}}||e^{t\Delta }f||_{L^{p}(\mathbb {R}^{3})}^{2}\mathrm{d}t\right]^{\frac{1}{2}}\\&\le \sup _{x\in \mathbb {R}^{3},R>0}\left[|B_{R}(x)|^{-\frac{2}{p}}\int _{0}^{R^{2}}t^{s}||f||^{2}_{\dot{B}^{s}_{p,\infty }(\mathbb {R}^{3})}\mathrm{d}t\right]^{\frac{1}{2}}\\&\le c||f||_{\dot{B}^{s}_{p,\infty }(\mathbb {R}^{3})}. \end{aligned}$$

The proof is thus finished.\(\square \)

4 Proof of Theorem 1.4

Define for any \( r>0\) such that \(Q_r(z_0)\subset Q_T,\)

$$\begin{aligned} C_2(u, r,z_0)=r^{-1}\int _{Q_r(z_0)}|u|^4,\quad C_2(\nabla d, r,z_0)=r^{-1}\int _{Q_r(z_0)}|\nabla d|^4. \end{aligned}$$

By interpolation inequality we have

$$\begin{aligned} C(\nabla d, R, z_0)\le A(\nabla d, R, z_0)^{6/7} C_2(\nabla d,R,z_0)^{4/7}\le c(c_0, R), \end{aligned}$$

and

$$\begin{aligned} C_2(u, R,z_0) \le c\, M^2\left[ A(u, R, z_0)+E(u,R,z_0)\right] , \end{aligned}$$

where \(\Vert u\Vert _{L^\infty (-1,0; L^{3,\infty }(\mathbb {R}^3))}\le M\). On the other hand, by Calderón-Zygmund theorem,

$$\begin{aligned} \Vert p(t)\Vert _{L^{s}(\mathbb {R}^3)}^{s}\le c\,\Vert (u,\nabla d)(t)\Vert _{L^{2s}(\mathbb {R}^3)}^{2s}\quad \text {for}\,\, 1<s<\infty , \end{aligned}$$

and

$$\begin{aligned} (u, \nabla d)\in L^4(0,T; L^3(\mathbb {R}^3)), \end{aligned}$$

which implies

$$\begin{aligned} p\in L^2(0,T; L^{3/2}(\mathbb {R}^3)), \end{aligned}$$

we have

$$\begin{aligned} D(p, R,z_0)\le c(c_0, R). \end{aligned}$$

Since for any \(0<r\le R,\)

$$\begin{aligned} C(u, r,z_0)\le c\, M^4, \end{aligned}$$

by Lemma 2.4, we have for any \(0<r\le R/2\) and \(z_0\in \Omega \times (0,T), \Omega \subset \subset \mathbb {R}^3\),

$$\begin{aligned}&A(u,\nabla d, r, z_0)+E(u,\nabla d,r,z_0)+C_1(u, \nabla d, r,z_0)+C_2(u, \nabla d, r,z_0)\nonumber \\&\qquad +C(\nabla d, r, z_0)+D(p, r,z_0)+D_1(p, r,z_0)\nonumber \\&\quad \le c(M, R, c_0)\equiv N. \end{aligned}$$
(4.1)

The number \(\varepsilon (M,N)\) of Proposition 3.1 can be determined.

Let S be a singular points set of (ud) at \(\left\{ (x,T): x\in \mathbb {R}^3\right\} \). Assume that it contains more than \(M^3\varepsilon ^{-4}\) elements. Letting \(P=[M^3\varepsilon ^{-4}]+1\), we can find P different singular points \(\left\{ (x_k, T): k = 1, 2,\ldots , P\right\} \) of the set S. We can choose \(R_0\le R\) such that \(B_{R_0}(x_k)\cap B_{R_0}(x_l)=\emptyset , k\ne l,\) and bounded domain \(\Omega \) such that \(\cup _{k=1}^PB_{R_0}(x_k)\subset \Omega \). According to Proposition 3.1, for all \(r\in (0,R_0/2]\), it holds true

$$\begin{aligned} \varepsilon \le \frac{1}{r^3} \left| \left\{ x\in B_r(x_k): |u(x,T)|>\frac{\varepsilon }{r}\right\} \right| \end{aligned}$$
(4.2)

for all \(k=1,2,\ldots , P.\) In particular, taking \(r=r_0=R_0/2,\) we have

$$\begin{aligned} P\varepsilon\le \,& {} \sum _{k=1}^P\frac{1}{r_0^3} \left| \left\{ x\in B_{r_0}(x_k): |u(x,T)|>\frac{\varepsilon }{r_0}\right\} \right| \\\le\, & {} \frac{1}{r_0^3} \left| \left\{ x\in \cup _{k=1}^P B_{r_0}(x_k): |u(x,T)|>\frac{\varepsilon }{r_0}\right\} \right| \\\le\, & {} \frac{1}{r_0^3} \left| \left\{ x\in \Omega : |u(x,T)|>\frac{\varepsilon }{r_0}\right\} \right| \\\le\, & {} \varepsilon ^{-3}\Vert u(\cdot , T)\Vert _{L^{3,\infty }(\Omega )}^3\\\le & {} \varepsilon ^{-3}M^3, \end{aligned}$$

i.e., \(P\le M^3\varepsilon ^{-4}<P,\) which is a contradiction. \(\square \)

5 Appendix: Proof of Proposition 1.2

According to the \(L^p\) theorem of Stokes system in [5], if \(\mathbf {f}\in W^{-1,q}(\Omega ;\mathbb {R}^3), 1<q<\infty \), and \(\Omega \) is a \(C^1\) bounded domain, the following Stokes equations

$$\begin{aligned} {\left\{ \begin{array}{ll}-\Delta v+\nabla p=\mathbf {f}, \\ \mathrm {div}\,v=0, \quad \int _\Omega p=0,\\ v|_{\partial \Omega }=0,\end{array}\right. } \end{aligned}$$
(5.1)

there exists exactly one solution \((v,p)\in W^{1,q}(\Omega )\times L^q(\Omega ),\) and

$$\begin{aligned} \Vert \nabla v\Vert _{L^{q}(\Omega )}+\Vert p\Vert _{L^q(\Omega )}\le c(q)\Vert \mathbf {f}\Vert _{W^{-1,q}(\Omega )}. \end{aligned}$$
(5.2)

Wolf’s the local pressure projection \(\mathcal {P}_q\) tell us,

$$\begin{aligned} \mathcal {P}_q: W^{-1,q}(\Omega )\rightarrow W^{-1,q}(\Omega ),\quad \mathcal {P}_q(\mathbf {f})=p. \end{aligned}$$

As in [39], we have following Lemma.

Lemma 5.1

Let (ud) be a weak solution of (1.1), then for every \(C^2\) bounded sub-domain \(\Omega \), and any \(\phi \in C_0^\infty (\Omega \times (0,T))\), there holds

$$\begin{aligned}&-\int _0^T\int _\Omega (u+\nabla p_h)\cdot \phi _t-\int _0^T\int _\Omega (u\otimes u+\nabla d\odot \nabla d+p_1\mathbf {I}): \nabla \phi \nonumber \\&\quad +\int _0^T\int _\Omega (\nabla u-p_2\mathbf {I}):\nabla \phi =0, \end{aligned}$$
(5.3)

i.e., set \(v_\Omega =v:=u+\nabla p_h\),

$$\begin{aligned} \partial _t v+\mathrm {div}\,(u\otimes u)+\nabla p_1+\nabla p_2=\Delta v-\nabla \cdot (\nabla d\odot \nabla d), \end{aligned}$$
(5.4)

where \(\mathbf {I}\) is identity matrix, and

$$\begin{aligned} {\left\{ \begin{array}{ll}p_h=-\mathcal {P}_2(u),\\ p_1=-\mathcal {P}_{3/2}(u\otimes u+\nabla d\odot \nabla d),\\ p_2=\mathcal {P}_2(\Delta u). \end{array}\right. } \end{aligned}$$

In addition, following estimates hold for a.e. \(t\in (0,T)\)

$$\begin{aligned} {\left\{ \begin{array}{ll}\Vert \nabla p_h(t)\Vert _{L^m(\Omega )}\le c\,\Vert u(t)\Vert _{L^m(\Omega )}, \quad 1<m\le 6,\\ \Vert p_1(t)\Vert _{L^{3/2}(\Omega )}\le c\,\Vert u\otimes u+\nabla d\odot \nabla d\Vert _{L^{3/2}(\Omega )},\\ \Vert p_2(t)\Vert _{L^2(\Omega )}\le c\,\Vert \nabla u(t)\Vert _{L^2(\Omega )}. \end{array}\right. } \end{aligned}$$
(5.5)

Here \(c>0\) depends on the geometry of \(\Omega \) and in (5.5)\(_1\) on m only. In particular, if \(\Omega \) is the ball \(B_R(x_0)\) then c in (5.5)\(_1\) depends only on m, while in (5.5)\(_2\) and (5.5)\(_3\) c is an absolute constant.

Hence, we have local energy inequality, for \(\varphi \in C_0^\infty (\Omega \times (0,T))\)

$$\begin{aligned}&\int _\Omega (|v(t)|^2+|\nabla d(t)|^2)\varphi +2\int _0^t\int _\Omega (|\nabla v|^2+|\nabla ^2 d|^2)\varphi \nonumber \\&\quad \le \int _0^t\int _\Omega (|v|^2+|\nabla d|^2)(\varphi _t+\Delta \varphi )+\int _0^t\int _\Omega (|u|^2u+|\nabla d|^2v)\cdot \nabla \varphi \nonumber \\&\qquad +\int _0^t\int _\Omega 2(p_1+p_2)v\cdot \nabla \varphi + 2\int _0^t\int _\Omega u^iu^j\partial _i(\partial _j p_h\varphi )\nonumber \\&\qquad +2\int _0^t\int _\Omega (u\cdot \nabla d)\cdot (\nabla d\nabla \varphi )\nonumber \\&\qquad + \int _0^t\int _\Omega 2\nabla ^2p_h:(\nabla d\odot \nabla d)\varphi -|\nabla d|^2\nabla p_h\cdot \nabla \varphi \nonumber \\&\qquad -2\int _0^t\int _\Omega \nabla _xf(d)\nabla d\,\varphi . \end{aligned}$$
(5.6)

Note that the suitable weak solution of (1.1) satisfies the local energy inequality (5.6).

From local energy inequality we can get the Caccioppoli-type estimates

$$\begin{aligned}&\Vert W\Vert _{L^{10/3}(Q_{R/2})}^2+\Vert \nabla W\Vert _{L^2(Q_{R/2})}^2\nonumber \\&\quad \le c\,R^{-1/3}\Vert W\Vert _{L^{3}(Q_R)}^2+ c\,R^{-1}\Vert W\Vert _{L^{3}(Q_R)}^3. \end{aligned}$$
(5.7)

Here, \(W=(u, \nabla d)\), and

$$\begin{aligned} \Vert W\Vert _{L^{k}(Q_r)}^2=\, & {} \Vert u\Vert _{L^{k}(Q_r)}^2+\Vert \nabla d\Vert _{L^{k}(Q_r)}^2,\\ \Vert \nabla W\Vert _{L^{2}(Q_r)}^2=\, & {} \Vert \nabla u\Vert _{L^{2}(Q_r)}^2+\Vert \nabla ^2 d\Vert _{L^{2}(Q_r)}^2. \end{aligned}$$

Obviously,

$$\begin{aligned} C_1(W, r, z_0)=C_1(u, \nabla d, r, z_0), \quad E(W,r,z_0)=E(u,\nabla d, r, z_0). \end{aligned}$$

Proposition 1.2 is an immediate result of following lemma and Theorem (A).

Lemma 5.2

Suppose that (ud) is a local suitable weak solution of (1.1). Then there exist universal constants \(\varepsilon ^*>0\) and \(\theta \in (0,\frac{1}{4}]\) with following property. For any \(\varepsilon \in (0, \varepsilon ^*]\) if

$$\begin{aligned} C_{1}(W, 1, z_0)\le \varepsilon , \end{aligned}$$

then

$$\begin{aligned} C_{1}(W,\theta , z_0)\le \varepsilon . \end{aligned}$$

Proof

We prove by contradiction. Let \(\theta \in (0,\frac{1}{4}]\) be a constant to be specified later. Suppose there exist a decreasing sequence \(\{\varepsilon _n\}\) converging to 0, and a sequence of pairs of local suitable weak solutions \((u_n, d_n, p_n )\) such that

$$\begin{aligned} C_{1}(W_n, 1, z_0)=\varepsilon _n^3, \end{aligned}$$
(5.8)

and

$$\begin{aligned} C_{1}(W_n, \theta , z_0)>\varepsilon _n^3. \end{aligned}$$
(5.9)

Define \((v_n, e_n, q_n)=(\frac{u_n}{\varepsilon _n}, \frac{d_n}{\varepsilon _n},\frac{p_n}{\varepsilon _n}),\) then they satisfy

$$\begin{aligned}&\partial _t v_n+\varepsilon _n v_n\cdot \nabla v_n+\nabla q_n=\Delta v_n-\varepsilon _n\mathrm {div}\,(\nabla e_n\odot \nabla e_n), \quad \mathrm {div} v_n=0,\\&\partial _t e_n+\varepsilon _n v_n\cdot \nabla e_n-\Delta e_n=-\sigma ^{-2}(|d_n|^2-1))e_n. \end{aligned}$$

Write \(w_n=(v_n, \nabla e_n)\), then

$$\begin{aligned} C_{1}(w_n, 1, z_0)=1, \end{aligned}$$
(5.10)

and

$$\begin{aligned} C_{1}(w_n, \theta , z_0)>1. \end{aligned}$$
(5.11)

Using the Caccioppoli estimate (5.7) we conclude

$$\begin{aligned} \Vert w_n\Vert _{L^{10/3}(Q_{1/2}(z_0))}+\Vert \nabla w_n\Vert _{L^2(Q_{1/2}(z_0))}\le c, \end{aligned}$$
(5.12)

which implies

$$\begin{aligned} \Vert w_n\cdot \nabla w_n\Vert _{L^{5/4}(Q_{1/2}(z_0))}\le \Vert \nabla w_n\Vert _{L^2(Q_{1/2}(z_0))}\Vert w_n\Vert _{L^{10/3}(Q_{1/2}(z_0))}\le c. \end{aligned}$$

The coercive estimate for the Stokes system (see, for instance, [27]) with a suitable cutoff function implies

$$\begin{aligned} \int _{Q_{\frac{1}{3}}} |\partial _tw_n|^{\frac{5}{4}}+|\nabla ^2w_n|^{\frac{5}{4}}+|\nabla q_n|^{\frac{5}{4}}+|w_n|^{\frac{5}{4}}\le c, \end{aligned}$$

where the constant c is independent of n. Thanks to the compact embedding theorem and (5.12), there exist \(w\in L^{3}(Q_{1/3}(z_0))\) and \(q\in L^{\frac{5}{4}}(Q_{1/3}(z_0))\) such that

$$\begin{aligned}&w_n\rightarrow w=(v,\nabla e) \quad \text {in}\,\, L^{3}(Q_{1/3}(z_0)),\\&q_n\rightharpoonup q \quad \text {in}\,\, L^{\frac{5}{4}}(Q_{1/3}(z_0)). \end{aligned}$$

Thus, (veq) satisfy

$$\begin{aligned}&\partial _t v-\Delta v+\nabla q=0, \quad \mathrm {div}\, v=0,\\&\partial _t e-\Delta e=\sigma ^{-2}e. \end{aligned}$$

Moreover

$$\begin{aligned} \Vert w\Vert _{L^{3}(Q_{1/3}(z_0))}+\Vert q\Vert _{L^{\frac{5}{4}}(Q_{3/4}(z_0))}\le c. \end{aligned}$$

By the classical estimate of the Stokes system [37], we get

$$\begin{aligned} \sup _{Q_{1/3}(z_0)}|w|\le c, \end{aligned}$$

which implies that for \(0<\theta \le 1/3\)

$$\begin{aligned} C_{1}(w,\theta , z_0)\le c\,\theta ^3. \end{aligned}$$

This contradicts (5.11), if we choose \(\theta \) sufficiently small. The lemma is proved.\(\square \)