Skip to main content

Blow-up criteria of the simplified Ericksen–Leslie system

Abstract

In this paper, we establish scaling invariant blow-up criteria for a classical solution to the simplified Ericksen–Leslie system in terms of the positive part of the second eigenvalue of the strain matrix and orientation field in mixed-norm Lebesgue spaces. Our result may be also regarded as an extension or improvement of the corresponding results obatined by Neustupa and Penel (Trends in Partial Differential Equations of Mathematical Physics, pp. 197–212, 2005), Miller (Arch. Ration. Mech. Anal. 235(1):99–139, 2020) and Huang and Wang (Commun. Partial Differ. Equ. 37(5):875–884, 2012).

1 Introduction

In this paper, we consider the following Cauchy problem for the three-dimensional simplified Ericksen–Leslie system:

$$ \textstyle\begin{cases} \partial _{t} u- \Delta u+(u \cdot \nabla ) u+\nabla p=- \nabla \cdot (\nabla d \odot \nabla d) & \text{in } \mathbb{R}^{3} \times \mathbb{R}_{+}, \\ \partial _{t} d+(u \cdot \nabla ) d=\Delta d+ \vert \nabla d \vert ^{2} d & \text{in } \mathbb{R}^{3} \times \mathbb{R}_{+}, \\ \nabla \cdot u=0, \qquad \vert d \vert =1 & \text{in } \mathbb{R}^{3} \times \mathbb{R}_{+}, \end{cases} $$
(1.1)

where \(u: \mathbb{R}^{3} \times \mathbb{R}_{+} \rightarrow \mathbb{R}^{3}\) is the unknown velocity field of the flow, \(p: \mathbb{R}^{3} \times \mathbb{R}_{+} \rightarrow \mathbb{R}\) is the scalar pressure and \(d: \mathbb{R}^{3} \times \mathbb{R}_{+} \rightarrow \mathbb{S}^{2}\), is the macroscopic average of the nematic liquid-crystal orientation field in \(\mathbb{R}^{3}\), \(\nabla \cdot u=0\) represents the incompressible condition, and the notation \(\nabla d \odot \nabla d\) denotes the \(3 \times 3\) matrix whose \((i, j)\)th component is given by \(\partial _{i} d \cdot \partial _{j} d \,(1 \leq i, j \leq 3)\). We will consider the Cauchy problem (1.1) with the initial conditions

$$ u |_{t=0}=u_{0}(x), \quad \text{and}\quad d |_{t=0}=d_{0}(x), \qquad \bigl\vert d_{0}(x) \bigr\vert =1 \quad \text{in } \mathbb{R}^{3} $$

and far-field behaviors

$$ u \rightarrow 0, \qquad d \rightarrow \bar{d}_{0} \quad \text{as } \vert x \vert \rightarrow \infty , $$

where \(u_{0}\) is a given initial velocity with \(\nabla \cdot u_{0}=0\) in the distribution sense, \(d_{0}: \mathbb{R}^{3} \rightarrow \mathbb{S}^{2}\) is a given initial liquid-crystal orientation field and \(\bar{d}_{0}\) is a constant vector with \(\vert \bar{d}_{0} \vert =1\).

The parabolic system (1.1) was first proposed by Lin [25] as a simplification of the general Ericksen–Leslie system that models the hydrodynamic flow of nematic liquid-crystal material [11, 20]. The simplified Ericksen–Leslie system (1.1) can be viewed as the incompressible Navier–Stokes equations coupling with the heat flow of a harmonic map, which has been successful in modeling various dynamical behaviors for nematic liquid crystals. When the velocity field u is identically vanishing, the system (1.1) becomes the heat flow of harmonic maps. Chang, Ding, and Ye [6] proved that strong solution blowup in finite time to the harmonic heat-flow equation. Wang [35] established a Serrin-type regularity criteria, which implies that if the solution d blowup at time \(T_{*}\), then

$$ \sup_{0 \leq t< T_{*}} \bigl\Vert \nabla d(\cdot , t) \bigr\Vert _{L^{n}}=\infty . $$

For a more detailed physical background, please refer to [22, 24] and the references therein.

From the mathematical point of view, the simplified Ericksen–Leslie system (1.1) has recently acquired much interest in the research community. Recently, Lin, Lin, and Wang [23] and Hong [17] independently proved the global existence of Leray–Hopf-type weak solutions to the system (1.1) for any smooth bounded domain in \(\mathbb{R}^{2}\) and the whole space \(\mathbb{R}^{2}\), respectively. For the case of three dimensions, Li and Wang [21] established the existence of a local strong solution with large initial value and the global strong solution with a small initial value for the initial boundary value problem of system (1.1). To characterize the first singular time, Huang and Wang [18] considered the so-called Beale–Kato–Majda-type blow-up criterion, more precisely, they proved \(0< T_{*}<+\infty \) is the maximal time interval if and only if

$$ n=3,\qquad \vert \omega \vert + \vert \nabla d \vert ^{2} \notin L_{t}^{1} L_{x}^{\infty} \bigl(\mathbb{R}^{3} \times [0, T_{*} ] \bigr); $$
(1.2)

and

$$ n=2,\qquad \vert \nabla d \vert ^{2} \notin L_{t}^{1} L_{x}^{\infty} \bigl( \mathbb{R}^{2} \times [0, T_{*} ] \bigr), $$
(1.3)

where \(\omega =\nabla \times u\). Liu and Zhao [26] showed that the smooth solution \((u, d)\) of system (1.1) blows up at the time \(T_{*}\) if and only if

$$ \int _{0}^{T_{*}} \frac{ \Vert \omega (\cdot , t) \Vert _{\dot{B}_{\infty , \infty}^{0}}+ \Vert \nabla d(\cdot , t) \Vert _{\dot{B}_{\infty , \infty}^{2}}^{2}}{\sqrt{1+\ln (e+ \Vert \omega (\cdot , t) \Vert _{\dot{B}_{\infty , \infty}^{0}}+ \Vert \nabla d(\cdot , t) \Vert _{\dot{B}_{\infty , \infty}^{0}} )}} \, {d} t= \infty . $$
(1.4)

Chen, Tan, and Wu [9] obtained the Serrin-type regularity criterion [34], which states that

$$ u \in L^{q} \bigl(0, T ; L^{p} \bigr), \qquad \nabla d \in L^{r} \bigl(0, T ; L^{s} \bigr), $$
(1.5)

with \(\frac{2}{q}+\frac{3}{p}=1\), \(\frac{2}{r}+\frac{3}{s}=1\), \(3< p \leq \infty \), \(3< s \leq \infty \). Lee [19] obtained the Beirão da Veiga-type blow-up criterion [10], which states that

$$ \operatorname{curl} u \in L^{q} \bigl(0, T ; L^{p} \bigr), \qquad \nabla d \in L^{r} \bigl(0, T ; L^{s} \bigr), $$
(1.6)

with \(\frac{2}{q}+\frac{3}{p}=2\), \(\frac{2}{r}+\frac{3}{s}=1\), \(\frac{3}{2}< p \leq \infty \), \(3< s \leq \infty \). For system (1.1), some refined blow-up criteria of (1.5) and (1.6) are later proven in [13, 14, 41].

When the macroscopic average of the nematic liquid-crystal orientation field is neglected, i.e., \(d=\text{constant vector}\), system (1.1) reduces to the incompressible Navier–Stokes equations (in short NSE). Many classical Serrin-type criteria and Beirão da Veiga-type criteria for the regularity of weak solutions have been proved, please refer to [24, 7, 8, 12, 16, 32, 33, 3740].

There is numerical evidence for the Navier–Stokes or Euler equation in [15] regarding the tendency of the vorticity to align with the eigenvector of the strain tensor corresponding to the intermediate eigenvalue \(\lambda _{2}\) and later Neustupa and Penel, Chae, and Miller independently gave the analytical evidence of this fact in [5, 27, 2931]. Specifically, Neustupa and Penel [2931] and Miller [27] proved that

$$ \lambda ^{+}_{2} \in L^{q} \bigl(0, T ; L^{p} (\Omega ) \bigr), \qquad \frac{2}{q}+ \frac{3}{p}=2,\qquad \frac{3}{2}< p \leq \infty $$
(1.7)

implies the smoothness of the solution to the Navier–Stokes equations, where \(\lambda ^{+}_{2}(x)=\max \{\lambda _{2}(x), 0\}\), Ω be a bounded domain or \(\Omega =\mathbb{R}^{3}\). Chae [5] proved that the dynamical behaviors of the \(L^{2}\) norm of vorticity is controlled completely by the second largest eigenvalue \(\lambda ^{+}_{2}\) of the deformation tensor for the 3D incompressible Euler equations. Recently, the second named author [36] extended the above regularity criteria to the Multiplier space and Besov space. More recently, Miller [28] extended the Serrin-type and the Beirão da Veiga-type criteria to the Lebeguse sum spaces for singularities of a local smooth solution.

From the physical point of view, the fluid behavior can be different in different directions. Therefore, understanding the asymptotic behaviors of solutions to the simplified Ericksen–Leslie system in anisotropic functional spaces seems to be an interesting topic. This leads us to focus on the blow-up criteria for the 3D simplified Ericksen–Leslie system (1.1) on the framework of the mixed-norm Lebesgue space.

It is well known that if the initial velocity \(u_{0} \in H^{s} (\mathbb{R}^{n}, \mathbb{R}^{n} )\) with \(\nabla \cdot u_{0}=0\) and \(d_{0} \in H^{s+1} (\mathbb{R}^{n}, \mathbb{S}^{2} )\) for \(s \geq n\), then there is \(T_{0}>0\) depending only on \(\Vert u_{0} \Vert _{H^{s}}\) and \(\Vert d_{0} \Vert _{H^{s+1}}\) such that (1.1) has a unique, classical solution \((u, d)\) in \(\mathbb{R}^{n} \times [0, T_{0} )\) satisfying

$$ \begin{gathered} u \in C \bigl([0, T], H^{s} \bigl(\mathbb{R}^{n} \bigr) \bigr) \cap C^{1} \bigl([0, T], H^{s-1} \bigl(\mathbb{R}^{n} \bigr) \bigr), \\ d \in C \bigl([0, T], H^{s+1} \bigl(\mathbb{R}^{n}, \mathbb{S}^{2} \bigr) \bigr) \cap C^{1} \bigl([0, T], H^{s} \bigl(\mathbb{R}^{n}, \mathbb{S}^{2} \bigr) \bigr) \end{gathered} $$
(1.8)

for any \(0< T< T_{0}\). Assume \(T_{*}>0\) is the maximum value such that (1.8) holds with \(T_{0}=T_{*} \). We would like to characterize such a \(T_{*}\). To facilitate the presentation of the result, let us first recall the definition of the mixed-norm Lebesgue space.

Definition 1.1

For a given \(\vec{p}= (p_{1}, p_{2}, p_{3} ) \in [1, \infty )^{3}\), the mixed norm Lebesgue space \(L^{\vec{p}} (\mathbb{R}^{d} )\) is defined to be the space consisting of all measurable functions \(f: \mathbb{R}^{3} \rightarrow \mathbb{R}\) such that the norm

$$ \Vert f \Vert _{L^{\vec{p}} (\mathbb{R}^{3} )}= \biggl( \int _{- \infty}^{\infty} \biggl( \int _{-\infty}^{\infty} \biggl( \int _{-\infty}^{ \infty} \bigl\vert f(x) \bigr\vert ^{p_{1}} \,dx _{1} \biggr)^{\frac{p_{2}}{p_{1}}} \,dx _{2} \biggr)^{\frac{p_{3}}{p_{2}}} \,dx _{3} \biggr)^{\frac{1}{p_{3}}}< \infty . $$

Similar definitions can be formulated if any of \(\{p_{1}, p_{2}, p_{3} \}\) is the same as ∞.

Now, we state our main result as follows:

Theorem 1.1

For \(s \geq 3\), \(u_{0} \in H^{s} (\mathbb{R}^{3}, \mathbb{R}^{3} )\) with \(\nabla \cdot u_{0}=0\) and \(d_{0} \in H^{s+1} (\mathbb{R}^{3}, \mathbb{S}^{2} )\), let \(T_{*}>0\) be the maximum value such that system (1.1) has a unique solution \((u, d)\) satisfying (1.8) with \(T_{0}\) replaced by \(T_{*}\), and let \(\lambda _{1}(x) \leq \lambda _{2}(x) \leq \lambda _{3}(x)\) be the eigenvalues of the strain tensor \(S=\nabla _{\operatorname{sym}} u=\frac{1}{2} (\partial _{i}u_{j}+ \partial _{j}u_{i} )\). Let \(\lambda _{2}^{+}(x)=\max \{\lambda _{2}(x), 0 \}\). If \(T_{*}<+\infty \), then

$$ \int ^{T_{*}}_{0} \frac{ \Vert \lambda _{2}^{+}(t) \Vert ^{q}_{L^{\vec{p}}} + \Vert \nabla d(t) \Vert ^{2q}_{L^{2\vec{p}}} }{1+\ln (e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} )}\,dt = \infty , $$
(1.9)

with \(2< p_{i}\leq \infty \), \(\frac{2}{q}+\sum^{3}_{i=1}\frac{1}{p_{i}}=2\), \(1-\sum^{3}_{i=1}\frac{1}{p_{i}}\geq 0\) and \(2< s_{3}<\infty \), \(\sum^{3}_{i=1}\frac{1}{s_{i}}=\frac{1}{2}\).

Remark 1.1

We note that when \(p_{1} = p_{2} = p_{3}=p\), the mixed-norm Lebesgue space \(L^{\vec{p}}\) is reduced to the usual Lebesgue space \(L^{p}\). Theorem 1.1 naturally extends and improves the blow-up criteria as stated in [18, 19, 27]. In addition, we show the logarithmic blow-up criterion. To the authors’ knowledge, Theorem 1.1 is the improvement result on blow-up criteria via the mixed Lebesgue norm in the denominator.

Remark 1.2

From Theorem 1.1, it is easy to see that if there exists a constant \(M > 0\) such that

$$ \int ^{T_{*}}_{0} \frac{ \Vert \lambda _{2}^{+}(t) \Vert ^{q}_{L^{\vec{p}}} + \Vert \nabla d(t) \Vert ^{2q}_{L^{2\vec{p}}} }{1+\ln (e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} )}\,dt \leq M, $$

with \(2< p_{i}\leq \infty \), \(\frac{2}{q}+\sum^{3}_{i=1}\frac{1}{p_{i}}=2\), \(1-\sum^{3}_{i=1}\frac{1}{p_{i}}\geq 0\), and \(2< s_{3}<\infty \), \(\sum^{3}_{i=1}\frac{1}{s_{i}}=\frac{1}{2}\), then the local smooth solution \((u, d)\) can be extended beyond the time \(T_{*}\).

The proof of Theorem 1.1 will be given in Sect. 2. Before concluding this section, we list the following lemmas that are needed in Sect. 2.

Lemma 1.1

([27])

For all \(-\frac{3}{2}<\alpha <\frac{3}{2}\) and for all u divergence free in the sense that \(\xi \cdot \hat{u}(\xi )=0\) almost everywhere,

$$ \Vert S \Vert _{\dot{H}^{\alpha}}^{2}= \Vert A \Vert _{\dot{H}^{\alpha}}^{2}=\frac{1}{2} \Vert \omega \Vert _{\dot{H}^{\alpha}}^{2}=\frac{1}{2} \Vert \nabla \otimes u \Vert _{ \dot{H}^{\alpha}}^{2}, $$
(1.10)

where the symmetric part \(S=S_{i j}=\frac{1}{2} (\frac{\partial u_{j}}{\partial x_{i}}+ \frac{\partial u_{i}}{\partial x_{j}} )\), which we refer to as the strain tensor, and the antisymmetric part \(A=A_{i j}=\frac{1}{2} (\frac{\partial u_{j}}{\partial x_{i}}- \frac{\partial u_{i}}{\partial x_{j}} )\), \(\omega =\nabla \times u\).

Lemma 1.2

([27])

Suppose \(S \in L^{\infty} ([0, T] ; L^{2} (\mathbb{R}^{3} ) ) \cap L^{2} ([0, T] : \dot{H}^{1} (\mathbb{R}^{3} ) )\) is a local strong solution to the Navier–Stokes strain equation and \(S(x)\) has eigenvalues \(\lambda _{1}(x) \leq \lambda _{2}(x) \leq \lambda _{3}(x)\). Define

$$ \lambda _{2}^{+}(x)=\max \bigl\{ \lambda _{2}(x), 0 \bigr\} , $$

then

$$ -\operatorname{det}(S) \leq \frac{1}{2} \vert S \vert ^{2} \lambda _{2}^{+}. $$

Lemma 1.3

([42])

For \(p_{1}, p_{2}, p_{3} \in [2, \infty )\) and \(0 \leq \sum^{3}_{i=1} \frac{1}{p_{i}}-\frac{1}{2} \leq 1\), there exists a positive constant C such that

$$ \begin{aligned} \Vert f \Vert _{L^{\vec{p}} (\mathbb{R}^{3} )} & \leq C \Vert \partial _{1} f \Vert _{L^{2} (\mathbb{R}^{3} )}^{ \frac{p_{1}-2}{2 p_{1}}} \Vert \partial _{2} f \Vert _{L^{2} ( \mathbb{R}^{3} )}^{\frac{p_{2}-2}{2 p_{2}}} \Vert \partial _{3} f \Vert _{L^{2} (\mathbb{R}^{3} )}^{ \frac{p_{3}-2}{2 p_{3}}} \Vert f \Vert _{L^{2} (\mathbb{R}^{3} )}^{ \sum ^{3}_{i=1} \frac{1}{p_{i}-\frac{1}{2}}} \\ & \leq C \Vert \nabla f \Vert _{L^{2} (\mathbb{R}^{3} )}^{ \frac{3}{2}-\sum _{i} \frac{1}{p_{i}}} \Vert f \Vert _{L^{2} (\mathbb{R}^{3} )}^{\sum _{i} \frac{1}{p_{i}}-\frac{1}{2}}. \end{aligned} $$

Lemma 1.4

([1])

Let \(\vec{s}= (s_{1}, s_{2}, s_{3} ) \in [2, \infty ]^{3}\) satisfy

$$ \frac{1}{s_{1}}+\frac{1}{s_{2}}+\frac{1}{s_{3}}= \frac{1}{2} \quad \textit{and} \quad s_{3} \in (2, \infty ). $$

Then, there exist constants \(C=C(\vec{s})\) such that

$$ \Vert u \Vert _{L^{\vec{s} } (\mathbb{R}^{3} )} \leq C \bigl[ \Vert D u \Vert _{L^{2} (\mathbb{R}^{3} )}+ \Vert u \Vert _{L^{2} (\mathbb{R}^{3} )} \bigr], \quad \forall u \in W^{1,2} \bigl(\mathbb{R}^{3} \bigr). $$

2 The proof of Theorem 1.1

This section is devoted to the proof of Theorem 1.1. We assume that (1.9) was not true, then there exists a positive constant K such that

$$ \int ^{T_{*}}_{0} \frac{ \Vert \lambda _{2}^{+}(t) \Vert ^{q}_{L^{\vec{p}}} + \Vert \nabla d(t) \Vert ^{2q}_{L^{2\vec{p}}} }{1+\ln (e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} )}\,dt \leq K. $$
(2.1)

Due to the Beale–Kato–Majda-type blowup criterion (1.2) in [18], it suffices to present the bound

$$ \int _{0}^{T_{*}} \bigl( \bigl\Vert \omega ( t) \bigr\Vert _{L^{\infty}}+ \bigl\Vert \nabla d( t) \bigr\Vert _{L^{ \infty}}^{2} \bigr) \,dt \leq C $$

under condition (2.1), which is enough to guarantee the extension of a local smooth solution \((u, d)\) beyond the time \(T_{*}\). That is to say, \([0, T_{*} )\) is not a maximal existence interval, and we obtain the desired contradiction.

Proposition 2.1

(Strain reformulation of the dynamics)

Suppose \((u, d)\) is a classical solution to the system (1.1). Then, strain tensor \(S=\nabla _{\mathrm{sym}}(u)\) sastifies

$$ \partial _{t} S+(u \cdot \nabla ) S-\Delta S+S^{2}+ \frac{1}{4} \omega \otimes \omega -\frac{1}{4} \vert \omega \vert ^{2} I_{3}+ \operatorname{Hess}(P)=-\nabla _{\mathrm{sym}}(\Delta d\cdot \nabla d), $$

where \(P=p+\frac{|\nabla d|^{2}}{2}\).

Proof

By Proposition 2.1 in [27] and noting that

$$ \nabla \cdot (\nabla d \otimes \nabla d)=\nabla \biggl( \frac{ \vert \nabla d \vert ^{2}}{2} \biggr)+\Delta d \cdot \nabla d, $$

this implies Proposition 2.1 immediately. □

Proof of Theorem 1.1

First, we give the basic energy estimate of system (1.1). Taking the inner product of (1.1)1 with u and (1.1)2 with \(-\Delta d\) in \(L^{2} (\mathbb{R}^{3} )\), respectively, and adding together, one has

$$ \frac{1}{2} \frac{d}{{d} t} \bigl\Vert (u, \nabla d) \bigr\Vert _{L^{2}}^{2}+ \Vert \nabla u \Vert _{L^{2}}^{2}=- \int _{\mathbb{R}^{3}} \Delta d \cdot \bigl( \Delta d+ \vert \nabla d \vert ^{2} d \bigr) \,dx $$

and then, we discover that

$$ \begin{aligned} & \frac{1}{2} \frac{{d}}{ {d} t} \bigl\Vert (u, \nabla d) \bigr\Vert _{L^{2}}^{2}+ \Vert \nabla u \Vert _{L^{2}}^{2}+ \bigl\Vert \Delta d+ \vert \nabla d \vert ^{2} d \bigr\Vert _{L^{2}}^{2} \\ &\quad = \int _{\mathbb{R}^{3}} \vert \nabla d \vert ^{2} d \cdot \bigl(\triangle d+ \vert \nabla d \vert ^{2} d \bigr) \, {d} x=0, \end{aligned} $$

where we have used the facts that

$$ \vert d \vert ^{2}=1\quad \Rightarrow \quad 0=\frac{1}{2} \Delta \vert d \vert ^{2}=d \cdot \triangle d+ \vert \nabla d \vert ^{2}. $$

Next, we derive the \(H^{1}\) estimate for \((u, \nabla d)\). Taking × on the first equation of (1.1), we obtain

$$ \partial _{t} \omega +(u \cdot \nabla ) \omega - \Delta \omega =S \omega -\nabla \times (\Delta d \cdot \nabla d) $$
(2.2)

and then taking the operator \(\nabla _{\mathrm{sym}}\) \((i.e., S=\nabla _{\mathrm{sym}}(u)_{i j}=\frac{1}{2} ( \frac{\partial u_{j}}{\partial x_{i}}+ \frac{\partial u_{i}}{\partial x_{j}} ) )\) to the (1.1)1, one obtains

$$ \partial _{t} S+(u \cdot \nabla ) S-\Delta S+S^{2}+\frac{1}{4} \omega \otimes \omega - \frac{1}{4} \vert \omega \vert ^{2} I_{3}+ \operatorname{Hess}(P)=-\nabla _{\mathrm{sym}}(\Delta d\cdot \nabla d), $$
(2.3)

where \(P=p+\frac{|\nabla d|^{2}}{2}\). Multiplying (2.2) by ω and integrating by parts over \(\mathbb{R}^{3}\), we find

$$ \begin{aligned} \frac{1}{2} \frac{d}{d t} \Vert \omega \Vert _{L^{2}}^{2}+ \Vert \nabla \omega \Vert _{L^{2}}^{2}& = \int _{\mathbb{R}^{3}} S\omega \cdot \omega \,dx - \int _{\mathbb{R}^{3}} \nabla \times (\Delta d \cdot \nabla d)\cdot \omega \,dx \\ & = \int _{\mathbb{R}^{3}} S\omega \cdot \omega \,dx + \int _{\mathbb{R}^{3}} (\Delta d \cdot \nabla d)\cdot \Delta u \,dx . \end{aligned} $$
(2.4)

Multiplying (2.3) by S, we deduce that

$$ \begin{aligned}& \frac{d}{d t} \Vert S \Vert _{L^{2}}^{2}+2 \Vert \nabla S \Vert _{L^{2}}^{2} \\ &\quad =-2 \int _{ \mathbb{R}^{3}} \operatorname{tr} \bigl(S^{3} \bigr) \,dx -\frac{1}{2} \int _{\mathbb{R}^{3}} \omega \otimes \omega \cdot S \,dx + \int _{ \mathbb{R}^{3}} (\Delta d \cdot \nabla d)\cdot \Delta u \,dx , \end{aligned} $$
(2.5)

where we used the following facts that were proved by Proposition 2.4 and Theorem 4.5 in [27]

$$\begin{aligned}& \bigl\langle \vert \omega \vert ^{2} I_{3}, S \bigr\rangle _{L^{2}}= 0,\\& \bigl\langle \operatorname{Hess}(P), S\bigr\rangle _{L^{2}}= 0 ,\\& \bigl\langle S^{2}, S \bigr\rangle _{L^{2}}= \int _{\mathbb{R}^{3}} \operatorname{tr} \bigl(S^{3} \bigr),\\& 2\bigl\langle \nabla _{\mathrm{sym}}(\Delta d \cdot \nabla d), S\bigr\rangle _{L^{2}}= \bigl\langle (\Delta d \cdot \nabla d), -\Delta u\bigr\rangle _{L^{2}}. \end{aligned}$$

From Lemma 1.1, for the identity (2.4), it follows that

$$ \begin{aligned} \frac{d}{d t} \Vert S \Vert _{L^{2}}^{2}+2 \Vert \nabla S \Vert _{L^{2}}^{2}& = \int _{ \mathbb{R}^{3}} S\omega \otimes \omega \,dx + \int _{\mathbb{R}^{3}} ( \Delta d \cdot \nabla d)\cdot \Delta u \,dx . \end{aligned} $$
(2.6)

Adding \(\frac{2}{3}\) (2.5) and \(\frac{1}{3}\) (2.6), we conclude that

$$ \begin{aligned} \frac{d}{d t} \Vert S \Vert _{L^{2}}^{2}+2 \Vert \nabla S \Vert _{L^{2}}^{2}& = - \frac{4}{3} \int _{\mathbb{R}^{3}} \operatorname{tr} \bigl(S^{3} \bigr)+ \int _{\mathbb{R}^{3}} (\Delta d \cdot \nabla d)\cdot \Delta u \,dx \\ &\stackrel{\mathrm{def}}{=} I_{1}+I_{2}, \end{aligned} $$
(2.7)

where

$$ I_{1}=-\frac{4}{3} \int _{\mathbb{R}^{3}} \operatorname{tr} \bigl(S^{3} \bigr),\qquad I_{2}= \int _{\mathbb{R}^{3}} (\Delta d \cdot \nabla d) \cdot \Delta u \,dx . $$
(2.8)

We first estimate the term \(I_{1}\), since \(\operatorname{tr} (S )=\nabla \cdot u=0\), it follows from Lemma 1.2 and Lemma 1.3 that

$$ \begin{aligned} I_{1}& = - \frac{4}{3} \int _{\mathbb{R}^{3}} \operatorname{tr} \bigl(S^{3} \bigr) \,dx \\ & = -\frac{4}{3} \int _{\mathbb{R}^{3}}\lambda _{1}^{3}+\lambda _{2}^{3}+ \lambda _{3}^{3} \,dx \\ & = -\frac{4}{3} \int _{\mathbb{R}^{3}}\lambda _{1}^{3}+\lambda _{2}^{3}+ (-\lambda _{1}-\lambda _{2} )^{3} \,dx \\ & = -4 \int _{\mathbb{R}^{3}} (-\lambda _{1}-\lambda _{2} ) \lambda _{1} \lambda _{2} \,dx =-4 \int _{\mathbb{R}^{3}} \lambda _{1} \lambda _{2} \lambda _{3} \,dx \\ & = -4 \int _{\mathbb{R}^{3}} \operatorname{det}(S) \,dx \leq 2 \int _{ \mathbb{R}^{3}} \vert S \vert ^{2} \lambda _{2}^{+} \,dx \\ & \leq C \bigl\Vert \lambda _{2}^{+} \bigr\Vert _{L^{\vec{p}}} \Vert S \Vert _{L_{1}^{\frac{2 p_{1}}{p_{1}-2}}L_{2}^{ \frac{2 p_{2}}{p_{2}-2}}L_{3}^{\frac{2 p_{3}}{p_{3}-2}}} \Vert S \Vert _{L^{2}} \\ & \leq C \bigl\Vert \lambda _{2}^{+} \bigr\Vert _{L^{\vec{p}}} \Vert \nabla S \Vert ^{\frac{1}{p_{1}}+\frac{1}{p_{2}} +\frac{1}{p_{3}}}_{L^{2}} \Vert S \Vert ^{2- (\frac{1}{p_{1}}+\frac{1}{p_{2}} + \frac{1}{p_{3}} )}_{L^{2}} \\ & \leq C \bigl\Vert \lambda _{2}^{+} \bigr\Vert ^{q}_{L^{\vec{p}}} \Vert S \Vert ^{2}_{L^{2}}+ \frac{1}{8} \Vert \nabla S \Vert ^{2}_{L^{2}}, \end{aligned} $$
(2.9)

where \(q= \frac{2}{2- (\frac{1}{p_{1}}+\frac{1}{p_{2}}+\frac{1}{p_{3}} )}\), \(2< p_{i}\leq \infty \).

For the term \(I_{2}\), it follows from Lemma 1.3, Hölder’s inequality, and Young’s inequality that

$$\begin{aligned} I_{2} & \leq C \Vert \nabla d \Vert _{L^{2\vec{p}}} \Vert \Delta d \Vert _{L_{1}^{\frac{2 p_{1}}{p_{1}-1}}L_{2}^{ \frac{2 p_{2}}{p_{2}-1}}L_{3}^{\frac{2 p_{3}}{p_{3}-1}}} \Vert \Delta u \Vert _{L^{2}} \\ & \leq C \Vert \nabla d \Vert _{L^{2\vec{p}}} \Vert \nabla \Delta d \Vert ^{\frac{1}{2p_{1}}+\frac{1}{2p_{2}} + \frac{1}{2p_{3}}}_{L^{2}} \Vert \Delta d \Vert ^{1- ( \frac{1}{2p_{1}}+\frac{1}{2p_{2}} +\frac{1}{2p_{3}} )}_{L^{2}} \Vert \Delta u \Vert _{L^{2}} \\ & \leq C \Vert \nabla d \Vert ^{2}_{L^{2\vec{p}}} \Vert \nabla \Delta d \Vert ^{\frac{1}{p_{1}}+\frac{1}{p_{2}} +\frac{1}{p_{3}}}_{L^{2}} \Vert \Delta d \Vert ^{2- (\frac{1}{p_{1}}+\frac{1}{p_{2}} + \frac{1}{p_{3}} )}_{L^{2}}+\frac{1}{8} \Vert \nabla S \Vert _{L^{2}}^{2} \\ & \leq C \Vert \Delta d \Vert _{L^{2}}^{2} \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}+ \frac{1}{8} \Vert \nabla \Delta d \Vert _{L^{2}}^{2}+\frac{1}{8} \Vert \nabla S \Vert _{L^{2}}^{2}, \end{aligned}$$
(2.10)

where \(q= \frac{2}{2- (\frac{1}{p_{1}}+\frac{1}{p_{2}}+\frac{1}{p_{3}} )}\), \(2< p_{i}\leq \infty \).

Inserting (2.9) and (2.10) into (2.7), it follows that

$$ \begin{aligned} \frac{d}{d t} \Vert S \Vert _{L^{2}}^{2}+ \Vert \nabla S \Vert _{L^{2}}^{2}\leq C \bigl\Vert \lambda _{2}^{+} \bigr\Vert ^{q}_{L^{\vec{p}}} \Vert S \Vert ^{2}_{L^{2}}+C \Vert \Delta d \Vert _{L^{2}}^{2} \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}+\frac{1}{8} \Vert \nabla \Delta d \Vert _{L^{2}}^{2}. \end{aligned} $$
(2.11)

For the estimate of Δd, taking Δ on the second equation of (1.1), multiplying by Δd, and integrating over \(\mathbb{R}^{3}\), one obtains

$$ \begin{aligned} \frac{1}{2} \frac{d}{d t} \Vert \Delta d \Vert _{L^{2}}^{2}+ \Vert \nabla \Delta d \Vert _{L^{2}}^{2}&=- \int _{\mathbb{R}^{3}} \Delta \bigl((u \cdot \nabla ) d\bigr) \cdot \Delta d\, d x+ \int _{\mathbb{R}^{3}} \Delta \bigl( \vert \nabla d \vert ^{2} d \bigr) \cdot \Delta d \, d x \\ &\stackrel{\mathrm{def}}{=} J_{1}+J_{2}, \end{aligned} $$
(2.12)

where

$$ J_{1} =- \int _{\mathbb{R}^{3}} \partial _{l l} (u_{i} \partial _{i} d_{k} ) \partial _{j j} d_{k} \,dx ,\qquad J_{2}= \int _{\mathbb{R}^{3}} \Delta \bigl( \vert \nabla d \vert ^{2} d \bigr) \cdot \Delta d\, d x. $$
(2.13)

We now obtain the estimate of the term \(J_{1}\). Since

$$ \int _{\mathbb{R}^{3}}(u \cdot \nabla \Delta d) \cdot \Delta d\, d x= \frac{1}{2} \int _{\mathbb{R}^{3}}(u \cdot \nabla ) \bigl( \vert \Delta d \vert ^{2} \bigr) \,dx =0, $$

it follows from Lemma 1.3 and Young’s inequality that

$$ \begin{aligned} J_{1} & = - \int _{\mathbb{R}^{3}} \partial _{l l} (u_{i} \partial _{i} d_{k} ) \partial _{j j} d_{k} \,dx \\ & = \int _{\mathbb{R}^{3}} \partial _{l} (u_{i} \partial _{i} d_{k} ) \partial _{l} \partial _{j j} d_{k} \,dx \\ & = \int _{\mathbb{R}^{3}} \partial _{l} u_{i} \partial _{i} d_{k} \partial _{l} \partial _{j j} d_{k} \,dx + \int _{\mathbb{R}^{3}} u_{i} \partial _{i} \partial _{l} d_{k} \partial _{l} \partial _{j j} d_{k} \,dx \\ & = \int _{\mathbb{R}^{3}} \partial _{l} u_{i} \partial _{i} d_{k} \partial _{l} \partial _{j j} d_{k} \,dx - \int _{\mathbb{R}^{3}} \partial _{l} u_{i} \partial _{l} \partial _{l} d_{k} \partial _{j j} d_{k} \,dx \\ & \leq C \int _{\mathbb{R}^{3}} \vert \nabla u \vert \vert \nabla d \vert \vert \nabla \Delta d \vert \,dx \\ & \leq C \Vert \nabla d \Vert _{L^{2\vec{p}}} \Vert \nabla u \Vert _{L_{1}^{\frac{2 p_{1}}{p_{1}-1}}L_{2}^{ \frac{2 p_{2}}{p_{2}-1}}L_{3}^{\frac{2 p_{3}}{p_{3}-1}}} \Vert \nabla \Delta d \Vert _{L^{2}} \\ & \leq C \Vert \nabla d \Vert _{L^{2\vec{p}}} \bigl\Vert \nabla ^{2} u \bigr\Vert ^{\frac{1}{2p_{1}}+\frac{1}{2p_{2}} +\frac{1}{2p_{3}}}_{L^{2}} \Vert \nabla u \Vert ^{1- (\frac{1}{2p_{1}}+ \frac{1}{2p_{2}} +\frac{1}{2p_{3}} )}_{L^{2}} \Vert \nabla \Delta d \Vert _{L^{2}} \\ & \leq C \Vert S \Vert _{L^{2}}^{2} \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}+\frac{1}{8} \Vert \nabla \Delta d \Vert _{L^{2}}^{2}+\frac{1}{4} \Vert \nabla S \Vert _{L^{2}}^{2}, \end{aligned} $$
(2.14)

where \(q= \frac{2}{2- (\frac{1}{p_{1}}+\frac{1}{p_{2}}+\frac{1}{p_{3}} )}\), \(2< p_{i}\leq \infty \).

For the term \(J_{2}\), after integration by parts, by using the Young inequality and Lemma 1.3, and the fact \(|d| = 1\), one has

$$ \begin{aligned} J_{2} & \leq \biggl\vert \int _{\mathbb{R}^{3}} \Delta \bigl( \vert \nabla d \vert ^{2}d \bigr) \cdot \Delta d\, d x \biggr\vert = \biggl\vert \int _{\mathbb{R}^{3}} \nabla \bigl( \vert \nabla d \vert ^{2}d \bigr) \cdot \nabla \Delta d\, d x \biggr\vert \\ & \leq \biggl\vert \int _{\mathbb{R}^{3}} \nabla d \vert \nabla d \vert ^{2} \cdot \nabla \Delta d\, d x \biggr\vert + \biggl\vert \int _{\mathbb{R}^{3}} d \nabla \bigl( \vert \nabla d \vert ^{2} \bigr) \cdot \nabla \Delta d \, d x \biggr\vert \\ & \leq \biggl\vert \int _{\mathbb{R}^{3}} \nabla \bigl(\nabla d \vert \nabla d \vert ^{2} \bigr) \cdot \Delta d \, d x \biggr\vert + \biggl\vert \int _{\mathbb{R}^{3}} d \nabla \bigl( \vert \nabla d \vert ^{2} \bigr) \cdot \nabla \Delta d\, d x \biggr\vert \\ & \leq C \biggl\vert \int _{\mathbb{R}^{3}} \vert \nabla d \vert ^{2} \bigl\vert \nabla ^{2} d \bigr\vert ^{2} \,dx \biggr\vert +C \biggl\vert \int _{ \mathbb{R}^{3}} \vert \nabla d \vert \biggr\vert \nabla ^{2} d \bigl\vert \vert \nabla \Delta d \vert \,dx \bigr\vert \\ & \leq C \Vert \nabla d \Vert ^{2}_{L^{2\vec{p}}} \bigl\Vert \nabla ^{2} d \bigr\Vert ^{2}_{L_{1}^{\frac{2 p_{1}}{p_{1}-1}}L_{2}^{ \frac{2 p_{2}}{p_{2}-1}}L_{3}^{\frac{2 p_{3}}{p_{3}-1}}}+ \frac{1}{8} \Vert \nabla \Delta d \Vert ^{2}_{L^{2}} \\ & \leq C \Vert \nabla d \Vert ^{2}_{L^{2\vec{p}}} \Vert \nabla \Delta d \Vert ^{\frac{1}{p_{1}}+\frac{1}{p_{2}} +\frac{1}{p_{3}}}_{L^{2}} \Vert \Delta d \Vert ^{2- (\frac{1}{p_{1}}+\frac{1}{p_{2}} + \frac{1}{p_{3}} )}_{L^{2}}+\frac{1}{8} \Vert \nabla \Delta d \Vert ^{2}_{L^{2}} \\ & \leq C \Vert \Delta d \Vert _{L^{2}}^{2} \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}+ \frac{1}{4} \Vert \nabla \Delta d \Vert _{L^{2}}^{2}. \end{aligned} $$
(2.15)

Combining (2.12) and (2.14) with (2.15), we obtain

$$ \begin{aligned} \frac{d}{d t} \Vert \Delta d \Vert _{L^{2}}^{2}+ \Vert \nabla \Delta d \Vert _{L^{2}}^{2} \leq C \bigl( \Vert S \Vert _{L^{2}}^{2}+ \Vert \Delta d \Vert _{L^{2}}^{2} \bigr) \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}+ \frac{1}{4} \Vert \nabla S \Vert _{L^{2}}^{2}. \end{aligned} $$
(2.16)

Adding (2.11) and (2.16) together and using Lemma 1.4, we arrive at

$$ \begin{aligned} &\frac{d}{d t} \bigl( \Vert S \Vert _{L^{2}}^{2}+ \Vert \Delta d \Vert _{L^{2}}^{2} \bigr)+\frac{1}{2} \Vert \nabla S \Vert _{L^{2}}^{2}+\frac{1}{2} \Vert \nabla \Delta d \Vert _{L^{2}}^{2} \\ &\quad \leq C \bigl\Vert \lambda _{2}^{+} \bigr\Vert ^{q}_{L^{\vec{p}}} \Vert S \Vert ^{2}_{L^{2}}+C \bigl( \Vert S \Vert _{L^{2}}^{2}+ \Vert \Delta d \Vert _{L^{2}}^{2} \bigr) \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q} \\ &\quad \leq C \frac{ \Vert \lambda _{2}^{+} \Vert ^{q}_{L^{\vec{p}}}+ \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}}{1+\ln (e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} )} \bigl(1+\ln \bigl(e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} \bigr) \bigr) \\ &\quad \quad {}\cdot\bigl( \Vert S \Vert _{L^{2}}^{2}+ \Vert \Delta d \Vert _{L^{2}}^{2} \bigr) \\ &\quad \leq C \frac{ \Vert \lambda _{2}^{+} \Vert ^{q}_{L^{\vec{p}}}+ \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}}{1+\ln (e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} )} \\ &\quad \quad {}\cdot\bigl(1+\ln \bigl(e+ \Vert u \Vert ^{2}_{L^{2}}+ \Vert S \Vert ^{2}_{L^{2}}+ \Vert \nabla d \Vert ^{2}_{L^{2}}+ \Vert \Delta d \Vert ^{2}_{L^{2}}+C \bigr) \bigr) \\ &\quad \quad{} \cdot \bigl( \Vert S \Vert _{L^{2}}^{2}+ \Vert \Delta d \Vert _{L^{2}}^{2} \bigr). \end{aligned} $$
(2.17)

Combining the basic energy estimates and (2.17), we see that

$$ \begin{aligned} &\frac{d}{dt} \bigl(1+\ln \bigl(e+ \Vert u \Vert ^{2}_{L^{2}}+ \Vert S \Vert ^{2}_{L^{2}}+ \Vert \nabla d \Vert ^{2}_{L^{2}}+ \Vert \Delta d \Vert ^{2}_{L^{2}}+C \bigr) \bigr) \\ &\quad \leq C \frac{ \Vert \lambda _{2}^{+} \Vert ^{q}_{L^{\vec{p}}}+ \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}}{1+\ln (e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} )} \\ &\quad \quad {}\cdot\bigl(1+\ln \bigl(e+ \Vert u \Vert ^{2}_{L^{2}}+ \Vert S \Vert ^{2}_{L^{2}}+ \Vert \nabla d \Vert ^{2}_{L^{2}}+ \Vert \Delta d \Vert ^{2}_{L^{2}}+C \bigr) \bigr), \end{aligned} $$
(2.18)

which together with the Grönwall inequality leads to

$$ \begin{aligned} &\sup_{0\leq t\leq T_{*}}\ln \bigl(e+ \Vert u \Vert ^{2}_{L^{2}}+ \Vert S \Vert ^{2}_{L^{2}}+ \Vert \nabla d \Vert ^{2}_{L^{2}}+ \Vert \Delta d \Vert ^{2}_{L^{2}}+C \bigr) \\ &\quad \leq \bigl(1+\ln \bigl(e+ \Vert u_{0} \Vert ^{2}_{L^{2}}+ \Vert S_{0} \Vert ^{2}_{L^{2}}+ \Vert \nabla d_{0} \Vert ^{2}_{L^{2}}+ \Vert \Delta d_{0} \Vert ^{2}_{L^{2}}+C \bigr) \bigr) \\ &\quad \quad {}\cdot\exp C \int ^{T_{*}}_{0} \frac{ \Vert \lambda _{2}^{+} \Vert ^{q}_{L^{\vec{p}}}+ \Vert \nabla d \Vert _{L^{2 \vec{p}}}^{2q}}{1+\ln (e+ \Vert u \Vert _{L^{\vec{s}}}+ \Vert \nabla d \Vert _{L^{\vec{s}}} )}\,dt , \end{aligned} $$
(2.19)

which implies that

$$ u, \nabla d\in L^{\infty}\bigl(0,T_{*};H^{1} \bigl(\mathbb{R}^{3}\bigr)\bigr)\cap L^{2} \bigl(0,T_{*};H^{2}\bigl( \mathbb{R}^{3} \bigr)\bigr). $$
(2.20)

To estimate

$$ \int _{0}^{T_{*}} \bigl( \bigl\Vert \omega ( t) \bigr\Vert _{L^{\infty}}+ \bigl\Vert \nabla d( t) \bigr\Vert _{L^{ \infty}}^{2} \bigr) \,dt \leq C, $$

let us establish a higher-order estimate for \((u,\nabla d)\). Applying Δ and Δ to the first equation of (1.1) and the second equation of (1.2), respectively, and then multiplying the resulting equations by Δu and \(\nabla \Delta d\), we have

$$ \begin{aligned} & \frac{1}{2} \frac{{d}}{ {d} t} \bigl( \Vert \Delta u \Vert _{L^{2}}^{2}+ \Vert \nabla \Delta d \Vert _{L^{2}}^{2} \bigr)+ \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+ \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2} \\ &\quad = - \int _{\mathbb{R}^{3}} \Delta (u \cdot \nabla u) \cdot \Delta u \, { d} x- \int _{\mathbb{R}^{3}} \Delta (\nabla d \cdot \Delta d ) \cdot \Delta u \, { d} x \\ &\quad \quad{} - \int _{\mathbb{R}^{3}} \nabla \Delta (u \cdot \nabla d) \cdot \nabla \Delta d \, { d} x+ \int _{\mathbb{R}^{3}} \nabla \Delta \bigl( \vert \nabla d \vert ^{2} d \bigr) \cdot \nabla \Delta d \, { d} x \\ &\quad \stackrel{\mathrm{def}}{=} K_{1}+K_{2}+K_{3}+K_{4}, \end{aligned} $$
(2.21)

where

$$ K_{1}=- \int _{\mathbb{R}^{3}} \Delta (u \cdot \nabla u) \cdot \Delta u \, { d} x,\qquad K_{2}=- \int _{\mathbb{R}^{3}} \Delta (\nabla d \cdot \Delta d ) \cdot \Delta u \, { d} x $$
(2.22)

and

$$ K_{3}=- \int _{\mathbb{R}^{3}} \nabla \Delta (u \cdot \nabla d) \cdot \nabla \Delta d \, { d} x,\qquad K_{4}= \int _{\mathbb{R}^{3}} \nabla \Delta \bigl( \vert \nabla d \vert ^{2} d \bigr) \cdot \nabla \Delta d \, { d} x. $$
(2.23)

Noting the fact that

$$ u, \nabla d\in L^{\infty}\bigl(0,T_{*};H^{1} \bigl(\mathbb{R}^{3}\bigr)\bigr)\cap L^{2} \bigl(0,T_{*};H^{2}\bigl( \mathbb{R}^{3} \bigr)\bigr), $$

we shall establish the bounds of \(K_{1}\), \(K_{2}\), \(K_{3}\), and \(K_{4}\). By integrating by parts, Hölder’s inequality, and Young’s inequality, we can estimate \(K_{1}\) as

$$ \begin{aligned} K_{1} & \leq \epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+C \bigl( \bigl\Vert (\nabla u \cdot \nabla ) u \bigr\Vert _{L^{2}}^{2}+ \bigl\Vert (u \cdot \nabla ) \nabla u \bigr\Vert _{L^{2}}^{2} \bigr) \\ & \leq \epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+C \bigl( \Vert \nabla u \Vert _{L^{4}}^{4}+ \Vert u \Vert _{L^{6}}^{2} \Vert \Delta u \Vert _{L^{3}}^{2} \bigr) \\ & \leq \epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+C \bigl( \Vert \nabla u \Vert _{L^{2}}^{ \frac{5}{2}} \Vert \nabla \Delta u \Vert _{L^{2}}^{\frac{3}{2}}+ \Vert \nabla u \Vert _{L^{2}}^{2} \Vert \Delta u \Vert _{L^{2}} \Vert \nabla \Delta u \Vert _{L^{2}} \bigr) \\ & \leq \epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+C \Vert \nabla u \Vert ^{10}_{L^{2}}+ \epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+C \Vert \nabla u \Vert _{L^{2}}^{4} \Vert \Delta u \Vert ^{2}_{L^{2}}+\epsilon \Vert \nabla \Delta u \Vert ^{2}_{L^{2}} \\ & \leq 3\epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+C \bigl( \Vert \Delta u \Vert _{L^{2}}^{2}+1 \bigr). \end{aligned} $$
(2.24)

For the term \(K_{2}\), by applying the Hölder inequality, the interpolation inequality, and the Young inequality, we obtain

$$ \begin{aligned} K_{2} & = \int _{\mathbb{R}^{3}} \nabla (\nabla d \cdot \Delta d ) \nabla \Delta u \,dx \\ & \leq \bigl( \Vert \nabla d \Vert _{L^{4}} \Vert \nabla \Delta d \Vert _{L^{4}}+ \Vert \Delta d \Vert _{L^{4}} \bigl\Vert \nabla ^{2} d \bigr\Vert _{L^{4}} \bigr) \Vert \nabla \Delta u \Vert _{L^{2}} \\ & \leq C \bigl( \Vert \Delta d \Vert ^{\frac{3}{4}}_{L^{2}} \Vert \nabla \Delta d \Vert _{L^{4}}+ \Vert \Delta d \Vert _{L^{4}} \bigl\Vert \nabla ^{2} d \bigr\Vert _{L^{4}} \bigr) \Vert \nabla \Delta u \Vert _{L^{2}} \\ & \leq C \Vert \nabla \Delta d \Vert _{L^{4}}^{2}+C \bigl\Vert \nabla ^{2} d \bigr\Vert _{L^{4}}^{4}+ \epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2} \\ & \leq C \Vert \nabla \Delta d \Vert _{L^{2}}^{\frac{1}{2}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{\frac{3}{2}}+C \Vert \Delta d \Vert _{L^{2}}^{\frac{5}{2}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{\frac{3}{2}}+\epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2} \\ &\leq C \bigl( \Vert \nabla \Delta d \Vert _{L^{2}}^{2}+1 \bigr)+\epsilon \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2}+\epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}. \end{aligned} $$
(2.25)

Similarly, for the terms \(K_{3}\) and \(K_{4}\), we discover that

$$ \begin{aligned} K_{3}& = \int _{\mathbb{R}^{3}} \Delta \bigl((u \cdot \nabla ) d\bigr) \cdot \Delta ^{2} d\, d x \\ & \leq \epsilon \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2}+C \bigl( \bigl\Vert ( \Delta u \cdot \nabla ) d \bigr\Vert _{L^{2}}^{2}+ \bigl\Vert (\nabla u \cdot \nabla ) \nabla d \bigr\Vert _{L^{2}}^{2}+ \bigl\Vert (u \cdot \nabla ) \Delta d \bigr\Vert _{L^{2}}^{2} \bigr) \\ & \leq \epsilon \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2}+C \bigl( \Vert \Delta u \Vert _{L^{3}}^{2} \Vert \nabla d \Vert _{L^{6}}^{2}+ \Vert \nabla u \Vert _{L^{4}}^{2} \Vert \Delta d \Vert _{L^{4}}^{2}+ \Vert u \Vert _{L^{6}}^{2} \Vert \nabla \Delta d \Vert _{L^{3}}^{2} \bigr) \\ & \leq \epsilon \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2}+C \bigl( \Vert \Delta u \Vert _{L^{2}} \Vert \nabla \Delta u \Vert _{L^{2}}+ \Vert \nabla u \Vert _{L^{2}}^{ \frac{5}{4}} \Vert \nabla \Delta u \Vert _{L^{2}}^{\frac{3}{4}} \Vert \Delta d \Vert _{L^{2}}^{ \frac{5}{4}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{\frac{3}{4}} \\ &\quad {}+ \Vert \nabla \Delta d \Vert _{L^{2}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}} \bigr) \\ & \leq 2\epsilon \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2}+\epsilon \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+C \bigl( \Vert \Delta u \Vert _{L^{2}}^{2}+ \Vert \nabla \Delta d \Vert _{L^{2}}^{2}+1 \bigr) \end{aligned} $$
(2.26)

and

$$\begin{aligned} K_{4} &=- \int _{\mathbb{R}^{3}} \Delta \bigl( \vert \nabla d \vert ^{2} d \bigr) \Delta ^{2} d \, dx \\ &= \int _{\mathbb{R}^{3}} \bigl[\Delta \bigl( \vert \nabla d \vert ^{2} \bigr) d+2 \nabla \vert \nabla d \vert ^{2} \nabla d+ \vert \nabla d \vert ^{2} \Delta d \bigr] \Delta ^{2} d \\ & \leq C \bigl( \Vert \Delta d \Delta d \Vert _{L^{2}}+ \Vert \nabla d \nabla \Delta d \Vert _{L^{2}}+ \Vert \nabla d \nabla d \Delta d \Vert _{L^{2}}+ \Vert d \Delta d \Delta d \Vert _{L^{2}} \bigr) \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}} \\ & \leq C \bigl( \Vert \Delta d \Vert _{L^{4}}^{2}+ \Vert \nabla d \Vert _{L^{4}} \Vert \nabla \Delta d \Vert _{L^{4}} \bigr) \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}} \\ & \leq C \Vert \Delta d \Vert _{L^{2}}^{\frac{5}{4}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{\frac{3}{4}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}+C \Vert \nabla \Delta d \Vert _{L^{2}}^{\frac{1}{4}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{\frac{3}{4}} \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}} \\ & \leq C \bigl( \Vert \nabla \Delta d \Vert _{L^{2}}^{2}+1 \bigr)+\epsilon \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2}. \end{aligned}$$
(2.27)

Inserting the above estimates (2.24)–(2.27) into (2.21), we conclude that

$$ \begin{aligned} & \frac{{d}}{ {d} t} \bigl( \Vert \Delta u \Vert _{L^{2}}^{2}+ \Vert \nabla \Delta d \Vert _{L^{2}}^{2} \bigr)+ \Vert \nabla \Delta u \Vert _{L^{2}}^{2}+ \bigl\Vert \Delta ^{2} d \bigr\Vert _{L^{2}}^{2} \\ &\quad \leq C \bigl( \Vert \Delta u \Vert _{L^{2}}^{2}+ \Vert \nabla \Delta d \Vert _{L^{2}}^{2}+1 \bigr), \end{aligned} $$
(2.28)

which leads to

$$ u, \nabla d\in L^{\infty}\bigl(0,T_{*};H^{2} \bigl(\mathbb{R}^{3}\bigr)\bigr)\cap L^{2} \bigl(0,T_{*};H^{3}\bigl( \mathbb{R}^{3} \bigr)\bigr). $$
(2.29)

Due to Sobolev embedding \(H^{2} (\mathbb{R}^{3} ) \rightarrow L^{\infty} ( \mathbb{R}^{3} )\), we have

$$ \int _{0}^{T_{*}} \bigl( \bigl\Vert \omega ( t) \bigr\Vert _{L^{\infty}}+ \bigl\Vert \nabla d( t) \bigr\Vert _{L^{ \infty}}^{2} \bigr) \,dt \leq C. $$

This completes the proof of Theorem 1.1. □

Availability of data and materials

There was no mention of data availability in this paper.

References

  1. Besov, O., II’in, V., Nikol’skii, S.: Integral Representations of Functions and Imbedding Theorems, Volume I. V.H. Winston & Sons, Washington (1978)

    Google Scholar 

  2. Cao, C., Qin, J., Titi, E.: Regularity criterion for solutions of three-dimensional turbulent channel flows. Commun. Partial Differ. Equ. 33(3), 419–428 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  3. Cao, C., Titi, E.: Regularity criteria for the three-dimensional Navier–Stokes equations. Indiana Univ. Math. J. 57(6), 2643–2661 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  4. Cao, C., Titi, E.: Global regularity criterion for the 3D Navier–Stokes equations involving one entry of the velocity gradient tensor. Arch. Ration. Mech. Anal. 202(3), 919–932 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  5. Chae, D.: On the spectral dynamics of the deformation tensor and new a priori estimates for the 3D Euler equations. Commun. Math. Phys. 263, 789–801 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  6. Chang, K., Ding, W., Ye, R.: Finite-time blow-up of the heat flow of harmonic maps from surfaces. J. Differ. Geom. 36, 507–515 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  7. Chemin, J.Y., Zhang, P.: On the critical one component regularity for 3-D Navier–Stokes system. Ann. Sci. Éc. Norm. Supér. 49(1), 131–167 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  8. Chemin, J.Y., Zhang, P., Zhang, Z.: On the critical one component regularity for 3-D Navier–Stokes system: general case. Arch. Ration. Mech. Anal. 224(3), 871–905 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  9. Chen, Q., Tan, Z., Wu, G.: LPS’s criterion for incompressible nematic liquid crystal flows. Acta Math. Sci. 34(4), 1072–1080 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  10. Da Veiga, B.H.: A new regularity class for the Navier–Stokes equations in \(\mathbb{R}^{n}\). Chin. Ann. Math., Ser. B 16(4), 407–412 (1995)

    MATH  Google Scholar 

  11. Ericksen, J.: Conservation laws for liquid. Trans. Soc. Rheol. 5(1), 23–34 (1961)

    Article  MathSciNet  Google Scholar 

  12. Escauriaza, L., Seregin, G., Šerák, V.: \(L_{3,\infty}\)-Solutions of the Navier–Stokes equations and backward uniqueness. Usp. Mat. Nauk 58(2(350)), 3–44 (2003)

    Google Scholar 

  13. Fan, J., Li, J.: Regularity criteria for the strong solutions to the Ericksen–Leslie system in \(\mathbb{R}^{3}\). J. Math. Anal. Appl. 425(2), 695–703 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  14. Gala, S., Ragusa, M.: A new regularity criterion for strong solutions to the Ericksen–Leslie system. Appl. Math. 43(1), 95–103 (2016)

    MathSciNet  MATH  Google Scholar 

  15. Galanti, B., Gibbon, J., Heritage, M.: Vorticity alignment results for the three-dimensional Euler and Navier–Stokes equations. Nonlinearity 10(6), 1675–1694 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  16. Guo, Z., Kucera, P., Skalák, Z.: Regularity criterion for solutions to the Navier–Stokes equations in the whole 3D space based on two vorticity components. J. Math. Anal. Appl. 458(1), 755–766 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  17. Hong, M.: Global existence of solutions of the simplified Ericksen–Leslie system in dimension two. Calc. Var. Partial Differ. Equ. 40(1), 15–36 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  18. Huang, T., Wang, C.: Blow up criterion for nematic liquid crystal flows. Commun. Partial Differ. Equ. 37(5), 875–884 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  19. Lee, J.: Scaling invariant blow-up criteria for simplified versions of Ericksen–Leslie system. Discrete Contin. Dyn. Syst., Ser. S 8(2), 381 (2015)

    MathSciNet  MATH  Google Scholar 

  20. Leslie, F.M.: Theory of flow phenomena in liquid crystals. Adv. Liq. Cryst. 4, 1–81 (1979)

    Article  Google Scholar 

  21. Li, X., Wang, D.: Global solution to the incompressible flow of liquid crystals. J. Differ. Equ. 252(1), 745–767 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  22. Lin, F.: Some analytical issues for elastic complex fluids. Commun. Pure Appl. Math. 65(7), 893–919 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  23. Lin, F., Lin, J., Wang, C.: Liquid crystal flows in two dimensions. Arch. Ration. Mech. Anal. 197(1), 297–336 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  24. Lin, F., Liu, C.: Existence of solutions for the Ericksen–Leslie system. Arch. Ration. Mech. Anal. 154(2), 135–156 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  25. Lin, F.H.: Nonlinear theory of defects in nematic liquid crystals: phase transition and flow phenomena. Commun. Pure Appl. Math. 42(6), 789–814 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  26. Liu, Q., Zhao, J.: Logarithmically improved blow-up criteria for the nematic liquid crystal flows. Nonlinear Anal., Real World Appl. 16, 178–190 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  27. Miller, E.: A regularity criterion for the Navier–Stokes equation involving only the middle eigenvalue of the strain tensor. Arch. Ration. Mech. Anal. 235(1), 99–139 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  28. Miller, E.: Navier–Stokes regularity criteria in sum spaces. Pure Appl. Anal. 3(3), 527–566 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  29. Neustupa, J., Penel, P.: Regularity of a weak solution to the Navier–Stokes equation in dependence on eigenvalues and eigenvectors of the rate of deformation tensor. In: Trends in Partial Differential Equations of Mathematical Physics, pp. 197–212. Birkhäuser, Basel (2005)

    Chapter  MATH  Google Scholar 

  30. Neustupa, J., Penel, P.: On regularity of a weak solution to the Navier–Stokes equation with generalized impermeability boundary conditions. Nonlinear Anal., Theory Methods Appl. 66(8), 1753–1769 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  31. Neustupa, J., Penel, P.: On regularity of a weak solution to the Navier–Stokes equations with the generalized Navier slip boundary conditions. Adv. Math. Phys. 2018, Article ID 4617020 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  32. Penel, P., Pokorny, M.: Some new regularity criteria for the Navier–Stokes equations containing gradient of the velocity. Appl. Math. 49(5), 483–493 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  33. Prodi, G.: Un teorema di unicita per le equazioni di Navier–Stokes. Ann. Mat. Pura Appl. 48(1), 173–182 (1959)

    Article  MathSciNet  MATH  Google Scholar 

  34. Serrin, J.: On the interior regularity of weak solutions of the Navier–Stokes equations. Arch. Ration. Mech. Anal. 9(1), 187–195 (1962)

    Article  MathSciNet  MATH  Google Scholar 

  35. Wang, C.: Heat flow of harmonic maps whose gradients belong to \(L^{n} _{x} L^{\infty} _{t} \). Arch. Ration. Mech. Anal. 188(2), 351–369 (2008)

    Article  MathSciNet  Google Scholar 

  36. Wu, F.: Conditional regularity for the 3D Navier–Stokes equations in terms of the middle eigenvalue of the strain tensor. Evol. Equ. Control Theory (2021). https://doi.org/10.3934/eect.2020078

    Article  MathSciNet  MATH  Google Scholar 

  37. Zhang, X.: A regularity criterion for the solutions of 3D Navier–Stokes equations. J. Math. Anal. Appl. 346(1), 336–339 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  38. Zhang, Z.: A Serrin-type regularity criterion for the Navier–Stokes equations via one velocity component. Commun. Pure Appl. Anal. 12(1), 117–124 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  39. Zhang, Z., Chen, Q.: Regularity criterion via two components of vorticity on weak solutions to the Navier–Stokes equations in \(\mathbb{R}^{3}\). J. Differ. Equ. 216(2), 470–481 (2005)

    Article  MATH  Google Scholar 

  40. Zhang, Z., Yao, Z., Li, P., et al.: Two new regularity criteria for the 3D Navier–Stokes equations via two entries of the velocity gradient tensor. Acta Appl. Math. 123(1), 43–52 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  41. Zhao, J.: BKM’s criterion for the 3D nematic liquid crystal flows via two velocity components and molecular orientations. Math. Methods Appl. Sci. 40(4), 871–882 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  42. Zheng, X.: A regularity criterion for the tridimensional Navier–Stokes equations in term of one velocity component. J. Differ. Equ. 256(1), 283–309 (2014)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Both authors would like to thank heartily the anonymous referees for their invaluable comments that helped to improve this paper’s quality, thank the editors for their hard work on the publication of the paper, and thank Professor Daomin Cao and Professor Qiuyi Dai for their useful guidance on nonlinear PDE theory and helpful comments on this work.

Funding

Z. Chen was supported by the China Postdoctoral Science Foundation (Grant: No. 2021M690773) and F. Wu was supported by the Jiangxi Provincial Natural Science Foundation (Grant: No. 20224BAB211003).

Author information

Authors and Affiliations

Authors

Contributions

F. Wu posed the problem and wrote the Sect. 1. Z. Chen wrote the Sect. 2. Both authors discussed and checked all the details. Both authors contributed equally to this work.

Corresponding author

Correspondence to Fan Wu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, Z., Wu, F. Blow-up criteria of the simplified Ericksen–Leslie system. Bound Value Probl 2023, 41 (2023). https://doi.org/10.1186/s13661-023-01729-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13661-023-01729-y

MSC

Keywords