Skip to main content

Interior regularity criterion for incompressible Ericksen-Leslie system

Abstract

An interior regularity criterion of suitable weak solutions is formulated for the Ericksen-Leslie system of liquid crystals. Such a criterion is point-wise, with respect to some appropriate norm of velocity u and the gradient of d, and it can be viewed as a sort of simply sufficient condition on the local regularity of suitable weak solutions.

1 Introduction and main results

In this paper, we investigate the local regularity of weak solutions to the following 3D incompressible Ericksen-Leslie liquid crystal system:

$$\begin{aligned} &\partial_{t}u+(u\cdot\nabla)u-\Delta u+\nabla P=-\nabla\cdot(\nabla d\odot\nabla d), \end{aligned}$$
(1.1a)
$$\begin{aligned} & \nabla\cdot u=0, \end{aligned}$$
(1.1b)
$$\begin{aligned} &\partial_{t}d+(u\cdot\nabla) d =\Delta d-f(d), \end{aligned}$$
(1.1c)

with the initial boundary conditions

$$\begin{aligned} \begin{aligned} &(u, d) (x,t)|_{t=0}= \bigl(u_{0}(x), d_{0}(x) \bigr),\qquad \nabla\cdot u_{0}=0, \quad x\in\Omega, \\ &(u,d) (x,t)|_{x\in\partial\Omega}= \bigl(0, d_{0}(x) \bigr),\qquad u_{0}(x)\in H_{0}^{1}(\Omega ),\qquad d_{0}(x)\in H_{0}^{2}(\Omega), \end{aligned} \end{aligned}$$
(1.2)

where \(u, d, P\) denote the velocity of the fluid, the uniaxial molecular direction, and the pressure, respectively, the \(i,j\)th element of \(\nabla d\odot\nabla d\) is \(\partial_{i}d^{k}\partial_{j}d^{k}\), \(d_{0}(x)\) is a unit vector, \(\Omega\subset \mathbb {R}^{3}\) is a smooth domain. Additionally, \(f(d)=\nabla F(d)\), and \(F(d)=\frac{1}{\zeta^{2}}( \vert d \vert ^{2}-1)^{2}, \zeta\) is a small number, formally speaking, as \(\zeta\to0, d\) tends to a unit vector.

The dynamic flows of liquid crystals have been successfully described by the Ericksen-Leslie theory [14]. System (1.1a)-(1.1c) is a coupled system of the Navier-Stokes equations with a parabolic system. It is Leray [5] and Hopf [6] that established the global existence of weak solutions to the 3D Navier-Stokes; however, the regularity of the weak solutions is still an open problem. Since the regularity of weak solutions to the 3D Navier-Stokes equations is hard to get, some related conditions or criteria for the regularity of the weak solutions are considered, such as the well-known Serrin type criterion [7] and the Beale-Kato-Majda type criterion [8]. Furthermore, based on the suitable weak solutions, some point-wise sufficient regularity criteria were imposed in [912].

The global existence of suitable weak solutions to system (1.1a)-(1.1c) was established in [13, 14] by Lin and Liu; however, noticing that system (1.1a)-(1.1c) contains the 3D Navier-Stokes equations as a subsystem, the uniqueness and regularity of these weak solutions are not known. In this paper, we would extend some point-wise sufficient conditions, which guarantee the local regularity of weak solutions for 3D Navier-Stokes equations, to the Ericksen-Leslie system (1.1a)-(1.1c). We would like to mention that when \(f(d)\) in system (1.1a)-(1.1c) is replaced by \(- \vert \nabla d \vert ^{2}d\), the global existence of weak solutions to the resulting system in three dimensions has only been known under the additional assumption that \(d_{3}\geq0\) or small initial data (see [15, 16]). Without these conditions, the general existence of weak solutions is still open. However, the Serrin type criterion and the Beale-Kato-Majda type criterion still hold true even for a weak solution (if it exists) (see [17, 18]).

The suitable weak solution established in [14] can be stated as below.

Definition 1.1

Suitable weak solutions in \(\Omega\times(0, T)\subset\mathbb {R}^{3}\times(0,\infty)\)

A pair \((u, d)\) is called a suitable weak solution to system (1.1a)-(1.1c) and (1.2) in an open set \(\mathcal {O}\subset\mathbb {R}^{3}\times(0,\infty)\) (we set \(\mathcal {O}_{t}=\mathcal {O}\cap(\mathbb {R}^{3}\times\{t\} )\)), if it satisfies the following properties:

  • \((u, d)\) is a weak solution in the sense of distribution;

  • \(u\in L^{\infty}(0,T;L^{2}(\Omega ))\cap L^{2}(0,T;H^{1}(\Omega)), d\in L^{\infty}(0,T;H^{1}(\Omega))\cap L^{2}(0,T;H^{2}(\Omega))\), or generally, there exist constants \(E_{1}, E_{2}\), such that

    $$\begin{aligned} &\int_{\mathcal {O}_{t}} \bigl[ \vert u \vert ^{2}+ \vert \nabla d \vert ^{2}+F(d) \bigr]\,\mathrm{ d}x< E_{1},\\ &\int\!\int_{\mathcal {O}} \bigl[ \vert \nabla u \vert ^{2}+ \bigl\vert \Delta d-f(d) \bigr\vert ^{2}+F(d) \bigr]\,\mathrm{ d}x\,\mathrm{ d}t< E_{2}; \end{aligned}$$
  • for any \(\varphi\in C_{c}^{\infty}(\mathcal {O})\), more specifically, for any \(\varphi\in C_{c}^{\infty}(B(x_{0}, R)\times(t_{0}-R^{2}, t_{0}))\), the following generalized energy inequality holds

    $$\begin{aligned} &\int_{B(x_{0}, R)} \bigl( \vert u \vert ^{2}+ \vert \nabla d \vert ^{2} \bigr) \varphi\,\mathrm{ d}x+2 \int _{t_{0}-R^{2}}^{t} \int_{B(x_{0}, R)} \bigl( \vert \nabla u \vert ^{2}+ \bigl\vert \nabla^{2} d \bigr\vert ^{2} \bigr)\varphi\,\mathrm{ d}x \,\mathrm{ d}\tau \\ &\quad\leq \int_{t_{0}-R^{2}}^{t} \int_{B(x_{0}, R)} \bigl\{ \bigl( \vert u \vert ^{2}+ \vert \nabla d \vert ^{2} \bigr) (\varphi _{t}+\Delta \varphi)+ \bigl( \vert u \vert ^{2}+ \vert \nabla d \vert ^{2}+2P \bigr)u\cdot\nabla\varphi \bigr\} \,\mathrm{ d}x\,\mathrm{ d}\tau \\ &\qquad{}+2 \int_{t_{0}-R^{2}}^{t} \int_{B(x_{0}, R)} \bigl((u\cdot\nabla) d\nabla d\nabla \varphi-\nabla f(d):\nabla d\varphi \bigr)\,\mathrm{ d}x\,\mathrm{ d}\tau. \end{aligned}$$
    (1.3)

In the following, we can take \(Q((x_{0}, t_{0}), R)\equiv B(x_{0}, R)\times(t_{0}-R^{2}, t_{0})\), \(B(x_{0}, R)\equiv\{y\in\mathbb {R}^{3}| \vert y-x_{0} \vert < R\}, z_{0}\equiv(x_{0}, t_{0})\) for simplicity.

We now state our main result of this paper.

Theorem 1.2

Let \((u, d)\) be a suitable weak solution to liquid crystal system (1.1a)-(1.1c) in \(Q(z_{0}, R)\). The real numbers \(l\geq1\) and \(s\geq1\) satisfy

$$ \frac{1}{2}\geq\frac{3}{s}+\frac{2}{l}-\frac{3}{2}>\max \biggl\{ \frac{1}{2l}, \frac{1}{2}-\frac{1}{s}, \frac{1}{s}- \frac{1}{6} \biggr\} . $$

Then there is a positive number \(\varepsilon=\varepsilon(s,l)\), such that if

$$ M^{s, l}(z_{0}, R)=\frac{1}{R^{\kappa}} \int_{t_{0}-R^{2}}^{t_{0}} \biggl( \int_{B(x_{0}, R)} \vert u \vert ^{s}+ \vert \nabla d \vert ^{s}\,\mathrm{ d}x \biggr)^{\frac{l}{s}}\,\mathrm{ d}t< \varepsilon,\quad \kappa=\frac{3l}{s}+2-l, $$

then \(z_{0}\) is a regular point of \((u,\nabla d)\), i.e. \((u,\nabla d)\) is Hölder continuous in \(Q(z_{0},r)\), for some \(r\in(0, R]\).

Throughout this paper, we use c to denote a generic positive constant which can be different from line to line.

2 Preliminaries

As the preparation for proving Theorem 1.2, we first give two auxiliary lemmas.

Lemma 2.1

We have

$$ D(z_{0}, r; p)\leq c \biggl[\frac{r}{\rho}D(z_{0}, \rho; p)+ \biggl(\frac{\rho}{r} \biggr)^{2}C(z_{0}, \rho; u, \nabla d) \biggr], $$
(2.1)

where

$$ C(z_{0},r;u,\nabla d)=\frac{1}{r^{2}} \int_{Q(z_{0},r)} \bigl( \vert u \vert ^{3}+ \vert \nabla d \vert ^{3} \bigr) \,\mathrm{ d}z, \qquad D(z_{0},r;p)= \frac{1}{r^{2}} \int_{Q(z_{0},r)} \vert p \vert ^{\frac{3}{2}}\,\mathrm{ d}z. $$

Proof

Step 1. For (1.1a), we choose the test function \(w=\chi\nabla q\), for any \(\chi\in C_{c}^{\infty}((t_{0}-\rho^{2}, t_{0})), q\in C_{c}^{\infty}(B(x_{0}, \rho))\), then it yields

$$ \int_{Q(z_{0}, \rho)}-u\cdot\partial_{t}\chi\nabla q-(u\otimes u+ \nabla d\odot\nabla d):\chi\nabla^{2} q-u\cdot\chi\nabla\Delta q\,\mathrm{ d}z= \int _{Q(z_{0}, \rho)}p\chi\Delta q \,\mathrm{d}z. $$

It follows from \(\nabla\cdot u=0\) that

$$\begin{aligned} - \int_{Q(z_{0}, \rho)}p\chi\Delta q \,\mathrm{d}z= \int_{Q(z_{0}, \rho)}\chi (u\otimes u+\nabla d\odot\nabla d): \nabla^{2} q\,\mathrm{ d}z. \end{aligned}$$

Therefore, for a.e. \(t\in(t_{0}-\rho^{2}, t_{0})\), we have

$$ - \int_{B(x_{0}, \rho)}p\Delta q \,\mathrm{d}x= \int_{B(x_{0}, \rho)}(u\otimes u+\nabla d\odot\nabla d): \nabla^{2} q\,\mathrm{ d}x,\quad \forall q\in C_{c}^{\infty}\bigl(B(x_{0}, \rho) \bigr). $$
(2.2)

Step 2. Approximate p with \(p_{1}\) by confining q in \(W^{2, 3}(B(x_{0}, \rho))\).

Set \(p_{1}\in L^{\frac{3}{2}}(Q(z_{0}, \rho))\) such that, for a.e. \(t\in(t_{0}-\rho ^{2}, t_{0})\),

$$ - \int_{B(x_{0}, \rho)}p_{1}\Delta q \,\mathrm{d}x= \int_{B(x_{0}, \rho )}(u\otimes u+\nabla d\odot\nabla d): \nabla^{2} q\,\mathrm{ d}x, $$
(2.3)

for any \(q(\cdot, t)\in W^{2, 3}(B(x_{0}, \rho))\), and \(q(\cdot, t)=0 \text{ on } \partial B(x_{0}, \rho)\). The existence of \(p_{1}\) is established due to the Lax-Milgram theorem with appropriate approximating process on u and d (see [11]).

Next, choose \(q_{0}(\cdot, t)\in W^{2, 3}(B(x_{0}, \rho))\), such that, for a.e. \(t\in(t_{0}-\rho^{2}, t_{0})\),

$$\begin{aligned} \Delta q_{0}(\cdot, t)=- \bigl\vert p_{1}(\cdot, t) \bigr\vert ^{\frac{1}{2}} \operatorname{sgn}p_{1}(\cdot, t),\quad \text{in } B(x_{0}, \rho),\qquad q_{0}(\cdot, t)=0, \quad\text{on } \partial B(x_{0}, \rho). \end{aligned}$$

Then, by the Calderon-Zygmund inequality, it yields

$$\begin{aligned} \biggl( \int_{B(x_{0}, \rho)} \bigl\vert \nabla^{2}q_{0}( \cdot, t) \bigr\vert ^{3}\,\mathrm{ d}x \biggr)^{\frac{1}{3}}\leq c \biggl( \int_{B(x_{0}, \rho)} \bigl\vert p_{1}(\cdot, t) \bigr\vert ^{\frac{3}{2}}\,\mathrm{ d}x \biggr)^{\frac{1}{3}}, \quad\mbox{a.e. }t\in \bigl(t_{0}- \rho^{2}, t_{0} \bigr). \end{aligned}$$

Therefore, it follows from (2.3) and the Hölder inequality that

$$\begin{aligned} \int_{B(x_{0}, \rho)} \bigl\vert p_{1}(\cdot, t) \bigr\vert ^{\frac{3}{2}}\,\mathrm{ d}x &\leq c \biggl( \int_{B(x_{0}, \rho)} \vert u \vert ^{3}+ \vert \nabla d \vert ^{3}\,\mathrm{ d}x \biggr)^{\frac{2}{3}} \biggl( \int_{B(x_{0}, \rho)} \bigl\vert \nabla^{2}q \bigr\vert ^{3}\,\mathrm{ d}x \biggr)^{\frac{1}{3}} \\ &\leq c \biggl( \int_{B(x_{0}, \rho)} \vert u \vert ^{3}+ \vert \nabla d \vert ^{3}\,\mathrm{ d}x \biggr)^{\frac{2}{3}} \biggl( \int_{B(x_{0}, \rho)} \vert p_{1} \vert ^{\frac{3}{2}} \,\mathrm{ d}x \biggr)^{\frac{1}{3}}, \end{aligned}$$

which yields \(\int_{Q(z_{0}, \rho)} \vert p_{1}(\cdot, t) \vert ^{\frac{3}{2}}\,\mathrm{ d}z\leq c\rho^{2}C(z_{0}, \rho; u, \nabla d)\).

Step 3. Estimates for the remainder \(p-p_{1}\).

For a.e. \(t\in(t_{0}-\rho^{2}, t_{0})\), let \(p_{2}=p-p_{1}\), then from (2.2)-(2.3) one infers that

$$\begin{aligned} \Delta p_{2}(\cdot, t)=0,\quad \text{in } B(x_{0}, \rho). \end{aligned}$$

By the harmonic property, one can get

$$\begin{aligned} \frac{1}{r^{3}} \int_{Q(z_{0}, r)} \vert p_{2} \vert ^{\frac{3}{2}} \,\mathrm{ d}z\leq \frac{c}{\rho ^{3}} \int_{Q(z_{0}, \rho)} \vert p_{2} \vert ^{\frac{3}{2}} \,\mathrm{ d}z,\quad \forall r< \rho, \end{aligned}$$

while

$$ \int_{Q(z_{0}, \rho)} \vert p_{2} \vert ^{\frac{3}{2}} \,\mathrm{ d}z \leq \int_{Q(z_{0}, \rho)} \bigl( \vert p \vert ^{\frac{3}{2}}+ \vert p_{1} \vert ^{\frac{3}{2}} \bigr)\,\mathrm{ d}z\leq c\rho ^{2} \bigl(D(z_{0}, \rho; p)+C(z_{0}, \rho; u, \nabla d) \bigr). $$

Step 4. Estimates for p.

We have

$$\begin{aligned} D(z_{0}, r; p)&\leq c \biggl(\frac{1}{r^{2}} \int_{Q(z_{0}, r)} \vert p_{1} \vert ^{\frac{3}{2}} \,\mathrm{ d}z+ \frac{r}{\rho^{3}} \int_{Q(z_{0}, \rho)} \vert p_{2} \vert ^{\frac{3}{2}} \,\mathrm{ d}z \biggr) \\ &\leq c \biggl(\frac{\rho^{2}}{r^{2}}\frac{1}{\rho^{2}} \int_{Q(z_{0}, r)} \vert p_{1} \vert ^{\frac{3}{2}} \,\mathrm{ d}z+ \frac{r}{\rho}\frac{1}{\rho^{2}} \int_{Q(z_{0}, \rho)} \vert p_{2} \vert ^{\frac{3}{2}} \,\mathrm{ d}z \biggr) \\ &\leq c \biggl[\frac{\rho^{2}}{r^{2}}C(z_{0}, \rho; u, \nabla d)+ \frac{r}{\rho } \bigl(D(z_{0}, \rho; p)+C(z_{0}, \rho; u, \nabla d) \bigr) \biggr] \\ &\leq c \biggl[\frac{r}{\rho}D(z_{0}, \rho; p)+ \biggl( \frac{\rho}{r} \biggr)^{2}C(z_{0}, \rho; u, \nabla d) \biggr]. \end{aligned}$$

 □

We denote

$$\begin{aligned} &A(\rho)=\operatorname{ess}\sup_{{t_{0}}-\rho ^{2}< t< {t_{0}}}\frac{1}{\rho}\int_{B(x_{0},\rho )} \bigl( \bigl\vert u(t) \bigr\vert ^{2}+ \bigl\vert \nabla d(t) \bigr\vert ^{2} \bigr)\,\mathrm{ d}x, \\ & E(\rho)=\frac{1}{\rho}\int_{Q(z_{0},\rho)} \bigl( \vert \nabla u \vert ^{2}+ \bigl\vert \nabla^{2} d \bigr\vert ^{2} \bigr)\,\mathrm{ d}z,\qquad H( \rho)=\frac{1}{\rho^{3}} \int_{Q(z_{0},\rho )} \bigl( \vert u \vert ^{2}+ \vert \nabla d \vert ^{2} \bigr) \,\mathrm{ d}z. \end{aligned}$$

Lemma 2.2

Under the assumptions of Theorem 1.2, we have

$$C(\rho)\leq c\epsilon^{\frac{1}{q}} \bigl(E(\rho)+A(\rho)+1 \bigr), $$

where \(q=2l(\frac{3}{s}+\frac{2}{l}-\frac{3}{2})\), and \(q'=\frac{q}{q-1}\).

Proof

With the help of the Hölder and Sobolev embedding inequalities, one gets

$$\begin{aligned} \int_{B(x_{0}, \rho)} \vert v \vert ^{3}\,\mathrm{ d}x={}& \int_{B(x_{0}, \rho)} \vert v \vert ^{\lambda s+2\mu+6\gamma}\,\mathrm{ d}x \\ \leq {}&\biggl( \int_{B(x_{0}, \rho)} \vert v \vert ^{2}\,\mathrm{ d}x \biggr)^{\mu}\biggl( \int _{B(x_{0}, \rho)} \vert v \vert ^{s}\,\mathrm{ d}x \biggr)^{\lambda}\biggl( \int_{B(x_{0}, \rho)} \vert v \vert ^{6}\,\mathrm{ d}x \biggr)^{\gamma}\\ \leq{}& \frac{c}{2}\rho^{\mu}\biggl(\operatorname{ess}\sup_{t_{0}-\rho^{2}< t< t_{0}} \frac{1}{\rho}\int _{B(x_{0},\rho)} \vert v \vert ^{2}\,\mathrm{ d}x \biggr)^{\mu}\biggl( \int_{B(x_{0}, \rho)} \vert v \vert ^{s}\,\mathrm{ d}x \biggr)^{\lambda}\\ &{}\times \biggl( \int_{B(x_{0}, \rho)} \vert \nabla v \vert ^{2}+ \frac{1}{\rho ^{2}} \vert v \vert ^{2} \,\mathrm{ d}x \biggr)^{3\gamma}, \end{aligned}$$

where \(\lambda s+2\mu+6\gamma=3, \lambda+\mu+\gamma=1\). Substituting v by u and d, respectively, then one can get the summation

$$\begin{aligned} \int_{B(x_{0}, \rho)} \vert u \vert ^{3}+ \vert \nabla d \vert ^{3}\,\mathrm{ d}x \leq{}&c\rho^{\mu}A^{\mu}( \rho) \biggl( \int_{B(x_{0}, \rho)} \bigl( \vert u \vert ^{s}+ \vert \nabla d \vert ^{s} \bigr) \,\mathrm{ d}x \biggr)^{\lambda}\\ &{}\times \biggl( \int_{B(x_{0}, \rho)} \bigl( \vert \nabla u \vert ^{2}+ \bigl\vert \nabla^{2} d \bigr\vert ^{2} \bigr)+ \frac{1}{\rho^{2}} \bigl( \vert u \vert ^{2}+ \vert \nabla d \vert ^{2} \bigr)\,\mathrm{ d}x \biggr)^{3\gamma}. \end{aligned}$$

Therefore, by choosing appropriate parameters \(\lambda=\frac{1}{2s(\frac{3}{s}+\frac{2}{l}-\frac{3}{2})}\), \(\mu=\frac {\frac{3}{s}+\frac{3}{l}-2}{2(\frac{3}{s}+\frac{2}{l}-\frac{3}{2})}\), \(\gamma= \frac{\frac{2}{s}+\frac{1}{l}-1}{2(\frac{3}{s}+\frac{2}{l}-\frac{3}{2})}\), and integrating from \(t_{0}-\rho^{2}\) to \(t_{0}\) with the variable t, it follows from the Hölder and Young inequalities that

$$\begin{aligned} C(\rho)\leq{}&c\rho^{\mu-2} A^{\mu}(\rho){ \biggl( \int_{Q(z_{0}, \rho )} \bigl( \vert \nabla u \vert ^{2}+ \bigl\vert \nabla^{2} d \bigr\vert ^{2} \bigr)+ \frac{1}{\rho^{2}} \bigl( \vert u \vert ^{2}+ \vert \nabla d \vert ^{2} \bigr)\,\mathrm{ d}z \biggr)}^{\frac{1}{q'}} \\ &{}\times \biggl[ \int_{t_{0}-\rho^{2}}^{t_{0}} \biggl( \int_{B(x_{0}, \rho )} \bigl( \vert u \vert ^{s}+ \vert \nabla d \vert ^{s} \bigr) \,\mathrm{ d}x \biggr)^{\frac{l}{s}} \,\mathrm{ d}t \biggr]^{\frac{1}{q}} \\ \leq{}&c\rho^{\mu-2} A^{\mu}(\rho)\rho^{\frac{1}{q'}} \bigl(E(\rho)+H( \rho) \bigr)^{\frac{1}{q'}} \bigl(\rho^{\kappa}M^{s,l}(\rho) \bigr)^{\frac{1}{q}} \\ \leq{}&c A^{\mu}(\rho) \bigl(E(\rho)+H(\rho) \bigr)^{\frac{1}{q'}} \bigl(M^{s,l}(\rho) \bigr)^{\frac{1}{q}} \\ \leq{}&c \epsilon^{\frac{1}{q}} A^{\mu}(\rho) \bigl(E(\rho)+H(\rho) \bigr)^{\frac{1}{q'}} \\ \leq{}&c \epsilon^{\frac{1}{q}} \bigl(A^{\mu q}(\rho)+E(\rho)+H(\rho) \bigr) \\ \leq{}&c \epsilon^{\frac{1}{q}} \bigl(E(\rho)+A(\rho)+1 \bigr), \end{aligned}$$

where \(\kappa=\frac{3l}{s}+2-l\) as in Theorem 1.2, and in the last step, we used the fact that \(\mu q\leq1, H(\rho)\leq A(\rho)\). □

3 Proof of Theorem 1.2

Due to the induction argument as Proposition 2.6 in [10] or Lemma 2.2 in [19] (the parabolic version of the Campanato criterion), to get the desired consequence, it suffices to prove \(C(\theta^{k})+D(\theta^{k})< \epsilon_{0}\) for some small \(\epsilon_{0}\). Here θ is a small number, which will be chosen later.

From the generalized energy inequality, it is easy to check that, for \(\rho\in(0, R]\),

$$ A \biggl(\frac{\rho}{2} \biggr)+E \biggl(\frac{\rho}{2} \biggr)\leq c \bigl[C^{\frac{2}{3}}(\rho)+C(\rho)+D(\rho) \bigr]. $$

Denoting \(G(\rho)=A(\rho)+E(\rho)+D(\rho)\), due to Lemmas 2.1-2.2, and the fact that \(C(2\theta\rho)\leq\frac{1}{4\theta^{2}}C(\rho)\), we can get

$$\begin{aligned} G(\theta\rho)\leq{}& c \biggl[C^{\frac{2}{3}}(2\theta\rho)+C(2\theta \rho)+D(2 \theta \rho)+\theta D(\rho)+\frac{1}{\theta^{2}}C(\rho) \biggr] \\ \leq{}& c \biggl[\frac{1}{\theta^{\frac{4}{3}}}C^{\frac{2}{3}}(\rho)+ \frac{1}{\theta^{2}}C( \rho )+\theta D(\rho) \biggr] \\ \leq{}& c \biggl[\frac{\epsilon^{\frac{2}{3q}}}{\theta^{\frac{4}{3}}} \bigl(G(\rho)+1 \bigr)^{\frac{2}{3}}+ \frac{\epsilon^{\frac{1}{q}}}{\theta^{2}} \bigl(G(\rho)+1 \bigr)+\theta G(\rho) \biggr] \\ \leq{}& c \biggl[ \biggl(\theta+\frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}} \biggr)G(\rho)+ \frac {\epsilon^{\frac{1}{2q}}}{\theta^{2}} \biggr], \end{aligned}$$

where in the last step we have used \(\frac{\epsilon^{\frac{2}{3q}}}{\theta^{\frac{4}{3}}}(G(\rho)+1)^{\frac{2}{3}}\leq c[\epsilon^{\frac{1}{q}}+\frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}}(G(\rho)+1)]\). Now choosing θ and ϵ such that \(c\theta<\frac{1}{4}\) and \(c\frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}}<\frac{1}{4}\), then it yields \(G(\theta\rho)\leq\frac{1}{2}G(\rho)+c\frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}}\). Iterating the above process, we obtain \(G(\theta^{k} \rho)\leq\frac{1}{2^{k}}G(\rho)+c\frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}}\), therefore,

$$ D \bigl(\theta^{k}\rho \bigr)\leq\frac{1}{2^{k}}G( \rho)+c\frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}}. $$
(3.1)

For \(C(\theta^{k}\rho)\), by Lemma 2.2, we have

$$\begin{aligned} C \bigl(\theta^{k} \rho \bigr)\leq c \epsilon^{\frac{1}{q}} \bigl[ G \bigl(\theta^{k} \rho \bigr)+1 \bigr]\leq c \epsilon^{\frac{1}{q}} \biggl[\frac{1}{2^{k}}G(\rho)+ \frac{\epsilon^{\frac{1}{2q}}}{\theta ^{2}}+1 \biggr]\leq c \biggl[\frac{1}{2^{k}}G(\rho)+ \frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}} \biggr], \end{aligned}$$
(3.2)

where in the last step we use the fact that \(\epsilon^{\frac{1}{q}}\leq \frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}}\) for ϵ small enough. With these inequalities in hand, for fixed ρ and \(\epsilon_{0}\), we can choose \(k_{0}\) large enough such that \(c\frac{1}{2^{k_{0}}}G(\rho)<\frac{\epsilon_{0}}{4}\), and choose ϵ small enough, such that \(c\frac{\epsilon^{\frac{1}{2q}}}{\theta^{2}}<\frac{\epsilon_{0}}{4}\). With these prerequisites and (3.1)-(3.2), it follows that \(D(\theta^{k}\rho)+C(\theta^{k}\rho)<\epsilon_{0}\).

References

  1. Ericksen, JL: Conservation laws for liquid crystals. Trans. Soc. Rheol. 5, 23-34 (1961)

    Article  MathSciNet  Google Scholar 

  2. Ericksen, JL: Liquid crystals with variable degree of orientation. Arch. Ration. Mech. Anal. 113, 97-120 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  3. Leslie, FM: Some constitutive equations for liquid crystals. Arch. Ration. Mech. Anal. 28, 265-283 (1968)

    Article  MathSciNet  MATH  Google Scholar 

  4. Leslie, FM: Theory of flow phenomenum in liquid crystals. Adv. Liq. Cryst. 4, 1-81 (1979)

    Article  Google Scholar 

  5. Leray, J: Sur le mouvement d’un liquide visqueux emplissant l’espace. Acta Math. 63, 183-248 (1934)

    Article  MathSciNet  MATH  Google Scholar 

  6. Hopf, E: Uber die Aufangswertaufgabe für die hydrodynamischen Grundgleichungen. Math. Nachr. 4, 213-231 (1951)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  8. Beale, JT, Kato, T, Majda, A: Remarks on the breakdown of smooth solutions for the 3-D Euler equation. Commun. Math. Phys. 94, 61-66 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  9. Cafferelli, L, Kohn, R, Nirenberg, L: Partial regularity of suitable weak solutions of Navier-Stokes equations. Commun. Pure Appl. Math. 35, 771-831 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  10. Ladyzhenskaya, OA, Seregin, G: On partial regularity of suitable weak solutions to the three-dimensional Navier-Stokes equations. J. Math. Fluid Mech. 1, 356-387 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  11. Seregin, G: On the number of singular points of weak solutions to the Navier-Stokes equations. Commun. Pure Appl. Math. 54, 1019-1028 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  12. Zajaczkowski, W, Seregin, G: Sufficient condition of local regularity for the Navier-Stokes equations. J. Math. Sci. 143, 2869-2874 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  13. Lin, FH, Liu, C: Nonparabolic dissipative systems modeling the flow of liquid crystals. Commun. Pure Appl. Math. 48, 501-537 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  14. Lin, FH, Liu, C: Partial regularities of the nonlinear dissipative systems modeling the flow of liquid crystals. Discrete Contin. Dyn. Syst. 2, 1-23 (1996)

    Google Scholar 

  15. Lin, FH, Wang, CY: Global existence of weak solutions of the nematic liquid crystal flow in dimension three. Commun. Pure Appl. Math. 69, 1532-1571 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  16. Ma, WY, Gong, HJ, Li, JK: Global strong solutions to incompressible Ericksen-Leslie system in \(\mathbb {R}^{3}\). Nonlinear Anal. 109, 230-235 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  17. Huang, T, Wang, CY: Blow up criterion for nematic liquid crystal flows. Commun. Partial Differ. Equ. 37, 875-884 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  18. Hong, MC, Li, JK, Xin, ZP: Blow up critera of strong solutions to the Ericksen-Leslie system in \(\mathbb{R}^{3}\). Commun. Partial Differ. Equ. 39, 1284-1328 (2014)

    Article  MATH  Google Scholar 

  19. Escauriaza, L, Seregin, G, Sverak, V: \(L_{3,\infty}\)-solutions of the Navier-Stokes equations and backward uniqueness. Russ. Math. Surv. 58(2), 211-250 (2003)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors thank Dr. Huajun Gong and Dr. Jinkai Li for helpful discussions and suggestions. Ma is supported by Fostering Talents of NSFC-Henan Province (U1404102) and NSFC (No. 11501174, 11626090). Feng is supported by NSFC (No. 61401283, 11601342, 61472257), GDPSTPP (No. 2013B040403005) and (No. GCZX-A1409).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jiqiang Feng.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

Both authors contributed equally in writing this paper. They both read and approved the final manuscript.

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 distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ma, W., Feng, J. Interior regularity criterion for incompressible Ericksen-Leslie system. Bound Value Probl 2017, 62 (2017). https://doi.org/10.1186/s13661-017-0792-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13661-017-0792-y

MSC

Keywords