1 Introduction

1.1 Formulation in Eulerian coordinates

In the paper, we consider a 3-D viscous compressible fluid in a moving domain \(\Omega (t)\) with an upper free surface \(\Gamma (t)\) and a fixed bottom \(\Gamma _b\). This model can be expressed by the 3-D compressible Navier–Stokes equations(CNS)

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t \rho +\nabla \cdot (\rho \,u)=0\quad \text{ in }\quad \Omega (t),\\ \rho (\partial _t u+u\cdot \nabla u)+\nabla p-\nabla \cdot \mathbb {S}(u)=0\quad \text{ in }\quad \Omega (t),\\ \rho >0 \quad \text{ in }\quad \Omega (t),\quad \rho =0 \quad \text{ on }\quad \Gamma (t),\\ \mathcal {V}(\Gamma (t))=u\cdot n \quad \text{ on }\quad \Gamma (t),\\ (\mathbb {S}(u)-p\, \mathbb {I})n=0 \quad \text{ on }\quad \Gamma (t),\\ u|_{\Gamma _{b}}=0 \quad \text{ on }\quad \Gamma _{b},\\ (\rho ,u)|_{t=0}=(\rho _0,u_0) \quad \text{ in }\quad \Omega (0),\quad \Omega (0)=\Omega _0, \end{array}\right. } \end{aligned}$$
(1.1)

where \(\mathcal {V}(\Gamma (t))\) denotes the normal velocity of the free surface \(\Gamma (t)\), and \(n= n(t)\) is the exterior unit normal vector of \(\Gamma (t)\), the vector-field u denotes the Eulerian velocity field, \(\rho \) is the density of the fluid, and \(p=p(\rho )\) denotes the pressure function. The stress tensor \(\mathbb {S}(u)\) is defined by \(\mathbb {S}(u)=\mu \mathbb {D}(u)+\lambda (\nabla \cdot u)\mathbb {I}\), where the strain tensor \(\mathbb {D}(u)=\nabla u+\nabla u^{T}\) and dynamic viscosity \(\mu \) and bulk viscosity \(\nu \) are constants which satisfy the following relationship

$$\begin{aligned} \mu >0,~\lambda +\frac{2}{3}\mu \ge 0. \end{aligned}$$
(1.2)

The deviatoric (trace-free) part of the strain tensor \(\mathbb {D}(u)\) is then \(\mathbb {D}^0(u)=\mathbb {D}(u)-\frac{2}{3}{\mathrm{div}}\, u\, \mathbb {I}\). The viscous stress tensor in fluid is then given by \(\mathbb {S}(u)=\mu \,\mathbb {D}^0(u)+(\lambda +\frac{2}{3}\mu ) (\nabla \cdot u)\, \mathbb {I}\). Moreover, the pressure obeys the \(\gamma \)-law: \(p(\rho )=K\,\rho ^{\gamma }\), where K is an entropy constant and \(\gamma >1\) is the adiabatic gas exponent.

Equation (1.1)\(_1\) is the conservation of mass; Eq. (1.1)\(_2\) means the momentum conserved; the boundary condition (1.1)\(_3\) states that the pressure (and hence the density function) vanishes along the moving boundary \(\Gamma (t)\), which indicates that the vacuum state appears on the boundary \(\Gamma (t)\); the kinematic boundary condition (1.1)\(_4\) states that the vacuum boundary \(\Gamma (t)\) is moving with speed equal to the normal component of the fluid velocity; (1.1)\(_5\) means the fluid satisfies the kinetic boundary condition on the free boundary, (1.1)\(_6\) denotes the fluid is no-slip, no-penetrated on the fixed bottom boundary, and (1.1)\(_{7}\) are the initial conditions for the density, velocity, and domain.

In the paper, we assume the bottom \(\Gamma _{b}=\{y_3=b(y_h)\}\), and the moving domain \(\Omega (t)\) is horizontal periodic by setting \(\mathbb {T}^2_{y_h}\) with \(y_h:=(y_1, y_2)^T\) for \(\mathbb {T}=\mathbb {R}/\mathbb {Z}\).

1.2 Known results

Whether or not the appearance of vacuum state is related to the regularity of the solution to the compressible Navier–Stokes equations. Even if there is no vacuum in initial data, it cannot guarantee that vacuum state will be not generated in finite time in high-dimensional system. Whence initial data is close to a non-vacuum equilibrium in some functional space, Matsumura and Nishida [35, 36] proved global well-posedness of strong solutions to the 3-D CNS. Moreover, for the one dimensional case, Hoff and Smoller [17] proved that if the vacuum is not included at the beginning, no vacuum will occur in the future. Hoff and Serre [16] showed some physical weak solution does not have to depend continuously on their initial data when vacuum occurs.

When the initial density may vanish in open sets or on the (part of) boundary of the domain, the flow density may contain a vacuum, the equation of velocity becomes a strong degenerate hyperbolic-parabolic system and the degeneracy is one of major difficulties in study of regularity and the solution’s behavior, which is completely different from the non-vacuum case. For the existence of solutions for arbitrary data (the far field density is vacuum, that is, \(\rho (t, x)\rightarrow 0\) as \(x\rightarrow \infty \)), the major breakthrough is due to Lions [27] (also see [8, 14, 22]), where he obtains global existence of weak solutions, defined as solutions with finite energy with suitable \(\gamma \). Recently, Li and Xin  [26] and Vasseur and Yu [39] independently studied global existence of weak solutions of CNS whence the viscosities depend on the density and satisfy the Bresch and Desjardins relation [1]. Yet little is known on the structure of such weak solutions except for the case that some additional assumptions are added (see [15] for example). Indeed, the works of Xin etc. [24, 40] showed that the homogeneous Sobolev space is as crucial as studying the well-posedness for the Cauchy problem of compressible Navier–Stokes equations in the presence of a vacuum at far fields even locally in time. Adding some compatible condition on initial data, Cho and Kim [3] develop local well-posedness for strong solutions. Moreover, if initial energy is small, Huang et al. [18] showed the global existence of classical solutions but with large oscillations to CNS.

Physically, the vacuum problem appears extensively in the fundamental free boundary hydrodynamical setting: for instance, the evolving boundary of a viscous gaseous star, formation of shock waves, vortex sheets, as well as phase transitions.

For free boundary problem of the multi-dimensional Navier–Stokes equations with non-vacuum state, there are many results concerning its local and global strong solutions, one may refer to  [43, 44] and references therein.

But when the vacuum (in particular, the physical vacuum [28]) appears, the system becomes much harder. To understand the difficulty of the vacuum, we introduce the sound speed \(c:=\sqrt{p'(\rho )} (=\sqrt{K\gamma } \rho ^{\frac{\gamma -1}{2}}\) for polytropic gases) of the gas or fluid to describe the behavior of the smoothness of the density connecting to vacuum boundary. A vacuum boundary \(\Gamma (t)\) is called physical vacuum if there holds

$$\begin{aligned} -\infty<\frac{\partial c^2}{\partial n}<0 \end{aligned}$$
(1.3)

near the boundary \(\Gamma (t)\), where n is the outward unit normal to the free surface. The physical vacuum condition (1.3) implies the pressure (or the enthalpy \(c^2\)) accelerates the boundary in the normal direction. Thus, the initial physical vacuum condition (1.3) is equivalent to the requirement that

$$\begin{aligned} -\infty<\partial _n (\rho _0^{\gamma -1})<0 \quad \text {on} \quad \Gamma (0) \end{aligned}$$
(1.4)

which means that \(\rho ^{\gamma -1}_0(x)\sim dist(x,\Gamma (0))\), in other words, the initial sound speed \(c_0\) is only \(C^{\frac{1}{2}}\)-Hölder continuous near the interface \(\Gamma (0)\).

Due to lack of sufficient smoothness of the enthalpy \(c^2\) at the vacuum boundary, a rigorous understanding of the existence of physical vacuum states in compressible fluid dynamics has been a challenging problem, especially in multi-dimensional cases.

Recently, the local well-posedness theory for compressible Euler system with physical vacuum singularity was established in [4, 20, 21], and also global existence of smooth solutions for the physical vacuum free boundary problem of the 3-D spherically symmetric compressible Euler equations with damping was showed in [32]. And more recently, Hadzic and Jang [13] proved global nonlinear stability of the affine solutions to the compressible Euler system with physical vacuum, and Guo et al. [9] constructed an infinite dimensional family of collapsing solutions to the Euler-Poisson system whose density is in general space inhomogeneous and undergoes gravitational blowup along a prescribed space-time surface, with continuous mass absorption at the origin.

The study of vacuum is important in understanding viscous surface flows [30]. Very little is rigorously known about well-posedness theories available about free boundary problems of CNS with physical vacuum boundary. For 1-D problem, global regularity for weak solutions to the vacuum free boundary problem of CNS was obtained in [30], which is further generalized by Zeng [45] which established the strong solutions. For the multidimensional case, regularity results related to spherically symmetric motions. Guo et al.  [11] obtain a global weak solution to the problem with spherically symmetric motions and a jump density connects to vacuum. Later Liu [29] gives the existence of global solutions with small energy in spherically symmetric motions with the density connected to vacuum continuously or discontinuously. Anyway, almost all the well-posedness results require additional strongly singular compatibility conditions on initial data in terms of the acceleration for gaining more regularities of the velocity. Some related works can refer to [2, 6, 7, 12, 19, 25, 28, 30, 31, 37, 41, 42] and references therein.

The purpose of this paper is to establish the local well-posedness of the 3-D compressible Navier–Stokes equations (1.1) with physical vacuum boundary condition without any compatibility conditions, more precisely, we do not need any initial condition on the material derivative \(D_t u\) or its derivatives. For simplicity, we set \(\gamma =2\) and \(K=1\) in this paper.

As mentioned above, the main difficulty in obtaining regularity for the vacuum free boundary problem (1.1) lies in the degeneracy of the system near vacuum boundaries. In order to solve the system (1.1), the first idea is that we use Lagrangian coordinates to transform it to a system with fixed domain. One of advantage of Lagrangian coordinates is that the density \(\rho \) is solved directly by initial data and we only focus on the equation of velocity with coefficients related to Lagrangian coordinates.

The second and also key idea in our paper is that we use the conormal derivatives to obtain the high-order regularity. Because the density vanishes on the boundary, we can not close the energy estimates if we directly take normal derivatives to the system. So another choose is to take time derivatives in [4, 21] solving the compressible Euler equations with the physical vacuum, where high-order enough time-derivative estimates as long as spatial-derivative estimates allow us to close the energy estimates and then get the local-in-time existence of the strong solution of the Euler system. This high-order energy estimate in it is reasonable since the pressure term may cancel the singularity near the vacuum boundary when consider compatibility conditions on initial data in terms of the acceleration and its derivatives. However, this method may not work for the Navier–Stokes system (1.1) with constant viscosity coefficients. In fact, a strong singular compatibility conditions on initial data in terms of the acceleration and its derivatives will appear in it when we consider the high-order energy estimate, which is mainly due to the non-degenerate of the viscosity, but it seems very hard to find such kind of initial data satisfying these compatibility conditions. In order to get rid of this difficulty, our strategy is that we use conormal Sobolev space introduced in [34] to get the tangential regularity. Based on that, we multiply \(\partial _t v\) on the both sides of equations of v to get the estimates of \(\rho ^{\frac{1}{2}}\partial _t v\) which implies the two-order derivative on the normal direction. Form this, together with high-order tangential derivatives estimates, we get the \(W^{1,\infty }\) estimates of v and its conormal derivatives, which in turn guarantees the propagation of conormal regularities of the velocity.

1.3 Derivation of the system in Lagrangian coordinates and main result

In this paper, we consider the case that the upper boundary does not touch the bottom which means that

$$\begin{aligned} dist(\Gamma (0), \Gamma _b)>0. \end{aligned}$$

Take \(\Omega =\{x \in \mathbb {T}^2\times \mathbb {R}|\quad 0<x_3<1\}\) as the domain of equilibrium. Let \(\eta (t,x)\) be the position of the gas particle x at time t so that

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t \eta (t,x)=u(t,\eta (t,x)) \quad \text{ for }\quad t>0,\\&\eta (0,x)=\eta _0(x) \quad \text{ in }\quad {\Omega }. \end{aligned} \right. \end{aligned}$$
(1.5)

Here \(\eta _0\) is a diffeomorphism from \({\Omega }\) to the initial moving domain \(\Omega (0)\) which satisfies that \(\Gamma (0)=\eta _0(\{x_3=1\})\) and \(\Gamma _{b}=\eta _0(\{x_3=0\})\). It is easy to construct a invertible transform \(\eta _0\) which satisfies that

$$\begin{aligned} \det (D\eta _0) >0. \end{aligned}$$

Due to (1.5), we introduce the displacement which satisfies the following ODE

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t \xi (t,x)=u(t, x+\xi (t,x)) \quad \text{ for }\quad t>0,\\ \xi (0,x)=\xi _0(x):=\eta _0(x)-x \quad \text{ in }\quad {\Omega }. \end{array}\right. } \end{aligned}$$
(1.6)

We define the following Lagrangian quantities:

$$\begin{aligned} v(t,x)&:=u(t,\eta (t,x)),\quad \, f(t,x):=\rho (t,\eta (t,x)),\\ \mathcal {A}&:=[D\eta ]^{-1},\quad J:=\det (D\eta ), \quad \mathcal {N}:=J\mathcal {A}\,e_3. \end{aligned}$$

Then, the system (1.1) is reformulated in Lagrangian coordinates as follows

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\xi =v \quad \text{ in }\quad \Omega ,\\&\partial _t f+f\nabla _{\mathcal {A}}\cdot v=0 \quad \text{ in }\quad \Omega ,\\&f\partial _t v+\nabla _{\mathcal {A}} (f^2)-\nabla _{\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}} (v)=0 \quad \text{ in }\quad \Omega \end{aligned} \right. \end{aligned}$$
(1.7)

with boundary conditions

$$\begin{aligned} {\left\{ \begin{array}{ll} f=0 \quad \text{ on }\quad \Gamma ,\\ \mathbb {S}_{\mathcal {A}} (v)\, \mathcal {N}=0,\quad \text{ on }\quad \Gamma ,\\ v|_{x_3=0}=0 \end{array}\right. } \end{aligned}$$
(1.8)

and initial data

$$\begin{aligned} (\xi , f,v)|_{t=0}=(\xi _0, \rho _0,u_0). \end{aligned}$$
(1.9)

One may readily check from the definition of J that

$$\begin{aligned} \partial _t J=\nabla _{J\mathcal {A}}\cdot v, \end{aligned}$$

which together with the equation of f in (1.7) yields

$$\begin{aligned} \partial _t(fJ)=J\partial _t f+f\partial _t J =-Jf\,\nabla _{\mathcal {A}}\cdot v+fJ\nabla _{\mathcal {A}}\cdot v =0. \end{aligned}$$

Hence, we find

$$\begin{aligned} Jf(t,x)=(Jf)(0,x)=\det (D\eta _0)\rho _0(\eta _0), \end{aligned}$$
(1.10)

where \(\rho _0\) is a given initial density function. We are interested in the initial density \(\rho _0\) satisfying

$$\begin{aligned}&\rho _0(\eta _0)\det (D\eta _0)=\overline{\rho }(x)\quad \text {in} \quad \Omega , \end{aligned}$$
(1.11)
$$\begin{aligned}&C^{-1}\,d(x)\le \overline{\rho }(x)\le C\,d(x)\quad \text {in} \quad \Omega , \end{aligned}$$
(1.12)
$$\begin{aligned}&|\nabla \overline{\rho }|\le C,~|\overline{\rho }^{-1}\nabla _h^k \overline{\rho }|\le C_k\quad \text {in} \quad \Omega \end{aligned}$$
(1.13)

with some given function \(\overline{\rho }(x)\) (\(x\in \Omega \)), for any \(k \in \mathbb {N}\) with \(\nabla _h=(\partial _1,\partial _2)\), where d(x) is the distance function to the boundary \(\{x_3=1\}\).

Thus, it follows from (1.10) that

$$\begin{aligned} Jf=\overline{\rho }(x), \end{aligned}$$
(1.14)

which implies that

$$\begin{aligned} f= J^{-1}\,\overline{\rho },\quad q=f^2=J^{-2}\,\overline{\rho }^{2}. \end{aligned}$$
(1.15)

Remark 1.1

For any smooth subdomain \(\mathcal {O}\) of \(\Omega \), we know that \(\eta _0(\mathcal {O})\) is a subdomain of \(\Omega (0)\) if \(\eta _0\) is a diffeomorphism from \(\Omega \) to \(\Omega (0)\). Hence, by using change of variables, we get

$$\begin{aligned} \int _{\eta _0(\mathcal {O})} \rho _0(y) \,dy= \int _{\mathcal {O}} \rho _0(\eta _0)\,\det (D\eta _0) \,dx. \end{aligned}$$
(1.16)

Hence, the assumption (1.11) is equivalent to the mass conservation law

$$\begin{aligned} \int _{\eta _0(\mathcal {O})} \rho _0(y) \,dy= \int _{\mathcal {O}} \overline{\rho }\,dx \quad \forall \,\, \mathcal {O} \subset \Omega . \end{aligned}$$
(1.17)

Multiplying the both side of equation v by J, we obtain the equivalent form of the system (1.7)–(1.9) as follows

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t\xi =v \quad \text{ in }\quad \Omega ,\\ \overline{\rho }\partial _t v+\nabla _{J\mathcal {A}}(J^{-2}\overline{\rho }^2)-\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}(v)=0 \quad \text{ in }\quad \Omega ,\\ \mathbb {S}_{\mathcal {A}} (v)\,\mathcal {N}=0,\quad \text{ on }\quad \Gamma ,\\ v|_{x_3=0}=0,\\ \xi |_{t=0}=\xi _0,\quad \, v|_{t=0}=v_0 \quad \text{ in }\quad \Omega . \end{array}\right. } \end{aligned}$$
(1.18)

Next, we give some useful equations which we often use in what follows. Since \(\mathcal {A}[D\eta ]=I,\) one obtains that

$$\begin{aligned} \partial _t \mathcal {A}_i^k=-\mathcal {A}_i^s\partial _s v^r\mathcal {A}^{k}_{r},\quad \partial _l \mathcal {A}_i^k=-\mathcal {A}_i^s\partial _s\partial _l\eta ^r\mathcal {A}^k_r. \end{aligned}$$
(1.19)

Differentiating the Jacobian determinant, we get

$$\begin{aligned} \partial _t J=J\mathcal {A}^{s}_{r}\partial _s v^r,\quad \partial _l J=J\mathcal {A}^s_r\partial _s\partial _l\eta ^r. \end{aligned}$$
(1.20)

Moreover, the following Piola identity holds:

$$\begin{aligned} \partial _j(J\mathcal {A}^j_i)=0, \end{aligned}$$
(1.21)

for any \(i=1,2,3.\)

1.4 Main results

Before we state our main results, we give some definitions of functional spaces. First, define the operators:

(1.22)

Using \(Z^m\) to denote \(Z_3^{m_2}Z_h^{m_1}=Z_3^{m_2}Z_1^{m_{11}}Z_2^{m_{12}}\) with \(m_1=(m_{11}, m_{12})\) and |m| to denote \(|m|=|m_1|+m_2=m_{11}+m_{12}+m_2.\) Moreover, we use \(Z_3^{m_2}\) to denote \(\overline{\rho }^{m_2}\partial _{3}^{m_2}.\) By (1.11)–(1.13), it is easy to see

$$\begin{aligned}{}[\partial _3,Z^m]\sim Z^{m-1}\partial _3,~[Z_h,Z_3]\sim Z_3. \end{aligned}$$

We recall the following conormal Sobolev space introduced by Masmoudi and Rousset [34].

$$\begin{aligned} \Vert f\Vert _{X^N_{\alpha }}^2:=\sum _{|m|=0}^N\Vert \overline{\rho }^{\alpha } Z^m f\Vert _{L^2}^2,\quad \Vert f\Vert _{\dot{X}^N_\alpha }^2:=\sum _{ |m|=1}^N\Vert \overline{\rho }^{\alpha } Z^m f\Vert _{L^2}^2, \end{aligned}$$

where \(\alpha \in \mathbb {R}\). In particular, when \(\alpha =0\), we the spaces \(X^N_{\alpha }\) and \(\dot{X}^N_{\alpha }\) will be denoted by \(X^N\) and \(\dot{X}^N\) respectively for simplicity.

For \(T>0\), we define the energy space \(E_T\) as

with the instantaneous energy \(\mathcal {E}(t)\) (in terms to the velocity v)

and the dissipation \(\mathcal {D}(t)\)

Given \(\kappa >0\), we also introduce the space \(F_{\kappa }\) in terms to the flow map \(\eta \) as follows:

equipped with the norm

Now, we are in the position to state our main results.

Theorem 1.2

Under the assumptions (1.11)–(1.13), assume that there exists a positive number \(\sigma _0\) such that

$$\begin{aligned}&dist(\Gamma (0), \Gamma _b)>0, \end{aligned}$$
(1.23)
$$\begin{aligned}&2\sigma _0\le J_0\le 3\sigma _0. \end{aligned}$$
(1.24)

If the initial data \((v_0,\, \eta _0)\in (X^{12}_{\frac{1}{2}}\cap H^1(\Omega )) \times \mathcal {F}_{\kappa }(\Omega )\) for some constant \(\kappa \in (0, \frac{1}{16})\), then the system (1.18) is locally well-posed. More precisely, there exists a positive time \(T>0\) such that the system (1.18) has a unique solution \((v,\,\eta )\in \,C([0, T]; X^{12}_{\frac{1}{2}}\cap H^1(\Omega )) \times C([0, T]; \mathcal {F}_{\kappa }(\Omega ))\) depending continuously on initial data \((v_0,\, \eta _0)\in (X^{12}_{\frac{1}{2}}\cap H^1(\Omega )) \times \mathcal {F}_{\kappa }(\Omega )\), and there hold

$$\begin{aligned}&\sup _{t\in [0,T]}(\Vert v\Vert ^2_{X^{12}_{\frac{1}{2}}}+\Vert v\Vert _{H^1}^2) +\int _0^T \left( \Vert \nabla v\Vert _{X^{12}}^2+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}^2\right) \, ds\le C,\nonumber \\&\sup _{t\in [0,T]}\Vert \xi (t)\Vert ^2_{\mathcal {F}_{\kappa }}\le C,\quad \,\sigma _0\le \sup _{(t, x)\in [0,T]\times \Omega }J(t, x)\le 4\sigma _0, \end{aligned}$$
(1.25)

where C depends on initial data.

Remark 1.3

The assumption (1.11)–(1.13) on \(\rho _0\) is reasonable. In fact, if \(\Omega (0)=\Omega :=\{x \in \mathbb {T}^2\times \mathbb {R}|\quad 0<x_3<1\}\) and \(\rho _0=dist(x,\partial \Omega )\sim x_3(1-x_3)\), then the assumptions (1.11)– (1.13) are automatically satisfied .

Remark 1.4

In this paper, we consider the case that \(\gamma =2\). But our method may still work for all the cases \(\gamma >1\).

Remark 1.5

For any \(t \in [0, T]\), since \(\sigma _0\le \sup _{(t, x)\in [0,T]\times \Omega }J(t, x)\le 4\sigma _0\), the flow-map \(\eta (t, x)\) defines a diffeomorphism from the equilibrium domain \(\Omega \) to the moving domain \(\Omega (t)\) with the boundary \(\Gamma (t)\). From this, together with the fact that \(\eta _0\) is a diffeomorphism from the equilibrium domain \(\Omega \) to the initial domain \(\Omega (0)\), we deduce a diffeomorphism from the initial domain \(\Omega (0)\) to the evolving domain \(\Omega (t)\) for any \(t \in [0, T]\). Denote the inverse of the flow map \(\eta (t, x)\) by \(\eta ^{-1}(t, y)\) for \(t \in [0, T]\) so that if \(y = \eta (t, x)\) for \(y \in \Omega (t)\) and \(t \in [0, T]\), then \(x = \eta ^{-1}(t, y) \in \Omega \).

For the strong solution \((\eta , v)\) obtained in Theorem 1.2, and for \(y \in \Omega (t)\) and \(t \in [0, T]\), we denote that

$$\begin{aligned} \rho (t,\,y):=J^{-1}(t, \eta ^{-1}(t, y))\overline{\rho }_0(\eta ^{-1}(t, y)),\quad \,u(t,\,y):=v(t,\,\eta ^{-1}(t, y)). \end{aligned}$$
(1.26)

Then the triple \((\rho (t,\,y), u(t, y), \Omega (t))\) (\(t \in [0, T]\)) defines a strong solution to the free boundary problem (1.1). Furthermore, we obtain the following theorem.

Theorem 1.6

Under the assumptions in Theorem 1.2, the free boundary problem (1.1) is locally well-posed, and the triple \((\rho (t,\,y), u(t, y), \Omega (t))\) (\(t \in [0, T]\)) defined in Remark 1.5 and (1.26) is the unique strong solution to the free boundary problem (1.1) satisfying \(\eta -Id \in \,C([0, T],\, \mathcal {F}_{\kappa })\).

The rest of the paper is organized as follows. In Sect. 2, we derive some preliminary estimates. Some necessary a priori estimates are obtained in Sect. 3. Finally in Sect. 4, the proof of Theorem 1.2 is proved.

Let us complete this section with some notations that we use in this context.

Notations Let AB be two operators, we denote \([A, B]=AB-BA,\) the commutator between A and B. For \(a\lesssim b\), we mean that there is a uniform constant C,  which may be different on different lines, such that \(a\le Cb\) and \(C_0\) denotes a positive constant depending on the initial data only.

2 Preliminary estimates

In what follows, we denote by C a positive constant which may depend on initial data \((v_0,\eta _0)\) if we don’t make a special explanation in it. This notation is allowed to change from one inequality to the next.

We first introduce the following inequality which we heavily use throughout the paper.

Lemma 2.1

(Hardy inequality, [23]) For any \(\varepsilon >0,\) there holds that

$$\begin{aligned} \Vert \overline{\rho }^{-\frac{1}{2}+\varepsilon }f\Vert _{L^2(\Omega )}\le C(\Vert \overline{\rho }^{\frac{1}{2}+\varepsilon }f\Vert _{L^2(\Omega )}+\Vert \overline{\rho }^{\frac{1}{2}+\varepsilon }\nabla f\Vert _{L^2(\Omega )}). \end{aligned}$$

With Hardy inequality in hand, we may get the following interpolation equalities.

Lemma 2.2

For any \(\kappa \in (0,\frac{1}{16})\), there hold that, for \(0\le \ell \le 6\),

$$\begin{aligned} \Vert Z^\ell \nabla f\Vert _{L^\infty _{x_3}(L^2_h)}\le C(\Vert \nabla f\Vert _{X^{12}}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}). \end{aligned}$$
(2.1)

and for \(0\le \ell \le 4\),

$$\begin{aligned} \Vert Z^\ell \nabla f\Vert _{L^\infty }\le C(\Vert \nabla f\Vert _{X^{12}}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}). \end{aligned}$$
(2.2)

Proof

For \(0\le \ell \le 6\), thanks to the Sobolev embedding theorem and Lemma 2.1, we have

$$\begin{aligned} \Vert Z^\ell \nabla f\Vert _{L^\infty _{x_3}(L^2_h)}&\le C_0 (\Vert \overline{\rho }^{-\frac{21}{44}}Z^\ell \nabla f\Vert _{L^2_{x_3}(L^2_h)}+ \Vert \overline{\rho }^{\frac{21}{44}}\partial _3 Z^\ell \nabla f\Vert _{L^2_{x_3}(L^2_h)}) \nonumber \\&\le C_0 (\Vert \overline{\rho }^{\frac{23}{44}}Z^\ell \nabla f\Vert _{L^2}+\Vert \overline{\rho }^{\frac{23}{44}}\nabla Z^\ell \nabla f\Vert _{L^2}\nonumber \\&\quad + \sum _{i=0}^\ell (\Vert \overline{\rho }^{\frac{21}{44}}Z^{i+1} \nabla f\Vert _{L^2}+ \Vert \overline{\rho }^{\frac{21}{44}}Z^i\triangle f\Vert _{L^2})\nonumber \\&\le C\Vert \nabla f\Vert _{X^{12}}+C \sum _{i=0}^\ell \Vert \overline{\rho }^{\frac{21}{44}}Z^i\triangle f\Vert _{L^2}. \end{aligned}$$
(2.3)

According to the fact \(|Z \bar{\rho }|\le C\bar{\rho }\), we deduce from integration by parts that

$$\begin{aligned} \Vert \overline{\rho }^{\frac{21}{44}}Z^i\triangle f\Vert _{L^2}&\le C\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}^{1-\frac{7}{11}} \Vert \triangle f\Vert _{\dot{X}^{11}_1}^{\frac{7}{11}}\nonumber \\&\le C(\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}+ \Vert \triangle f\Vert _{\dot{X}^{11}_1}),\quad \forall \,\,\, i\le 5, \end{aligned}$$
(2.4)

where we used that \(\frac{14}{22}+(-\frac{1}{2}+\kappa )\frac{4}{11}\le \frac{21}{44}<\frac{1}{2}\) with \(\kappa \in (0, \frac{1}{16})\).

While by using integration by parts again, one can see that

$$\begin{aligned}&\Vert \overline{\rho }^{\frac{21}{44}}Z^6\triangle f\Vert ^2_{L^2}=\int _{\Omega }\bar{\rho }^{\frac{21}{22}} Z^6\triangle f\cdot Z^6\triangle f\, dx\\&\quad =-\int _{\Omega }\bar{\rho }^{\frac{21}{22}} Z\triangle f\cdot Z^{11}\triangle f\, dx-\int _{\Omega } [\rho ^{\frac{21}{22}}; \,Z^5]\triangle f\cdot Z^6\triangle f\, dx, \end{aligned}$$

which follows from the fact \(|Z \bar{\rho }|\le C\bar{\rho }\) that

$$\begin{aligned} \Vert \overline{\rho }^{\frac{21}{44}}Z^6\triangle f\Vert ^2_{L^2} \le C\Vert \triangle f\Vert _{X^{11}_1}(\Vert \bar{\rho }^{-\frac{1}{22}}\triangle f\Vert _{L^2}+\Vert \bar{\rho }^{-\frac{1}{22}}Z\triangle f\Vert _{L^2}). \end{aligned}$$

Next, we deal with the last term in the above inequality. In fact, we may get from integration by parts that

$$\begin{aligned} \Vert \bar{\rho }^{-\frac{1}{22}}Z\triangle f\Vert _{L^2}^2&=\int _{\Omega }\bar{\rho }^{-\frac{1}{11}}Z\triangle f\cdot Z\triangle f\,dx \le C\Vert \rho ^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2} \,\sum _{k=0}^2\Vert \rho ^{\frac{1}{2}-\frac{1}{11}-\kappa } Z^k\triangle f \Vert _{L^2}\\&\le C\Vert \rho ^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2} (\Vert \triangle f\Vert _{X^{11}_1}+C_0\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}), \end{aligned}$$

which implies

$$\begin{aligned} \Vert \bar{\rho }^{-\frac{1}{22}}Z\triangle f\Vert _{L^2}\le C(\Vert \triangle f\Vert _{X^{11}_1}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}). \end{aligned}$$

Hence, one has

$$\begin{aligned} \Vert \overline{\rho }^{\frac{21}{44}}Z^6\triangle f\Vert ^2_{L^2} \le C\Vert \triangle f\Vert _{X^{11}_1}(\Vert \bar{\rho }^{-\frac{1}{22}}\triangle f\Vert _{L^2}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}+\Vert \triangle f\Vert _{X^{11}_1}). \end{aligned}$$
(2.5)

Inserting (2.42.5) into (2.3) ensures that for \(0\le \ell \le 6\)

$$\begin{aligned} \Vert Z^\ell \nabla f\Vert _{L^\infty _{x_3}(L^2_h)} \le C(\Vert \nabla f\Vert _{X^{12}}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}), \end{aligned}$$

that is, the inequality (2.1) holds.

The second inequality (2.2) comes from the Sobolev embedding theorem and (2.1):

$$\begin{aligned} \Vert Z^\ell \nabla f\Vert _{L^\infty }\le C\,\Vert Z^\ell \nabla f\Vert _{L^\infty _{x_3}H^2_h}\le C(\Vert \nabla f\Vert _{X^{12}}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle f\Vert _{L^2}) \end{aligned}$$

for \(0\le \ell \le 4\), which ends the proof of Lemma 2.2. \(\square \)

To deal with nonlinear term, we need the following product laws in the spaces \(X^{N}\).

Lemma 2.3

There hold true that

$$\begin{aligned} \Vert g\,f\Vert _{X^{12}}\le C\Vert g\Vert _{X^{12}} \sum _{|\ell |\le 6}\Vert Z^\ell f\Vert _{L^\infty _{x_3}(L^2_h)}+C \Vert f\Vert _{X^{12}}\sum _{|\ell |\le 6}\Vert Z^\ell g\Vert _{L^\infty _{x_3}(L^2_h)}, \end{aligned}$$
(2.6)

and

$$\begin{aligned} \sum _{0\le |j|\le 1}\Vert g~Z^{j}f\Vert _{X^{11}}\le C\Vert g\Vert _{X^{11}} \sum _{|\ell |\le 6} \Vert Z^\ell f\Vert _{L^\infty _{x_3}(L^2_h)}+C\Vert f\Vert _{X^{12}}\sum _{|\ell |\le 6} \Vert Z^\ell g\Vert _{L^\infty _{x_3}(L^2_h)}. \end{aligned}$$
(2.7)

Proof

By the Leibnitz formula, one can see that

$$\begin{aligned} \Vert g\,f\Vert _{X^{12}}\le C\,\sum _{|m_1|+|m_2|=0}^{12}\Vert Z^{m_1}g~Z^{m_2}f\Vert _{L^2}. \end{aligned}$$

Now, we focus only on the proof of the most difficulty case: \(|m_1|+|m_2|=12\). The others can be treated by a similar way. In fact, we divide its proof into three cases.

  • Case 1. \(8\le |m_1|\le 12\). By Hölder’s inequality, we prove

    $$\begin{aligned} \Vert Z^{m_1}g~ Z^{m_2}f\Vert _{L^2}\le & {} \Vert Z^{m_1}g \Vert _{L^2}\Vert Z^{m_2}f\Vert _{L^\infty } \le C\Vert Z^{m_1}g \Vert _{L^2}\Vert Z^{m_2}f \Vert _{L^\infty _{x_3}(H^2_h)}\\\le & {} C\Vert g\Vert _{X^{12}} \sum _{|\ell |\le 6}\Vert Z^\ell f\Vert _{L^\infty _{x_3}(L^2_h)}, \end{aligned}$$

    where we used \(|m_2|+2\le 6\).

  • Case 2. \(6\le |m_1|\le 7\). Thanks to the Sobolev embedding theorem and Hölder’s inequality, one can obtain that

    $$\begin{aligned} \Vert Z^{m_1}g~ Z^{m_2}f\Vert _{L^2}\le & {} \Vert Z^{m_1}g \Vert _{L^2_{x_3}(L^\infty _{h})}\Vert Z^{m_2}f\Vert _{L^\infty _{x_3}(L^2_h)}\\\le & {} C\Vert g\Vert _{X^{12}} \sum _{|\ell |\le 6}\Vert Z^\ell f\Vert _{L^\infty _{x_3}(L^2_h)}. \end{aligned}$$
  • Case 3. \( 0\le |m_1|\le 5\). For this case, we only need to exchange the position of f and g and apply the same argument as in the above two cases to get that

    $$\begin{aligned} \Vert Z^{m_1}g~ Z^{m_2}f\Vert _{L^2}\le & {} C\Vert f\Vert _{X^{12}} \sum _{|\ell |\le 6}\Vert Z^\ell g\Vert _{L^\infty _{x_3}(L^2_h)}. \end{aligned}$$

Collecting all the above cases together, we obtain

$$\begin{aligned} \Vert g~f\Vert _{X^{12}}\le C\Vert g\Vert _{X^{12}} \sum _{|\ell |\le 6}\Vert Z^\ell f\Vert _{L^\infty _{x_3}(L^2_h)}+C\sum _{|\ell |\le 6}\Vert Z^\ell g\Vert _{L^\infty _{x_3}(L^2_h)}\Vert f\Vert _{X^{12}}, \end{aligned}$$

which follows (2.6).

Next, since we the highest order in (2.7) is 11, we may readily verify (2.7) by the same process above, which ends the proof of Lemma 2.3. \(\square \)

We introduce a new quantity \(\mathfrak { D}(v)(t)\) which controls \(\Vert \nabla v\Vert _{L^\infty }\) according to Lemma 2.2:

(2.8)

In what follows, \(\mathcal {P}(\cdot )\) stands for some polynomial function which coefficients may depend on initial data.

Lemma 2.4

Assume that

$$\begin{aligned} \xi _0\in \mathcal {F}_{\kappa }, \quad \Vert \mathfrak { D}(v)\Vert _{L^2(0, T)}\le \mathfrak {C},\quad \sigma _0\le J\le 4\sigma _0. \end{aligned}$$

Then there hold that for any \(t \in [0, T]\)

$$\begin{aligned} \Vert \nabla v:\nabla v(t)\Vert _{X^{12}}\le C\mathfrak { D}(v)^2(t), \end{aligned}$$
(2.9)

and

$$\begin{aligned}&\sum _{0\le |\ell |\le 6}\Vert Z^\ell (J\mathcal {A})(t)\Vert _{L^\infty _{x_3}(L^2_h)}\le C (1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})),\nonumber \\&\sum _{0\le |\ell |\le 6}\Vert Z^\ell \mathcal {A}(t)\Vert _{L^\infty _{x_3}(L^2_h)} \le C (1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})),\nonumber \\&\Vert J\mathcal {A}(t)\Vert _{X^{12}}\le C (1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})),\quad \Vert \mathcal {A}(t) \Vert _{X^{12}}\le C (1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})), \end{aligned}$$
(2.10)

where the constant C depends on \(\Vert \xi _0\Vert _{\mathcal {F}_{\kappa }}\) and \(\sigma _0\).

Proof

Before giving the proof of this lemma, we state some estimates as preliminary.

First, taking \(f=g=\nabla v\) in (2.6), we obtain

$$\begin{aligned} \Vert \nabla v:\nabla v\Vert _{X^{12}}\le C\Vert \nabla v\Vert _{X^{12}} \sum _{|\ell |\le 6} \Vert Z^\ell \nabla v\Vert _{L^\infty _{x_3}(L^2_h)}. \end{aligned}$$
(2.11)

While by Lemma 2.2, one can prove that

$$\begin{aligned}&\sum _{0\le |\ell |\le 6} \Vert Z^\ell (\nabla v:\nabla v)\Vert _{L^\infty _{x_3}(L^2_h)}\le C \sum _{0\le |\ell |\le 6}\Vert Z^\ell \nabla v\Vert _{L^\infty _{x_3}(L^2_h)} \sum _{0\le |\ell |\le 4} \Vert Z^\ell \nabla v\Vert _{L^\infty }\nonumber \\&\quad \le \,C (\Vert \nabla v\Vert _{X^{12}}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle v\Vert _{L^2})^2\le \,C\mathfrak { D}(v)^2, \end{aligned}$$
(2.12)

which along with (2.11) ensures (2.9).

Now we are in the position to prove the estimates in terms of \(J\mathcal {A}\) and \(\mathcal {A}\). Notice that

$$\begin{aligned} J\mathcal {A}=(D\eta )^{-1}=\left( \nabla \eta _0+\int _{0}^{t}\nabla v\,ds\right) ^{-1}, \end{aligned}$$

and every entry in \(J\mathcal {A}\) is a linear combination of

$$\begin{aligned} \nabla \eta _0,\, \nabla \eta _0\int _{0}^{t}\nabla vds,\,\left( \int _{0}^{t}\nabla vds\right) ^2. \end{aligned}$$

Then, thanks to Lemmas 2.22.3, (2.12) and Minkowski’s inequality, one has

$$\begin{aligned}&\sum _{0\le |\ell |\le 6} \Vert Z^\ell (J\mathcal {A})\Vert _{L^\infty _{x_3}(L^2_h)}\nonumber \\&\quad \le \sum _{0\le |\ell |\le 6} \Vert Z^\ell \nabla \eta _0\Vert _{L^\infty _{x_3}(L^2_h)}+\sum _{0\le |\ell |\le 6} \Vert Z^\ell \left( \nabla \eta _0\int _{0}^{t}\nabla vds\right) \Vert _{L^\infty _{x_3}(L^2_h)}\nonumber \\&\qquad +\sum _{0\le |\ell |\le 6} \Vert Z^\ell \left( \left( \int _{0}^{t}\nabla vds\right) ^2\right) \Vert _{L^\infty _{x_3}(L^2_h)}\nonumber \\&\quad \le C \,\Vert \xi _0\Vert _{\mathcal {F}_{\kappa }}+C \,\Vert \xi _0\Vert _{\mathcal {F}_{\kappa }}\,t^{\frac{1}{2}}\Vert \mathfrak { D}(v)\Vert _{L^2_t}+C t\Vert \mathfrak { D}(v)\Vert _{L^2_t}^2 \le C \,(1+t^\frac{1}{2}\,\mathcal {P}(\mathfrak {C})), \end{aligned}$$
(2.13)

which proves the first inequality in (2.10).

Similarly, we deduce that

$$\begin{aligned}&\sum _{0\le |\ell |\le 6} \Vert Z^\ell (\nabla \eta _0\int _{0}^{t}\nabla vds)\Vert _{L^\infty _{x_3}(L^2_h)}+\sum _{0\le |\ell |\le 6} \Vert Z^\ell \left( \left( \int _{0}^{t}\nabla v\,ds\right) ^2\right) \Vert _{L^\infty _{x_3}(L^2_h)} \nonumber \\&\quad \le C\,t^{\frac{1}{2}}\Vert \mathfrak { D}(v)\Vert _{L^2_t}+C\,t\Vert \mathfrak { D}(v)\Vert _{L^2_t}^2\le C\,t^{\frac{1}{2}} \mathcal {P}(\mathfrak {C}). \end{aligned}$$
(2.14)

Recalling the definition of J: \(J=\det (\nabla \eta _0+\int _0^t\nabla vds)\), J is a linear combination of the terms

$$\begin{aligned} (\nabla \eta _0)^3,\,\nabla \eta _0\left( \int _{0}^{t}\nabla vds\right) ^2,\, (\nabla \eta _0)^2\int _{0}^{t}\nabla vds,\,\left( \int _{0}^{t}\nabla vds\right) ^3. \end{aligned}$$

Hence, similar to the proof of the first inequality in (2.10) in terms of \(J\mathcal {A}\), we may obtain

$$\begin{aligned} \sum _{0\le |\ell |\le 6} \Vert Z^\ell J\Vert _{L^\infty _{x_3}(L^2_h)}\le C\,(1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})). \end{aligned}$$
(2.15)

Owing to the fact \(J\ge \sigma _0\) and the formula to the composition of two functions, we obtain

$$\begin{aligned} \sum _{0\le |\ell |\le 6} \Vert Z^\ell (J^{-1})\Vert _{L^\infty _{x_3}(L^2_h)}\le&C\,\sum _{0\le |\ell |\le 6} \Vert \prod _{\sum _j |k_j| m_j\le |\ell |}(Z^{k_j}J)^{m_j}\Vert _{L^\infty _{x_3}(L^2_h)}. \end{aligned}$$

We put \(\Vert \cdot \Vert _{L^\infty _{x_3}(L^2_h)}\) on the highest order term \(Z^{k_j}J\) and put \(\Vert \cdot \Vert _{L^\infty }\) to other lower terms (not more than order 4) with similar process to (2.12). It follows from Lemma 2.2 and (2.15) that

$$\begin{aligned} \sum _{0\le |\ell |\le 6}\delta ^{|\ell |}\Vert Z^\ell (J^{-1})\Vert _{L^\infty _{x_3}(L^2_h)}\le C\,(1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})). \end{aligned}$$
(2.16)

Therefore, due to (2.13) and (2.16), we find

$$\begin{aligned} \sum _{|\ell |\le 6} \Vert Z^{\ell }\mathcal {A}\Vert _{L^\infty _{x_3}(L^2_h)}\le&C+C\sum _{|\ell |\le 6} \Vert Z^{\ell }(J\mathcal {A})\Vert _{L^\infty _{x_3}(L^2_h)}\sum _{|\ell |\le 4} \Vert Z^{\ell }(J^{-1})\Vert _{L^\infty }\\&+C\sum _{|\ell |\le 6}\Vert Z^{\ell }(J^{-1})\Vert _{L^\infty _{x_3}(L^2_h)} \sum _{|\ell |\le 4}\Vert Z^{\ell }(J\mathcal {A})\Vert _{L^\infty } \\ \le&C\,(1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})). \end{aligned}$$

For the high order estimate, similar to the proof of (2.9), by using Lemma 2.2, we achieve

$$\begin{aligned} \Vert \nabla \eta _0\left( \int _{0}^{t}\nabla vds\right) ^2\Vert _{X^{12}}+\Vert (\nabla \eta _0)^2\int _{0}^{t}\nabla vds\Vert _{X^{12}}+\Vert \left( \int _{0}^{t}\nabla vds\right) ^3\Vert _{X^{12}}\le C\,t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}), \end{aligned}$$
(2.17)

and then

$$\begin{aligned} \Vert (J,\,J\mathcal {A})\Vert _{X^{12}}&\le C\,\Big (\Vert \nabla \eta _0\Vert _{X^{12}}+\Vert \nabla \eta _0\Big (\int _{0}^{t}\nabla vds\Big )^2\Vert _{X^{12}}\nonumber \\&\quad +\Vert (\nabla \eta _0)^2\int _{0}^{t}\nabla vds\Vert _{X^{12}}+\Vert \Big (\int _{0}^{t}\nabla vds\Big )^3\Vert _{X^{12}}\Big )\nonumber \\&\le C+C\,t^{\frac{1}{2}}\Vert \mathfrak { D}(v)\Vert _{L^2_t}+C\,t\Vert \mathfrak { D}(v)\Vert _{L^2_t}^2+C\,t^\frac{3}{2}\Vert \mathfrak { D}(v)\Vert _{L^2_t}^3\nonumber \\&\le C\,(1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})). \end{aligned}$$
(2.18)

While by virtue of (2.15), (2.18) and Lemma 2.3, we deduce that

$$\begin{aligned} \Vert J^{-1}\Vert _{X^{12}}\le C\,(1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})), \end{aligned}$$

and

$$\begin{aligned} \Vert \mathcal {A}\Vert _{X^{12}}=\Vert J\mathcal {A}~J^{-1}\Vert _{X^{12}} \le C\,(1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})), \end{aligned}$$

which completes the proof of Lemma 2.4. \(\square \)

Based on the above lemma, we may get the following estimates.

Lemma 2.5

Under the assumptions in Lemma 2.4, there hold

$$\begin{aligned} \sum _{0\le |j|\le 1}\Vert Z^j(J\mathcal {A})~\nabla v \Vert _{X^{11}}&\le C\,\Vert \nabla v\Vert _{X^{11}}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v),\nonumber \\ \sum _{0\le |j|\le 1}\Vert Z^j(\mathcal {A})~\nabla v \Vert _{X^{11}}&\le C\,\Vert \nabla v\Vert _{X^{11}}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v),\nonumber \\ \Vert \nabla _{\mathcal {A}} v\Vert _{X^{12}}&\le C\,\Vert \nabla v\Vert _{X^{12}}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v),\nonumber \\ \Vert \mathrm {S}_{J\mathcal {A}} (v)\Vert _{X^{12}}&\le C\,\Vert \nabla v\Vert _{X^{12}}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v). \end{aligned}$$
(2.19)

Proof

We mainly utilize Lemmas 2.3, 2.4 to prove (2.19). So one may focus only on the proof of the first inequality in (2.19), and the proofs of the others are the same as it, whose details will be omitted here.

First, by the definition of \(J\mathcal {A}\), we split \(\sum _{0\le |j|\le 1}\Vert \nabla v ~Z^j(J\mathcal {A})\Vert _{X^{11}}\) into three parts:

$$\begin{aligned}&\sum _{0\le |j|\le 1}\Vert Z^j(J\mathcal {A})~\nabla v \Vert _{X^{11}}\nonumber \\&\quad \le C\sum _{0\le |j|\le 1} \Vert \nabla v\, Z^j\nabla \eta _0 \Vert _{X^{11}} +C\sum _{0\le |j|\le 1} \Vert \nabla v\, Z^j\left( \nabla \eta _0\int _{0}^{t}\nabla vds\right) \Vert _{X^{11}}\nonumber \\&\qquad + C\sum _{0\le |j|\le 1} \Vert \nabla v\, Z^j\left( \int _{0}^{t}\nabla vds \right) ^2\Vert _{X^{11}} \triangleq \sum _{i=1}^3 I_i. \end{aligned}$$
(2.20)

For \(I_1\), we have

$$\begin{aligned} I_1\le C\,\Vert \nabla v\Vert _{X^{11}}. \end{aligned}$$
(2.21)

For \(I_2\), taking \(g=\nabla v\) and \(f=J\mathcal {A}\) in (2.7) in Lemma 2.3 to obtain that

$$\begin{aligned} I_2&\le C\Vert \nabla v \Vert _{X^{11}}\sum _{|\ell |\le 6}\Vert Z^j\nabla \eta _0\int _{0}^{t}\nabla vds \Vert _{L^\infty _{x_3}(L^2_h)}\\&\quad +C \sum _{|\ell |\le 6}\Vert Z^\ell \nabla v\Vert _{L^\infty _{x_3}(L^2_h)}\Vert Z^j\nabla \eta _0\int _{0}^{t}\nabla vds \Vert _{X^{12}}. \end{aligned}$$

Applying Lemma 2.2 and (2.14), (2.17) in Lemma 2.4 to get

$$\begin{aligned} I_2\le t^{\frac{1}{2}} \mathcal {P}(\mathfrak {C})\Vert \nabla v \Vert _{X^{11}}+ Ct^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v) \le \,t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v). \end{aligned}$$
(2.22)

Similarly, we have

$$\begin{aligned} I_3 \le t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v). \end{aligned}$$
(2.23)

Plugging the estimates (2.21)–(2.23) into (2.20), we prove

$$\begin{aligned} \sum _{0\le |j|\le 1}\Vert Z^j(J\mathcal {A})~\nabla v \Vert _{X^{11}}\le&C\,\Vert \nabla v\Vert _{X^{11}}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v), \end{aligned}$$

which ends our proof. \(\square \)

Next we recall a version of Korn’s inequality involving only the deviatoric part \(\mathbb {D}^0\).

Lemma 2.6

(Korn’s lemma, Theorem 1.1 in [5]) Let \(n \ge 3\) and U be a Lipschitz domain in \(\mathbb {R}^n\), then there exists a constant C, independent of f, such that

$$\begin{aligned} \Vert f\Vert _{H^1(U)} \le C\,(\Vert \mathbb {D}^0(f)\Vert _{L^2(U)}+\Vert f\Vert _{L^2(U)}) \end{aligned}$$

for all \(f \in H^1(U)\).

3 A priori estimates

In this section, we give a priori estimates of the system (1.18). The main result of the section is as follows:

Proposition 3.1

Assume \((\xi ,\,v)\) is a smooth solution of system (1.18) on \([0,\bar{T}]\) with initial data \((\xi _0,\,v_0) \in \mathcal {F}_{\kappa } \times (X^{12}_{\frac{1}{2}}\cap \,H^1)\) and \(0<2\sigma _0\le J_0\le 3\sigma _0\), and \(\overline{\rho }\) satisfies (1.11)–(1.13). Then, there exists a positive constant \(T\le \bar{T}\) which depends on the initial data such that

$$\begin{aligned} \sup _{t\in [0,T]}\mathcal {E}(t)+\int _0^T\mathcal {D}(s)ds\le 2\mathcal {E}(0). \end{aligned}$$

Here, we use the bootstrap argument to prove this proposition. Now, we define a T such that there holds that

$$\begin{aligned} \Vert \mathfrak { D}(v)\Vert _{L^2(0, T)}\le \mathfrak {C},\quad \sigma _0\le \sup _{t\in [0, T]} J\le 4\sigma _0. \end{aligned}$$
(3.1)

Before, we give the proof of the proposition, we prove some useful lemmas.

Lemma 3.2

Under the assumption of Proposition 3.1, we have

$$\begin{aligned} \Vert \nabla v\Vert _{L^1(0, t; L^\infty )} \le t^\frac{1}{2}\mathcal {P}(\mathfrak {C}),\quad \Vert (J, A)(t)\Vert _{L^\infty } \le C\,(1+t^{\frac{1}{2}} \mathcal {P}(\mathfrak {C})) ,\quad \forall \,\, t\in [0, T). \end{aligned}$$

Proof

It is a direct result from Lemma 2.2 and Lemma 2.4. \(\square \)

Lemma 3.3

Under the assumption of Proposition 3.1, the following holds

$$\begin{aligned} \Vert \nabla v\Vert _{X^{N}}\le C\,(\Vert \mathbb {D}^0(v)\Vert _{X^{N}}+\Vert v\Vert _{X^N_{\frac{1}{2}}}). \end{aligned}$$
(3.2)

Proof

Thanks to Korn’s lemma (Lemma 2.6), we have

$$\begin{aligned} \Vert v\Vert _{H^1}\le C_0(\Vert \mathbb {D}^0(v)\Vert _{L^2}+\Vert v\Vert _{L^2}). \end{aligned}$$

For any function f(s), by Lemma 2.1, we have

$$\begin{aligned} \int _0^1f^2ds\le C_0\int _0^1s^2(f^2+f'^2)ds \end{aligned}$$

By scaling, we have

$$\begin{aligned} \int _{1-\varepsilon }^1 f^2ds\le \frac{C_0}{\varepsilon ^2}\int _{1-\varepsilon }^1 (1-s)^2f^2ds+C_0\int _{1-\varepsilon }^1 ({1-s})^2f'^2ds. \end{aligned}$$

Then (1.12) gives that

$$\begin{aligned} \Vert f\Vert _{L^2}^2\le C\,\Vert {\overline{\rho }}^{-1}\Vert _{L^\infty (0\le x_3\le 1-\varepsilon )}\int _{\Omega }\overline{\rho }f^2dx+C\,\varepsilon ^2\Vert f\Vert _{H^1}^2\le \frac{C}{\varepsilon }\int _{\Omega }\overline{\rho }f^2dx+C\,\varepsilon ^2\Vert f\Vert _{H^1}^2. \end{aligned}$$
(3.3)

Taking \(\varepsilon \) small enough and \(f:=v\), we combine with Lemma 2.6 to get that

$$\begin{aligned} \Vert v\Vert _{H^1}\le C\,(\Vert \mathbb {D}^0(v)\Vert _{L^2}+\Vert \overline{\rho }^\frac{1}{2}v\Vert _{L^2}). \end{aligned}$$

For given \(m\in \mathbb {N}^3\): \(1\le |m| \le N\),

$$\begin{aligned} \Vert Z^mv\Vert _{H^1}\le&C\,(\Vert \mathbb {D}^0(Z^mv)\Vert _{L^2}+\Vert \overline{\rho }^\frac{1}{2}Z^mv\Vert _{L^2})\\ \le&C\,(\Vert Z^m\mathbb {D}^0v\Vert _{L^2} +\Vert [\mathbb {D}^0,Z^m]v\Vert _{L^2}+\Vert \overline{\rho }^\frac{1}{2} Z^mv\Vert _{L^2}), \end{aligned}$$

which follows from the fact \([\mathbb {D}^0,Z^m]v\sim Z^{m-1}\, \nabla \,v\) that

$$\begin{aligned} \Vert Z^mv\Vert _{H^1}\le C\,(\Vert Z^m\mathbb {D}^0v\Vert _{L^2} +\Vert Z^{m-1}\nabla \,v\Vert _{L^2}+\Vert \overline{\rho }^\frac{1}{2} Z^mv\Vert _{L^2}). \end{aligned}$$

Therefore, by a standard inductive argument in terms of \(m=0,1,\ldots ,N\) and the definition of space \(X^{N}\), we prove (3.2). \(\square \)

Lemma 3.4

Let the initial flow map \(\eta _0=Id+\xi _0: \Omega \rightarrow \Omega (0)\) satisfy its Jacobian \(2\sigma _0\le \,J_0\le 3\sigma _0\) and \(\xi _0 \in \mathcal {F}_{\kappa }\), and its inverse map \(\eta _0^{-1}: \Omega (0)\rightarrow \Omega \), \(v(x)=\widetilde{u}(\eta _0(x))\) with \(x \in \Omega \) and \(\widetilde{u}(y)=v(\eta _0^{-1}(y))\) with \(y \in \Omega (0)\), then there is a positive constant \(C_1\ge 1\) such that

$$\begin{aligned} C_1^{-1} (1+\Vert \xi _0\Vert _{\mathcal {F}_{\kappa }}^2)^{-1}\int _{\Omega }|\nabla \,v|^2\,dx \le \int _{\Omega (0)}|\nabla _y\,\widetilde{u}(y)|^2\,dy\le \,C_1 (1+\Vert \xi _0\Vert _{\mathcal {F}_{\kappa }}^2)\int _{\Omega }|\nabla \,v|^2\,dx. \end{aligned}$$
(3.4)

Proof

First, taking changes of variables \(y=\eta _0(x)\), we have

$$\begin{aligned} \int _{\Omega (0)}|\nabla _y\,\widetilde{u}(y)|^2\,dy=\int _{\Omega }|\nabla _y\,v(x)|^2\,d(\eta _0(x)) ={\int _{\Omega }}|(D_y(\eta _0^{-1}))(\eta _0(x))\nabla _x\,v(x)|^2\,J_0dx, \end{aligned}$$

which along with the assumptions \(2\sigma _0\le \,J_0\le 3\sigma _0\), \(\xi _0 \in \mathcal {F}_{\kappa }\), and (2.2) implies

$$\begin{aligned} \int _{\Omega (0)}|\nabla _y\,\widetilde{u}(y)|^2\,dy&\le C \Vert (D_y(\eta _0^{-1}))(\eta _0(x))\Vert _{L^\infty }^2\int _{\Omega }|\nabla _x\,v(x)|^2\,dx\\&\le C_1(1+\Vert \xi _0\Vert _{\mathcal {F}_{\kappa }}^2)\int _{\Omega }|\nabla \,v|^2\,dx. \end{aligned}$$

Similarly, one may readily check

$$\begin{aligned} \int _{\Omega }|\nabla _x\,v(x)|^2\,dx&=\int _{\Omega (0)}|(D_x\eta _0)(\eta _0^{-1}(y))\nabla _y\,\widetilde{u}(y)|^2\,J_0^{-1}dy\\&\le \,C_1(1+\Vert \xi _0\Vert _{\mathcal {F}_{\kappa }}^2) \int _{\Omega (0)}|\nabla _y\,\widetilde{u}(y)|^2\,dy. \end{aligned}$$

Therefore, we get (3.4), and complete the proof of Lemma 3.4. \(\square \)

Lemma 3.5

Under the assumption of Proposition 3.1, if (3.1) holds, then we have

$$\begin{aligned} (c_0- t^2\mathcal {P}(\mathfrak {C}))\Vert \nabla v\Vert _{L^2}^2-C_0\Vert \overline{\rho }^\frac{1}{2} v\Vert _{L^2}^2\le \Vert \mathbb {D}_{\mathcal {A}}^0 v\Vert _{L^2}^2\le C_0(1+t^2\mathcal {P}(\mathfrak {C}))\Vert \nabla v\Vert _{L^2}^2. \end{aligned}$$

Moreover, if T is small enough such that \(T^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})<\frac{c_0}{2}\), then we have

$$\begin{aligned} \int _\Omega J\mathbb {S}_\mathcal {A}v: \nabla _{\mathcal {A}}v \,dx\ge c_1\Vert \mathbb {D}_\mathcal {A}^0 v\Vert _{L^2}^2 \ge \frac{c_0c_1}{2}\Vert \nabla v\Vert _{L^2}^2-C_0\Vert \overline{\rho }^\frac{1}{2} v\Vert _{L^2}^2. \end{aligned}$$

Proof

We first to prove the first result. According to the fact

$$\begin{aligned} J\mathcal {A}-J_0\mathcal {A}_0\sim \left( \int _0^t \nabla vds\right) ^2, \end{aligned}$$
(3.5)

and \(\mathcal {A}_0^{-1}=D\eta _0,\) combining Lemmas 2.2, 3.2 with (3.1) , we have

$$\begin{aligned} \Vert J\mathcal {A}-J_0\mathcal {A}_0\Vert _{L^\infty }\le&C\,\Vert \nabla v\Vert _{L_t^1L^\infty }^2\le C\,t \mathcal {P}(\mathfrak {C}),\quad \Vert (\mathcal {A}_0^{-1},\mathcal {A}_0)\Vert _{L^\infty }\le C\,(1+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})), \end{aligned}$$
(3.6)

which imply that

$$\begin{aligned} \Vert \mathbb {D}^0_{J\mathcal {A}-J_0\mathcal {A}_0}(v)\Vert _{L^2}^2&\le C\,\Vert J\mathcal {A}-J_0\mathcal {A}_0\Vert _{L^\infty }^2\Vert \nabla v\Vert _{L^2}^2\le t^2\mathcal {P}(\mathfrak {C}) \Vert \nabla v\Vert _{L^2}^2, \end{aligned}$$
(3.7)
$$\begin{aligned} \Vert \mathbb {D}_{J_0\mathcal {A}_0}^0(v)\Vert _{L^2}^2&\le C\,\Vert \nabla v\Vert _{L^2}^2. \end{aligned}$$
(3.8)

On the other hand, we use (3.1), the coordinate transformation from \(\Omega \) to \(\Omega (0)\) and Lemmas 2.6, 3.4 to get that

$$\begin{aligned} \int _{\Omega }|\mathbb {D}^0_{\mathcal {A}_0}(v)|^2\,J_0\,dx&=\int _{\Omega (0)}|\mathbb {D}^0(\widetilde{u})|^2\,dx\ge c_1\Vert \nabla \,\widetilde{u}\Vert _{L^2(\Omega (0))}^2-C_1\Vert \widetilde{u}\Vert _{L^2(\Omega (0))}^2\\&\ge c_0\Vert v\Vert _{H^1}^2-C_0\Vert v\Vert _{L^2}^2, \end{aligned}$$

where \(\widetilde{u}=v\circ \eta _0^{-1}\). Hence, according to (3.1) and (3.3), we obtain that

$$\begin{aligned} \Vert \mathbb {D}^0_{J_0\mathcal {A}_0} (v)\Vert _{L^2}^2\ge c_0\Vert v\Vert _{H^1}^2-C_0\Vert \overline{\rho }^\frac{1}{2} v\Vert _{L^2}^2, \end{aligned}$$

which combining with (3.7) gives rise to

$$\begin{aligned} (c_0- t^2\mathcal {P}(\mathfrak {C}) )\Vert \nabla v\Vert _{L^2}^2-C_0\Vert \overline{\rho }^\frac{1}{2} v\Vert _{L^2}^2\le \Vert \mathbb {D}_{\mathcal {A}}^0 v\Vert _{L^2}^2\le (C_0+ t^2\mathcal {P}(\mathfrak {C}) )\Vert \nabla v\Vert _{L^2}^2, \end{aligned}$$

which we complete the first result. For the second one, we deduce

$$\begin{aligned} \int _\Omega J\mathbb {S}_\mathcal {A}v: \nabla _{\mathcal {A}}v dx&=\frac{1}{2}\int _{\Omega }(\frac{\mu }{2}|\mathbb {D}^0_{\mathcal {A}} v|^2+(\lambda +\frac{2}{3}\mu ) |\nabla _\mathcal {A}\cdot v|^2)\,Jdx\\&\ge c_1\Vert \mathbb {D}^0_{\mathcal {A}} v\Vert _{L^2}^2\ge (c_0c_1- t^2\mathcal {P}(\mathfrak {C}) )\Vert \nabla v\Vert _{L^2}^2-C_0\Vert \overline{\rho }^\frac{1}{2} v\Vert _{L^2}^2, \end{aligned}$$

here we used (3.1) in the last step and assumption \(\mu >0,~\lambda +\frac{2}{3}\mu \ge 0\). Combining with the first result, we finish this proof. \(\square \)

3.1 Zeroth-order estimate of v

Now, we are in a position to give a priori estimates. First, multiplying by v on the first equation of (1.18) and integrating over \(\Omega \), from the Piola identity (1.21) and boundary conditions, we get the basic energy estimate:

Proposition 3.6

Assume v is a smooth solution of system (1.18) on [0, T]. Then, we have

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\left( \int _{\Omega }\overline{\rho }|v|^2dx+2\int _{\Omega }\overline{\rho }^2J^{-1}dx\right) +\frac{1}{2}\int _{\Omega }\left( \frac{\mu }{2}|\mathbb {D}^0_{\mathcal {A}} v|^2+\left( \lambda +\frac{2}{3}\mu \right) |\nabla _\mathcal {A}\cdot v|^2\right) \,Jdx=v0. \end{aligned}$$

3.2 First-order estimate of v

Here, to get the higher regularity of the v. We multiply \(\partial _t v\) on the both sides of (1.18) to get that

Proposition 3.7

Assume that (3.1) holds and v is a smooth solution of system (1.18) on [0, T], then there holds that for \(t\in [0, T]\)

$$\begin{aligned}&\frac{1}{2}\frac{d}{dt}\int _{\Omega } \left( \frac{\mu }{2}|\mathbb {D}^0_{\mathcal {A}} v|^2+\left( \lambda +\frac{2}{3}\mu \right) |\nabla _\mathcal {A}\cdot v|^2\right) \,J\,dx+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}^2\\&\quad \le ( C +t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(\mathfrak {D}(v) \Vert \nabla v\Vert _{L^2}^2 +1). \end{aligned}$$

Proof

Taking \(L^2\) product with \(\partial _t v\) to the first equation of (1.18) to get that

$$\begin{aligned} \Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}^2+\int _{\Omega }\nabla _{J\mathcal {A}}(\overline{\rho }^2J^{-2})\cdot \partial _t vdx-\int _{\Omega }\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}} (v) \cdot \partial _t vdx=0. \end{aligned}$$

Due to the Piola identity (1.21) and the boundary condition \(\mathrm {S}_{\mathcal {A}}(v)\cdot \mathcal {N}|_{x_3=1}=0\) and \(v|_{x_3=0}=0\), integration by parts yields

$$\begin{aligned} -\int _{\Omega }\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}} (v)\cdot \partial _t vdx=&\int _{\Omega }\mathbb {S}_{J\mathcal {A}} (v): \partial _t(\nabla _\mathcal {A}v)dx-\int _{\Omega }\mathbb {S}_{J\mathcal {A}}(v):\nabla _{\partial _t\mathcal {A}} vdx. \end{aligned}$$

Since \(\mathbb {D}_\mathcal {A}(v)\) and \((\nabla _\mathcal {A}\cdot v)\mathbb {I}\) are symmetric, it implies that

$$\begin{aligned}&\int _{\Omega }\mathbb {S}_{J\mathcal {A}} (v): \partial _t(\nabla _\mathcal {A}v)dx=\int _{\Omega }(\mu \mathbb {D}_{J\mathcal {A}}(v)+\lambda (\nabla _{J\mathcal {A}} \cdot v)\mathbb {I}): \partial _t(\nabla _\mathcal {A}v)dx\\&\quad =\frac{\mu }{2}\int _{\Omega }\mathbb {D}_{J\mathcal {A}}(v): \partial _t\mathbb {D}_\mathcal {A}(v)dx+\lambda \int _{\Omega }\nabla _{J\mathcal {A}} \cdot v~\partial _t(\nabla _\mathcal {A}\cdot v)\\&\quad =\frac{1}{2}\frac{d}{dt}\int _{\Omega }J \Big (\frac{\mu }{2}|\mathbb {D}^0_{\mathcal {A}} v|^2+\Big (\lambda +\frac{2}{3}\mu \Big ) |\nabla _\mathcal {A}\cdot v|^2\Big )\,dx\\&\qquad -\frac{1}{2}\int _{\Omega }\partial _tJ\Big (\frac{\mu }{2} |\mathbb {D}_\mathcal {A}(v)|^2+\lambda |\nabla _\mathcal {A}\cdot v|^2\Big )dx\\&\quad =\frac{1}{2}\frac{d}{dt}\int _{\Omega }\mathbb {S}_{J\mathcal {A}} (v): \nabla _{\mathcal {A}}v\,dx-\frac{1}{2}\int _{\Omega }\mathbb {S}_\mathcal {A}(v): \nabla _{\mathcal {A}}v\partial _t J\,dx, \end{aligned}$$

which gives that

$$\begin{aligned} -\int _{\Omega }\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}} (v)\cdot \partial _t vdx&=\frac{1}{2}\frac{d}{dt}\int _{\Omega }J \left( \frac{\mu }{2}|\mathbb {D}^0_{\mathcal {A}} v|^2+\left( \lambda +\frac{2}{3}\mu \right) |\nabla _\mathcal {A}\cdot v|^2\right) \,dx\\&\quad -\frac{1}{2}\int _{\Omega }\mathbb {S}_\mathcal {A}(v): \nabla _{\mathcal {A}}v\partial _t J\,dx -\int _{\Omega }\mathbb {S}_{J\mathcal {A}}(v):\nabla _{\partial _t\mathcal {A}} v\,dx. \end{aligned}$$

To estimate the last two terms of right hand of the above equation, we recall that formula (1.19)–(1.20), Lemmas 2.2 and 3.2 to get that

$$\begin{aligned} \Vert \partial _tJ,\,\partial _t \mathcal {A}\Vert _{L^\infty }\le C\,\Vert \mathcal {A}\Vert _{L^\infty }^2\Vert \nabla v\Vert _{L^\infty }\le (C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\mathfrak {D}(v), \end{aligned}$$

which implies that

$$\begin{aligned}&\left| \int _{\Omega }\mathbb {S}_\mathcal {A}(v): \nabla _{\mathcal {A}}v\partial _t J\,dx\right| +\left| \int _{\Omega }\mathbb {S}_{J\mathcal {A}}(v):\nabla _{\partial _t\mathcal {A}} v\,dx\right| \\&\quad \le ( C +t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\mathfrak {D}(v)\Vert J\mathcal {A}\Vert _{L^\infty }\Vert \nabla v\Vert _{L^2}^2\le (C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\mathfrak {D}(v) \Vert \nabla v\Vert _{L^2}^2. \end{aligned}$$

For the pressure term, we notice it contains \(\overline{\rho }^2\). Thus, we have

$$\begin{aligned} \overline{\rho }^{-\frac{1}{2}}\nabla _{J\mathcal {A}}(\overline{\rho }^2J^{-2})=\overline{\rho }^{-\frac{1}{2}}\partial _k(J^{-1}\mathcal {A}^k_i\overline{\rho }^2)=\overline{\rho }^{-\frac{1}{2}}\Big ( J^{-1}\mathcal {A}^k_i\partial _k\overline{\rho }^2+\partial _k(J^{-2})J\mathcal {A}^k_i\overline{\rho }^2 \Big ), \end{aligned}$$

which implies that for all \(t\in [0, T]\), we have

$$\begin{aligned} \Vert \overline{\rho }^{-\frac{1}{2}}\nabla _{J\mathcal {A}}(\overline{\rho }^2J^{-2})\Vert _{L^2}\le & {} \Vert \overline{\rho }^{\frac{1}{2}}\overline{\rho }'\Vert _{L^\infty }(C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})) \Vert \mathcal {A}\Vert _{L^2}\\&+(C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Vert ZJ\Vert _{L^\infty }\Vert J\mathcal {A}\Vert _{L^2}\\\le & {} C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}), \end{aligned}$$

where we used Lemma 2.4. Thus, by Hölder’s inequality, we get

$$\begin{aligned} \Big |\int _{\Omega }\nabla _{J\mathcal {A}}(\overline{\rho }^2J^{-2})~\partial _t vdx\Big |\le & {} \Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}\Vert \overline{\rho }^{-\frac{1}{2}}\nabla _{J\mathcal {A}}(\overline{\rho }^2J^{-2})\Vert _{L^2}\\\le & {} (C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}\le C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})+ \frac{1}{2}\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}^2. \end{aligned}$$

This ends the proof of Proposition 3.7. \(\square \)

3.3 High-order estimates of v

In this subsection, we use the conormal derivative to get the regularity of the horizontal direction. For this, we recall the conormal Sobolev space with a parameter \(\delta \) introduced by Masmoudi and Rousset [34].

$$\begin{aligned} \Vert f\Vert _{X^N_{\alpha , \delta }}^2:=\sum _{|m|=0}^N\delta ^{2|m|} \Vert \overline{\rho }^{\alpha } Z^m f\Vert _{L^2}^2,\quad \Vert f\Vert _{\dot{X}^N_{\alpha , \delta }}^2:=\sum _{ |m|=1}^N\delta ^{2|m|}\Vert \overline{\rho }^{\alpha } Z^m f\Vert _{L^2}^2, \end{aligned}$$

where \(\delta \) is a small positive constant which will be determined later on and \(\alpha \in \mathbb {R}\). In particular, when \(\delta =1\), the spaces \(X^N_{\alpha , \delta }\) and \(\dot{X}^N_{\alpha , \delta }\) will be denoted by \(X^N_{\alpha }\) and \(\dot{X}^N_{\alpha }\) respectively for simplicity.

For \(T>0\), \(\delta >0\), and \(t \in [0, T]\), we define the modified instantaneous energy \(\mathcal {E}_{\delta }(t)\) (in terms to the velocity v)

and the modified dissipation \(\mathcal {D}_{\delta }(t)\)

In particular, if \(\delta =1\), then \(\mathcal {E}_{\delta }(t)\) and \(\mathcal {D}_{\delta }(t)\) become the usual instantaneous energy \(\mathcal {E}(t)\) and the dissipation \(\mathcal {D}(t)\) respectively.

Let’s now state our main results of this subsection:

Proposition 3.8

Assume that (3.1) holds and v is a smooth solution of system (1.18) on [0, T], then it holds that

$$\begin{aligned}&\frac{d}{dt}\Vert v\Vert _{X^{12}_{\frac{1}{2}, \delta }}^2+\Big (c_0-\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Big )\Vert \nabla v\Vert _{X^{12}_{0, \delta }}^2\\&\quad \le t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})\mathfrak {D}^2(v)+C_0\Vert v\Vert _{X^{12}_{\frac{1}{2}, \delta }}^2+C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}), \end{aligned}$$

where the positive constants \(c_0\) and \(C_0\) are independent of \(\delta \), and \(\mathcal {P}(\mathfrak {C})\) may depend on \(\delta \).

Proof

Acting \(Z^m\) on the first equation of (1.18) and taking \(L^2\) inner product with \(\delta ^{2|m|}Z^mv\), then summing \(\sum _{|m|=0}^{12}\), we obtain

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\Vert v\Vert _{X^{12}_{\frac{1}{2}, \delta }}^2-\sum _{|m|=0}^{12}\delta ^{2|m|}\int _{\Omega }Z^m\big (\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}v\big )\cdot Z^m v\,dx= I_1+I_2 \end{aligned}$$

with

$$\begin{aligned} I_1&=\sum _{|m|=1}^{12}\delta ^{2|m|} \int _{\Omega }[\overline{\rho },Z^m]\partial _t v\cdot Z^m v \, dx,\quad \,\\ I_2&=-\sum _{|m|=0}^{12}\delta ^{2|m|}\int _{\Omega }Z^m\big (\nabla _{J\mathcal {A}}(J^{-2}\overline{\rho }^2)\big )\cdot Z^m v \,dx. \end{aligned}$$

Estimate of dissipation term. For the dissipation term, by using integration by parts, we split it into three parts:

$$\begin{aligned}&-\sum _{|m|=0}^{12}\delta ^{2|m|} \int _{\Omega }Z^m\big (\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}v\big )\cdot Z^m vdx\\&\quad =\sum _{|m|=0}^{12}\delta ^{2|m|} \int _\Omega J\mathbb {S}_\mathcal {A}(Z^mv): \nabla _{\mathcal {A}}Z^m v \,dx+\sum _{|m|=1}^{12}\delta ^{2|m|} \int _{\Omega }[Z^m,\mathbb {S}_{\mathcal {A}}]v: \nabla _{J\mathcal {A}}(Z^m v) dx\\&\qquad -\sum _{|m|=1}^{12}\delta ^{2|m|} \bigg (\int _{x_3=1} \mathcal {N}\cdot Z_h^m\mathbb {S}_{\mathcal {A}}v\cdot Z_h^m v dS+ \int _{\Omega }[Z^m,\nabla _{J\mathcal {A}}]\cdot \mathbb {S}_{\mathcal {A}}v \cdot Z^m v \,dx\bigg )\\&\quad =:I_3+I_4+I_5. \end{aligned}$$

Next, we deal with the commutators \(I_3\), \(I_4\) and \(I_5\) step by step.

  • \({ {Estimates\, of\, } I_3}\). Thanks to Lemma 3.5, one can see that for any \(m:|m|=0, 1,\ldots ,12\)

    $$\begin{aligned}&\int _\Omega J\mathbb {S}_\mathcal {A}(Z^mv): \nabla _{\mathcal {A}}Z^m v \,dx\\&\quad =\int _{\Omega } \left( \frac{\mu }{4}|\mathbb {D}^0_{\mathcal {A}} Z^mv|^2+\frac{\lambda +\frac{2}{3}\mu }{2} |\nabla _\mathcal {A}\cdot Z^mv|^2\right) \,J\,dx\\&\quad \ge c_1\Vert \mathbb {D}^0_{\mathcal {A}} Z^mv\Vert _{L^2}^2 \ge c_1\bigg ((c_0- t^2 \mathcal {P}(\mathfrak {C}))\Vert \nabla (Z^mv)\Vert _{L^2}^2-C_0\Vert \overline{\rho }^\frac{1}{2} Z^mv\Vert _{L^2}^2\bigg ), \end{aligned}$$

which implies

$$\begin{aligned}&\sum _{|m|=0}^{12}\delta ^{2|m|}\int _\Omega J\mathbb {S}_\mathcal {A}(Z^mv): \nabla _{\mathcal {A}}Z^m v \,dx\ge \sum _{|m|=0}^{12}\delta ^{2|m|}c_1\bigg (\frac{1}{2}(c_0- t^2 \mathcal {P}(\mathfrak {C}))\Vert Z^m\nabla \,v\Vert _{L^2}^2\nonumber \\&\quad -\frac{1}{2}(c_0+ t^2 \mathcal {P}(\mathfrak {C}))\Vert [\nabla ,Z^m]v\Vert _{L^2}^2-C_0\Vert \overline{\rho }^\frac{1}{2} Z^mv\Vert _{L^2}^2\bigg ), \end{aligned}$$
(3.9)

For \(|m|\ge 1\), by a direct calculation, we have

$$\begin{aligned}{}[\nabla ,Z^m]=m\nabla \overline{\rho }Z^{m-1}\partial _{3}, \end{aligned}$$
(3.10)

which implies that

$$\begin{aligned} \sum _{|m|=1}^{12}\delta ^{2|m|}\frac{c_1}{2}(c_0+ t^2 \mathcal {P}(\mathfrak {C}))\Vert [\nabla ,Z^m]v\Vert _{L^2}^2\le (C_0+ t^2 \mathcal {P}(\mathfrak {C}))\delta ^2\Vert \nabla v\Vert _{X^{11}_{0, \delta }}^2. \end{aligned}$$
(3.11)

Plugging (3.11) into (3.9) shows

$$\begin{aligned}&\sum _{|m|=0}^{12}\delta ^{2|m|}\int _\Omega J\mathbb {S}_\mathcal {A}(Z^mv): \nabla _{\mathcal {A}}Z^m v \,dx\\&\quad \ge (c_2- t^2 \mathcal {P}(\mathfrak {C}))\Vert \nabla \,v\Vert _{X^{12}_{0, \delta }}^2-(C_0+ t^2 \mathcal {P}(\mathfrak {C}))\delta ^2\Vert \nabla v\Vert _{X^{11}_{0, \delta }}^2-C_0\Vert v\Vert _{X^{12}_{\frac{1}{2}, \delta }}^2. \end{aligned}$$
  • \({ {Estimates\, of\, } I_4}\). For \(|m|\ge 1,\) by a direct calculation, we have

    $$\begin{aligned}{}[Z^m,\mathbb {D}_{\mathcal {A}}]v&= Z^m\Big (\mathcal {A}_{i}^k\partial _kv_j+\mathcal {A}_{j}^k\partial _kv_i \Big )-\Big (\mathcal {A}_{i}^k\partial _k(Z^mv_j)+\mathcal {A}_{j}^k\partial _k(Z^mv_i) \Big )\\&=\mathcal {A}_{i}^k[Z^m,\partial _k]v_j+\mathcal {A}_{j}^k[Z^m,\partial _k]v_i\\&\quad +\sum _{\begin{array}{c} |m_1|+|m_2|=|m|,\\ |m_1|\ge 1 \end{array}}(Z^{m_1} \mathcal {A}_{i}^kZ^{m_2}\partial _kv_j+Z^{m_1}\mathcal {A}_{j}^kZ^{m_2}\partial _kv_i)\\&=m\partial _k \overline{\rho }\mathcal {A}_{i}^3Z^{m-1}\partial _3v_j+m\partial _k \overline{\rho }\mathcal {A}_{j}^3Z^{m-1}\partial _3v_i\\&\quad +\sum _{\begin{array}{c} \begin{array}{c} |m_1|+|m_2|=|m|,\\ |m_1|\ge 1 \end{array} \end{array}}(Z^{m_1} \mathcal {A}_{i}^kZ^{m_2}\partial _kv_j+Z^{m_1}\mathcal {A}_{j}^kZ^{m_2}\partial _kv_i). \end{aligned}$$

By Lemmas 2.5, 3.2, we have

$$\begin{aligned} \delta ^{|m|}\Vert [Z^m,\mathbb {D}_{\mathcal {A}}]v\Vert _{L^2}\le & {} \delta \Big ((C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Vert \nabla v\Vert _{X^{11}_{0, \delta }}+C_0\Vert \nabla v\Vert _{X^{11}_{0, \delta }}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)\Big )\\\le & {} \delta \Big ( C_0\Vert \nabla v\Vert _{X^{11}_{0, \delta }}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)\Big ). \end{aligned}$$

By the same argument, we have

$$\begin{aligned} \delta ^{|m|}\Vert [Z^m,\text{ div }_{\mathcal {A}}]v\Vert _{L^2}\le \delta \Big ( C_0\Vert \nabla v\Vert _{X^{11}_{0, \delta }}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)\Big ). \end{aligned}$$

Combining the above two estimates, we have

$$\begin{aligned} I_4\le&\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(\delta \Vert \nabla v\Vert _{X^{11}_{0, \delta }}+\Vert \nabla v\Vert _{X^{12}_{0, \delta }})(C_0\Vert \nabla v\Vert _{X^{11}_{0, \delta }}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v))\\ \le&\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Vert \nabla v\Vert _{X^{12}_{0, \delta }}^2+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)^2. \end{aligned}$$
  • \({ {Estimates\, of \,} I_5}\). A direct calculation gives that

    $$\begin{aligned}{}[Z^m,\nabla _{J\mathcal {A}}]\cdot \mathbb {S}_{\mathcal {A}}v&=Z^m(J\mathcal {A}_i^k\partial _k(\mathbb {S}_{\mathcal {A}} v)^i)-\partial _k(J\mathcal {A}_i^k(Z^m(\mathbb {S}_{\mathcal {A}} v))^i)\nonumber \\&=\partial _k\Big (Z^m(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} v)^i)-J\mathcal {A}_i^k(Z^m(\mathbb {S}_{\mathcal {A}} v))^i\Big )+[Z^m,\partial _k](J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} v)^i). \end{aligned}$$
    (3.12)

For the commutator term, we see

$$\begin{aligned}{}[Z^m,\partial _3]=-m\partial _3\overline{\rho }Z^{m-1}\partial _3\sim Z^{m-1}\partial _3,~[Z^m,\partial _h]=-m\partial _h\overline{\rho }Z^{m-1}\partial _3\sim Z^{m}, \end{aligned}$$
(3.13)

where we used (1.13). Then one has

$$\begin{aligned}&\Big |\int _{\Omega }[Z^m,\partial _k](J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} v)^i)\cdot Z^m v dx\Big |\\&\quad \le C_0\Big |\int _{\Omega }Z^{m-1}\partial _3(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} v)^i)\cdot Z_3 Z^{m-1} v dx\Big |\\&\qquad +C_0\Big |\int _{\Omega }Z^{m}(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} v)^i)\cdot Z^m v dx\Big |\le C_0\Big |\int _{\Omega } Z^{m}(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} v)^i)\cdot Z^{m-1}\nabla vdx\Big |, \end{aligned}$$

which combining with Lemma 2.5 follows

$$\begin{aligned}&\Big |\sum _{|m|=1}^{12}\delta ^{2|m|}\int _{\Omega }[Z^m,\partial _k](J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} v)^i)\cdot Z^m v dx\Big |\\&\quad \le \delta ( C_0\Vert \nabla v\Vert _{X^{12}_{0, \delta }}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)) \Vert \nabla v\Vert _{X^{11}_{0, \delta }}. \end{aligned}$$

Now, we deal with the first term of the right hand of (3.12). By using integration by parts, one has

$$\begin{aligned}&\sum _{|m|=1}^{12}\delta ^{2|m|} \int _{\Omega }\partial _k\Big (Z^m(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} (v))^i)-J\mathcal {A}_i^k(Z^m(\mathbb {S}_{\mathcal {A}} (v)))^i\Big )\cdot Z^m v dx\\&\quad =-\sum _{|m|=1}^{12}\delta ^{2|m|} \int _{\Omega }\Big (Z^m(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}} (v))^i)-J\mathcal {A}_i^k(Z^m(\mathbb {S}_{\mathcal {A}}(v)))^i\Big )\cdot \partial _kZ^m v dx\\&\qquad +\sum _{|m|=1}^{12}\delta ^{2|m|}\int _{x_3=1}\Big (Z_h^m(J\mathcal {A}_i^3e_3(\mathbb {S}_{\mathcal {A}}(v))^i)-J\mathcal {A}_i^3e_3(Z_h^m(\mathbb {S}_{\mathcal {A}} (v)))^i\Big )\cdot Z_h^m vdS. \end{aligned}$$

Because of \(\mathbb {S}_{\mathcal {A}} (v) \mathcal {N}=0\) on the boundary \(\{x_3=1\}\), \(J\mathcal {A}_i^3e_3= \mathcal {N}\), and \(Z_h^m(\mathbb {S}_{\mathcal {A}} v \mathcal {N})=0\) on \(\{x_3=1\}\), the second term on the above equality plus the second term of \(I_5\) is zero:

$$\begin{aligned}&\sum _{|m|=1}^{12}\delta ^{2|m|}\int _{x_3=1}\Big (Z_h^m(\mathcal {N}(\mathbb {S}_{\mathcal {A}}(v)))-\mathcal {N}(Z_h^m(\mathbb {S}_{\mathcal {A}} (v)))\Big )~Z_h^m vdS\\&\quad +\sum _{|m|=1}^{12}\delta ^{2|m|} \int _{x_3=1}\mathcal {N}Z_h^m\mathbb {S}_{\mathcal {A}}(v)~Z_h^m v dS=0. \end{aligned}$$

Hence, all we left is to deal with the commutator

$$\begin{aligned} \int _{\Omega }\Big (Z^m(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}}(v))^i)-J\mathcal {A}_i^k(Z^m(\mathbb {S}_{\mathcal {A}}(v)))^i\Big )\cdot \partial _kZ^m v dx. \end{aligned}$$

By the same arguments as \(I_4\) and using Lemma 2.22.5, we deduce that

$$\begin{aligned}&\Big | \sum _{|m|=1}^{12}\delta ^{2|m|}\int _{\Omega }\Big (Z^m(J\mathcal {A}_i^k(\mathbb {S}_{\mathcal {A}}(v))^i)-J\mathcal {A}_i^k(Z^m(\mathbb {S}_{\mathcal {A}} (v)))^i\Big )\cdot \partial _kZ^m vdx \Big | \\&\quad \le \delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Vert \nabla v\Vert _{X^{12}_{0, \delta }}^2+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)^2. \end{aligned}$$

Combining all the above estimates, we get that

$$\begin{aligned} I_5\le \delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Vert \nabla v\Vert _{X^{12}_{0, \delta }}^2+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)^2. \end{aligned}$$

So far, we obtain

$$\begin{aligned}&-\sum _{|m|=0}^{12}\delta ^{2|m|}\int _{\Omega }Z^m\big (\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}v\big )\cdot Z^m vdx\\&\quad \ge \Big (c_2-\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Big )\Vert \nabla v\Vert _{X^{12}_{0, \delta }}^2-C_0\Vert v\Vert _{X^{12}_{\frac{1}{2}, \delta }}^2-t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)^2. \end{aligned}$$

Estimate of \(I_2\). Now, we deal with the pressure.

$$\begin{aligned} I_2=&\sum _{|m|=0}^{12}\delta ^{2|m|}\int _{\Omega }\partial _kZ^m\big (\mathcal {A}_i^kJ^{-1}\overline{\rho }^2\big )\cdot Z^m v^idx\\&+\sum _{|m|=1}^{12}\delta ^{2|m|}\int _{\Omega }[Z^m,\partial _k]\big (\mathcal {A}_i^kJ^{-1}\overline{\rho }^2\big )\cdot Z^mv^i\,dx \triangleq I_{21}+I_{22}. \end{aligned}$$
  • \({ {Estimates\, of\, } I_{22}}\). Since \(Z^m \overline{\rho }^2\sim \overline{\rho }^2\) for any m, we use (3.10) and Lemmas 2.32.4 to get

    $$\begin{aligned} I_{22} \le \delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Vert \nabla v\Vert _{X^{11}_{0, \delta }}. \end{aligned}$$
  • \({ {Estimates\, of\, } I_{21}}\). Because of \(\overline{\rho }|_{x_3=1}=0,\) the boundary terms vanish when we integrate by parts. By the same argument as \(I_5\), it is easy to see \(I_{21}\) is bounded by

    $$\begin{aligned} I_{21}\le (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(\Vert \nabla v\Vert _{X^{12}_{0, \delta }}+\delta \Vert \nabla v\Vert _{X^{11}_{0, \delta }}). \end{aligned}$$

Combining the two estimates, we get

$$\begin{aligned} I_2\le (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(\Vert \nabla v\Vert _{X^{12}_{0, \delta }}+\delta \Vert \nabla v\Vert _{X^{11}_{0, \delta }}). \end{aligned}$$

Estimate of \(I_1\). For \(m\ge 1,\) it holds that

$$\begin{aligned}{}[\overline{\rho },Z^m]\sim \sum _{k=0}^{m-1} f_kZ^{k}(\overline{\rho }\cdot ) . \end{aligned}$$

where \(f_k\) are smooth functions which are defined by \(\overline{\rho }\). Thus

$$\begin{aligned} I_1&\le C_0\sum _{\begin{array}{c} |m|=1 \end{array}}^{12}\sum _{k=0}^{m-1}\delta ^{2|m|} \Big |\int _{\Omega }Z^k(\overline{\rho }\partial _t v)\cdot Z^m vdx\Big |\\&=C_0\sum _{\begin{array}{c} |m|=1 \end{array}}^{12}\sum _{k=0}^{m-1}\delta ^{2|m|} \Big |\int _{\Omega }Z^k(-\nabla _{J\mathcal {A}}(J^{-2}\overline{\rho }^2)+\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}v)\cdot Z^m vdx\Big |. \end{aligned}$$

From the formula above, \(I_1\) can be regarded as lower term to \(I_2\) plus dissipation term with the highest order 11. Since \(k\le m-1,\) extra \(\delta \) is left. Thus, we have

$$\begin{aligned} I_1&\le \delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))( \Vert \nabla v\Vert _{X^{12}_{0, \delta }} + \delta \Vert \nabla v\Vert _{X^{11}_{0, \delta }})\\&\quad +\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(C_0\Vert \nabla v\Vert _{X^{11}_{0, \delta }}+t^\frac{1}{2}\mathcal {P}(\mathfrak {C})\mathfrak { D}(v)) \Vert \nabla v\Vert _{X^{12}_{0, \delta }}. \end{aligned}$$

Collecting all estimates together, we finally obtain

$$\begin{aligned}&\frac{d}{dt}\Vert v\Vert _{X^{12}_{\frac{1}{2}, \delta }}^2+\Big (c_0-\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Big )\Vert \nabla v\Vert _{X^{12}_{0, \delta }}^2\\&\quad \le t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})\mathfrak {D}^2(v)+C_0\Vert v\Vert _{X^{12}_{\frac{1}{2}, \delta }}^2+C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}), \end{aligned}$$

which implies the desired results. \(\square \)

3.4 Estimate for \(\mathfrak {D}(v)\)

To close the energy estimates, all we left is the estimate of \(\mathfrak {D}(v)\) which should be controlled by the energy.

Lemma 3.9

Assume that (3.1) holds. Then there exists \(0<T\le \bar{T}\) and \(\delta _0>0\) which depend on the initial data, \(\sigma _0\) and \(\mathfrak {C}\) such that for any \(t\in [0,T]\) and \(\delta \in (0,\delta _0)\), there holds that

$$\begin{aligned} \mathfrak {D}(v)\le C\mathcal {D}^{\frac{1}{2}}_{\delta }(t)+(C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(1+t^\frac{1}{2}\mathfrak {D}(v)). \end{aligned}$$

Proof

Here we only need to control the term \(\Vert \overline{\rho }^{\kappa -\frac{1}{2}}\triangle v\Vert _{L^2}\). To do that, we go back to the equation of v. Since

$$\begin{aligned} \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v)&=\overline{\rho }^{-\frac{1}{2}+\kappa }\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}(v)+\overline{\rho }^{-\frac{1}{2}+\kappa }\big (\nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v)-\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}(v)\big )\\&=\overline{\rho }^{-\frac{1}{2}+\kappa }\Big (-\overline{\rho }\partial _t v-\nabla _{J\mathcal {A}}(J^{-2}\overline{\rho }^2) \Big )\\&\quad +\overline{\rho }^{-\frac{1}{2}+\kappa }\big (\nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v)-\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}}(v)\big ), \end{aligned}$$

which implies that

$$\begin{aligned} \Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v)\Vert _{L^2}&\le \Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla _{J\mathcal {A}}(J^{-2}\overline{\rho }^2)\Vert _{L^2}\\&\quad +\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla _{J_0\mathcal {A}_0-J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}_0}(v)\Vert _{L^2}\\&\quad +\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla _{J\mathcal {A}}\cdot \mathbb {S}_{\mathcal {A}_0-\mathcal {A}}(v)\Vert _{L^2}\\&\triangleq \Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}+I_1+I_2+I_3. \end{aligned}$$

Owing to Lemma 2.4, we have

$$\begin{aligned} I_1 \le C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}). \end{aligned}$$

For \(I_2\), by Lemma 2.1, Lemma 2.4 and (3.5)–(3.6), we have

$$\begin{aligned} I_2&\le \Vert J_0\mathcal {A}_0-J\mathcal {A}\Vert _{L^\infty }\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla \cdot \mathbb {S}_{\mathcal {A}_0}(v)\Vert _{L^2}\\&\le t^\frac{1}{2}(C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla v\Vert _{L^2}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla ^2 v\Vert _{L^2})\\&\le t^\frac{1}{2}(C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(\Vert \overline{\rho }^{-\frac{1}{2}+\kappa } \triangle v\Vert _{L^2}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa } Z_h \partial _3 v\Vert _{L^2}+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa } Z_h^2 v\Vert _{L^2}+\mathfrak {D}(v))\\&\le t^\frac{1}{2}(C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(\Vert \overline{\rho }^{-\frac{1}{2}+\kappa } \triangle v\Vert _{L^2}+\Vert \overline{\rho }^{\frac{1}{2}+\kappa } Z_h \triangle v\Vert _{L^2}+\Vert \overline{\rho }^{\frac{1}{2}+\kappa } Z_h^2 \nabla v\Vert _{L^2}+\mathfrak {D}(v))\\&\le t^\frac{1}{2}(C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\mathfrak {D}(v). \end{aligned}$$

Similarly, by the fact that

$$\begin{aligned} \mathcal {A}-\mathcal {A}_0=(\mathcal {A}J-\mathcal {A}_0 J_0)J^{-1}+J^{-1}(J_0-J)\mathcal {A}_0 \end{aligned}$$

and

$$\begin{aligned} J-J_0=\int _0^t \partial _t Jds=\int _0^t J\nabla _{\mathcal {A}}vds, \end{aligned}$$

combine (3.5) with Lemma 3.2 to get

$$\begin{aligned} I_3\le t^\frac{1}{2} (C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\mathfrak {D}(v). \end{aligned}$$

Collecting all above estimates to obtain

$$\begin{aligned} \Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v)\Vert _{L^2}\le&\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}+(C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(1+t^\frac{1}{2}\mathfrak {D}(v)). \end{aligned}$$
(3.14)

Next, we give the relationship between \(\triangle v\) and \(\nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v)\). It is easy to find that

$$\begin{aligned} \nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v) =\left( \begin{array}{c}\mu J^{-1}_0\partial _3^2 v_1 \\ \mu J^{-1}_0\partial _3^2 v_2\\ (2\mu +\lambda ) J^{-1}_0\partial _3^2 v_3 \end{array}\right) +\text{ some } \text{ terms } \text{ likes }~ Z\nabla v. \end{aligned}$$

By Lemma 2.1 and the interpolation inequality, we have

$$\begin{aligned} \Vert \overline{\rho }^{-\frac{1}{2}+\kappa } Z_h\nabla v\Vert _{L^2}&\le C_0\Vert \overline{\rho }^{\frac{1}{2}+\kappa } Z_h\nabla v\Vert _{L^2}+C_0\Vert \overline{\rho }^{\frac{1}{2}+\kappa } Z_h\nabla ^2 v\Vert _{L^2}\nonumber \\&\le C_0\Vert \nabla v\Vert _{L^2_{x_3}(H^2_h)}+C_0\Vert \overline{\rho }^{-\frac{1}{2}+\kappa } \triangle v\Vert _{L^2}^\theta \Vert \overline{\rho }\triangle v\Vert _{L^2}^{1-\theta }\nonumber \\&\le C_\varepsilon \Vert \nabla v\Vert _{L^2_{x_3}(H^2_h)}+\varepsilon \Vert \overline{\rho }^{-\frac{1}{2}+\kappa } \triangle v\Vert _{L^2}, \end{aligned}$$
(3.15)

where we use Young inequality in the last step and \(\theta \in (0,1).\)

Taking \(\varepsilon \) small enough and using (3.1), (3.14), (3.15), we have

$$\begin{aligned} \Vert \overline{\rho }^{-\frac{1}{2}+\kappa } \triangle v\Vert _{L^2}\le \Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2}+(C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))(1+t^\frac{1}{2}\mathfrak {D}(v))+\frac{C_0}{\delta ^2}\Vert \nabla v\Vert _{X^{12}_{0, \delta }}. \end{aligned}$$
(3.16)

Combining (3.14) and (3.16), we obtain the desired results. \(\square \)

3.5 Proof of Proposition 3.1

Now, from Propositions 3.7 to 3.8, we obtain that

$$\begin{aligned}&\Big (c_0-\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Big )\Big (\sup _{\tau \in [0, t]} \mathcal {E}_{\delta }(\tau ) + \int _{0}^t \mathcal {D}_{\delta }(\tau )\Big )\\&\quad \le \Big (c_0-\delta (C_0+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C}))\Big ) \mathcal {E}(0)+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})(1+ \sup _{\tau \in [0, t]} \mathcal {E}_{\delta }(\tau )). \end{aligned}$$

Now, we give the estimates of J. By the definition of J, we have

$$\begin{aligned} J-J_0=\int _0^t \partial _t Jds=\int _0^t J\nabla _{\mathcal {A}}vds, \end{aligned}$$

which implies that

$$\begin{aligned} |J-J_0|\le \Vert J\Vert _{L^\infty }\Vert \mathcal {A}\Vert _{L^\infty }\Vert \nabla v\Vert _{L^1_tL^\infty }\le t^\frac{1}{2}(C+t^{\frac{1}{2}}\mathcal {P}(\mathfrak {C})). \end{aligned}$$

Then by the Lemma 3.9 and standard bootstrap argument, we finish the proof of Proposition 3.1.

4 Local well-posedness

In this section, we will first give existence and uniqueness of strong solutions of system (1.18), which is motivated by the method in [10]. First, we give some definitions of functional spaces. Given \(T>0,\) let \(\widetilde{Y}_T\) and \({Y}_T\) are defined by

$$\begin{aligned} \widetilde{Y}_T&\triangleq C([0, T], X^{0}_{\frac{1}{2}})\cap \,L^2([0, T],\,H^1),\\ Y_T&\triangleq \{v\in \widetilde{Y}_T\cap C([0, T], X^{12}_{\frac{1}{2}}\cap \,H^1) :\,\, \Vert v\Vert _{Y_T}<+\infty \}, \end{aligned}$$

where \(\Vert v\Vert _{\widetilde{Y}_T}^2:=\sup _{t\in [0,T]}\Vert \overline{\rho }^{\frac{1}{2}}v\Vert _{L^2}^2+\Vert v\Vert _{L^2_T (H^1)}^2\) and \(\Vert v\Vert _{Y_T}^2:=\sup _{t\in [0,T]}(\Vert v\Vert _{X^{12}_{\frac{1}{2}}}^2+\Vert \nabla v\Vert _{L^2}^2 )+\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle v\Vert _{L^2_T(L^2)}^2 +\Vert \nabla v\Vert _{L_T^2(X^{12})}^2+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L_T^2(L^2)}^2\).

Now, we define map \(\Theta : Y_T\rightarrow Y_T\) as follows. For any given \(\widetilde{v}\in Y_T,\) \(v:=\Theta (\widetilde{v})\) is the solution of the following linear \(\mathcal {A}\)-equations:

$$\begin{aligned} \left\{ \begin{aligned}&\overline{\rho }\partial _t v+\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}((\widetilde{J})^{-2}\overline{\rho }^2)-\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot \mathbb {S}_{\widetilde{\mathcal {A}}}(v)=0\quad \text{ in }\quad \Omega ,\\&\mathrm {S}_{\widetilde{\mathcal {A}}}(v)~ \widetilde{\mathcal {N}}=0,\quad \text{ on }\quad \Gamma ,\\&v|_{x_3=0}=0,\\&v|_{t=0}=v_0 \quad \text{ in }\quad \Omega . \end{aligned} \right. \end{aligned}$$
(4.1)

4.1 Existence and uniqueness of the strong solution to (4.1).

Our aim in this subsection is to construct strong solutions to linear \(\mathcal {A}\)-equations (4.1).

Lemma 4.1

Assume that \(\overline{\rho }^{\frac{1}{2}} v_0,\nabla v_0\in L^2\) and \(\widetilde{v}\in Y_T,\) then there exists a positive time \(T_1 \in (0, T]\) such that the system (4.1) has a unique strong solution v with

$$\begin{aligned}&\overline{\rho }^{\frac{1}{2}}\,v \in C([0,T_1],L^2),\quad \,v\in C([0,T_1],H^1),\\&\overline{\rho }\partial _t v\in L^2(0,T_1; (H^1)^*),\quad \, \overline{\rho }^{\frac{1}{2}}\partial _t v\in L^2([0,T_1],L^2),\quad \, \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle v\in L^2([0,T_1],L^2). \end{aligned}$$

Moreover, the solution satisfies the following estimate

$$\begin{aligned}&\sup _{t\in [0,T_1]}(\Vert \overline{\rho }^{\frac{1}{2}}v\Vert _{L^2}^2+\Vert \nabla v\Vert _{L^2}^2)+\Vert v\Vert _{L^2_{T_1}(H^1)}^2+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v\Vert _{L^2_{T_1}(L^2)}^2\\&\quad +\Vert \overline{\rho }^{-\frac{1}{2}+\kappa }\triangle v\Vert _{L^2_{T_1}(L^2)}^2+\Vert \overline{\rho }\partial _t v\Vert _{L^2_{T_1}(H^1)^*}^2 \le C_0\Vert \overline{\rho }^{\frac{1}{2}}v_0\Vert _{L^2}^2+C_0\Vert \nabla v_0\Vert _{L^2}^2+C_0(1+T_1). \end{aligned}$$

Proof

We split the proof of the lemma into four steps.

Step 1: Galerkin approximation. We first use Galerkin method to construct approximate solutions of the system (4.1). Let \(\{w_k\}_{k=1}^\infty \) are orthonormal basis of \(H^1(\Omega )\) which satisfy boundary condition \(\mathrm {S}_{\widetilde{\mathcal {A}}}(w_k)~ \widetilde{\mathcal {N}}|_{x_3=1}=0\) and \(w_k|_{x_3=0}=0\) and set approximate solution with the form

$$\begin{aligned} v^m(t,x):=\sum _{k=1}^m d^m_k(t)w_k(x), \quad d^m_k(t)\quad \text {will be determined later on}, \end{aligned}$$

which solves the linear system

$$\begin{aligned} \left\{ \begin{aligned}&\overline{\rho }\partial _t v^m+\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}((\widetilde{J})^{-2}\overline{\rho }^2)-\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot \mathbb {S}_{\widetilde{\mathcal {A}}}(v^m)=0.\quad \text{ in }\quad \Omega ,\\&\mathrm {S}_{\widetilde{\mathcal {A}}}(v^m)~ \widetilde{\mathcal {N}}=0,\quad \text{ on }\quad \Gamma ,\\&v^m|_{x_3=0}=0,\\&v^m|_{t=0}=v_0^m=\sum _{k=1}^m d^m_k(0)w_k(x) \quad \text{ in }\quad \Omega \end{aligned} \right. \end{aligned}$$
(4.2)

in the sense of the distribution, where \(d^m_k(0)=\int _{\Omega } v_0 w_k\) for \(k=1,\ldots , m\).

Taking the test function \(\phi =w_\ell \), \(\ell =1,\ldots , m\), from the weak formula of the system (4.2), we obtain the following ordinary differential equations

$$\begin{aligned} {\left\{ \begin{array}{ll} \sum _{k=1}^m\int _{\Omega }\overline{\rho }w_k w_\ell dx~ d^m_k(t)'+\sum _{k=1}^m\int _\Omega \mathbb {S}_{\widetilde{\mathcal {A}}}w_k:\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}w_\ell dx~d^m_k(t)\\ \quad =\int _\Omega \overline{\rho }^2\widetilde{J}^{-2}\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot w_\ell dx,\\ d^m_k(0)=\int _{\Omega } v_0 w_k. \end{array}\right. } \end{aligned}$$
(4.3)

Notice that the matrix \(\Big (\int _{\Omega }\overline{\rho }w_k w_\ell dx \Big )_{m\times m}\) is invertible for any \(m\ge 1\), and the coefficient \(\int _\Omega \mathbb {S}_{\widetilde{\mathcal {A}}}w_k:\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}w_\ell dx\) (in front of \(d^m_k(t)\)) is continuous in terms of \(t\in [0,T]\) because of \(\widetilde{v}\in Y_T\), we know that (4.3) is a non-generate linear ODE system with continuous coefficients. Due to the classical theory of ODE, we find solutions \(d^m_k(t)\in C^1([0,T]), k=1,\ldots , m\), which means approximate solutions \(v^m(t,x)\) exist and belong to the space \(C^1([0,T],H^1(\Omega )).\)

Step 2: Uniform estimates for \(v^m\). Multiplying \(d^m_\ell (t)\) on the both sides of (4.3) and taking the summation in terms of \(\ell =1,\ldots , m\), one has

$$\begin{aligned} \int _{\Omega }\overline{\rho }\partial _t v^m \cdot \,v^m+\int _{\Omega }\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m):\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}v^m dx=\int _\Omega \overline{\rho }^2\widetilde{J}^{-2}\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot v^m dx. \end{aligned}$$

Then Lemmas 2.4 and 3.3 give that

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\Vert \overline{\rho }^{\frac{1}{2}}v^m\Vert _{L^2}^2+c_0\Vert v^m\Vert _{H^1}^2\le&\int _\Omega \overline{\rho }^2\widetilde{J}^{-2}\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot v^m dx+C_0\Vert \overline{\rho }^{\frac{1}{2}}v^m\Vert _{L^2}^2\nonumber \\ \le&C_0\Vert \nabla v^m\Vert _{L^2}+C_0\Vert \overline{\rho }^{\frac{1}{2}}v^m\Vert _{L^2}^2, \end{aligned}$$
(4.4)

for t small enough.

By Gronwall’s inequality, we know there exists \(T_1>0\) independent of m such that

$$\begin{aligned} \sup _{t\in [0,T_1]}\Vert \overline{\rho }^{\frac{1}{2}}v^m\Vert _{L^2}^2+\int _0^{T_1}\Vert v^m\Vert _{H^1}^2ds\le C_0\Vert \overline{\rho }^{\frac{1}{2}}v^m_0\Vert _{L^2}^2+C_0T_1. \end{aligned}$$
(4.5)

For any test function \(\phi \in C([0, T], H^1)\) with \(\phi |_{x_3=0}=0\) and \(\Vert \phi \Vert _{L^2_TH^1}\le 1,\) owing to the weak formula of the system (4.2), we deduce from (4.5) that

$$\begin{aligned}&|\int _0^{T_1}\langle \overline{\rho }\partial _t v^m, \,\phi \,\rangle \,ds| =|-\int _0^{T_1}\int _{\Omega }\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m):\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\phi \,dxds+\int _0^{T_1}\int _\Omega \overline{\rho }^2\widetilde{J}^{-2}\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot \phi \,dxds|\\&\quad \le \,(C_0+C_0\Vert \nabla v^m\Vert _{L^2_{T_1}L^2})\Vert \phi \Vert _{L^2_TH^1}\le (C_0(1+T_1^\frac{1}{2})+C_0\Vert \overline{\rho }^{\frac{1}{2}}v^m_0\Vert _{L^2})\Vert \phi \Vert _{L^2_TH^1}, \end{aligned}$$

which follows from the dual argument that

$$\begin{aligned} \Vert \overline{\rho }\partial _t v^m\Vert _{L^2_{T_1}(H^1)^*}\le C_0(1+T_1^\frac{1}{2})+C_0\Vert \overline{\rho }^{\frac{1}{2}}v^m_0\Vert _{L^2}. \end{aligned}$$
(4.6)

Multiplying \(d^m_\ell (t)'\) on the both sides of (4.3) and taking the summation in terms of \(\ell =1,\ldots , m\), we have

$$\begin{aligned} \int _{\Omega }\overline{\rho }|\partial _t v^m|^2+\int _{\Omega }\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m):\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\partial _tv^m dx=\int _\Omega \overline{\rho }^2\widetilde{J}^{-2}\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot \partial _tv^m dx. \end{aligned}$$

Similar estimate in Proposition 3.7 implies that

$$\begin{aligned}&\frac{1}{2}\frac{d}{dt}\int _{\Omega }\widetilde{J}\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m): \nabla _{\widetilde{\mathcal {A}}}v^mdx+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v^m\Vert _{L^2}^2\\&\quad \le \Big | \frac{1}{2}\int _{\Omega }\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m): \nabla _{\widetilde{\mathcal {A}}}v^m\partial _t \widetilde{J}dx\Big | +\Big | \int _{\Omega }\widetilde{J}\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m):\nabla _{\partial _t{\widetilde{\mathcal {A}}}} v^mdx \Big |\\&\qquad +\Big |\int _\Omega \overline{\rho }^2\widetilde{J}^{-2}\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot \partial _tv^m dx\Big |. \end{aligned}$$

Since \(\widetilde{v}\in Y_T\), we infer that

$$\begin{aligned} \Big | \frac{1}{2}\int _{\Omega }\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m): \nabla _{\widetilde{\mathcal {A}}}v^m\partial _t \widetilde{J}dx\Big | +\Big | \int _{\Omega }\widetilde{J}\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m):\nabla _{\partial _t{\widetilde{\mathcal {A}}}} v^m\,dx \Big |\le C\Vert \nabla v^m\Vert _{L^2}^2\mathfrak {D}(\widetilde{v}) \end{aligned}$$

and

$$\begin{aligned} \Big |\int _\Omega \overline{\rho }^2\widetilde{J}^{-2}\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\cdot \partial _tv^m dx\Big |\le C_0\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v^m\Vert _{L^2}. \end{aligned}$$

As a result, we get

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\int _{\Omega }\widetilde{J}\mathbb {S}_{\widetilde{\mathcal {A}}}(v^m): \nabla _{\widetilde{\mathcal {A}}}v^mdx+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v^m\Vert _{L^2}^2\le&C\Vert \nabla v^m\Vert _{L^2}^2\mathfrak {D}(\widetilde{v})+C_0\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v^m\Vert _{L^2}. \end{aligned}$$

Integrating time from 0 to \(T_1\) and using \(\widetilde{v}\in Y_T\) and Lemma 3.9, we obtain

$$\begin{aligned}&\sup _{t\in [0,T_1]}\Vert \nabla v^m\Vert _{L^2}^2+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v^m\Vert _{L^2_{T_1}L^2}^2\\&\quad \le C_0\Vert \nabla v_0^m\Vert _{L^2}^2+C_0T_1^\frac{1}{2}\sup _{t\in [0,T_1]}\Vert \nabla v^m\Vert _{L^2}^2(\int _0^{T_1}\mathfrak {D}(\widetilde{v})^2ds)^{\frac{1}{2}}+C_0T_1. \end{aligned}$$

Taking \(T_1\) small enough such that the second term on the right hand side absorbed by the left hand side, we obtain

$$\begin{aligned} \sup _{t\in [0,T_1]}\Vert \nabla v^m\Vert _{L^2}^2+c_0\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v^m\Vert _{L^2_{T_1}L^2}^2\le 2C_0\Vert \nabla v_0^m\Vert _{L^2}^2+C_0T_1. \end{aligned}$$
(4.7)

Combining estimate (4.5), (4.6) and (4.7) together, there holds that

$$\begin{aligned}&\sup _{t\in [0,T_1]}(\Vert \overline{\rho }^{\frac{1}{2}}v^m\Vert _{L^2}^2+\Vert \nabla v^m\Vert _{L^2}^2)+\Vert v^m\Vert _{L_{T_1}^2 H^1}^2+\Vert \overline{\rho }^{\frac{1}{2}}\partial _t v^m\Vert _{L^2_{T_1}L^2}^2+\Vert \overline{\rho }\partial _t v^m\Vert _{L^2_{T_1}(H^1)^*}^2\nonumber \\&\quad \le 2C_0\Vert \overline{\rho }^{\frac{1}{2}}v^m_0\Vert _{L^2}^2+2C_0\Vert \nabla v_0^m\Vert _{L^2}^2+C_0(1+T_1). \end{aligned}$$
(4.8)

Step 3: Passing to the limit. Since

$$\begin{aligned} \sup _{t\in [0,T_1]}(\Vert \overline{\rho }^{\frac{1}{2}}v^m\Vert _{L^2}^2+\Vert \nabla v^m\Vert _{L^2}^2)+\Vert v^m\Vert _{L^2_{T_1} H^1}^2+\Vert \rho ^{\frac{1}{2}}\partial _t v^m\Vert _{L^2_{T_1}L^2}^2+\Vert \overline{\rho }\partial _t v^m\Vert _{L^2_{T_1}(H^1)^*}^2 \end{aligned}$$

is uniformly bounded, up to the extraction of a subsequence, we know as \(m\rightarrow \infty \)

$$\begin{aligned} \left\{ \begin{aligned}&\overline{\rho }^{\frac{1}{2}}v^m\rightharpoonup ^* \overline{\rho }^{\frac{1}{2}}v\quad \text{ in } L^\infty _{T_1} L^2,\\&\nabla v^m\rightharpoonup ^*\nabla v\quad \text{ in } L^\infty _{T_1} L^2,\\&\overline{\rho }\partial _t v^m \rightharpoonup \overline{\rho }\partial _t v \quad \text{ in } L^2_{T_1} (H^1)^*,\\&v^m \rightharpoonup v \quad \text{ in } L^2_{T_1} H^1. \end{aligned} \right. \end{aligned}$$
(4.9)

By lower semicontinuity and energy estimate (4.8), we use the fact \(\Vert v^m(0)-v_0\Vert _{L^2(\Omega )}\rightarrow 0\) as \(m\rightarrow \infty \) to infer that

$$\begin{aligned}&\sup _{t\in [0,T_1]}(\Vert \overline{\rho }^{\frac{1}{2}}v\Vert _{L^2}^2+\Vert \nabla v\Vert _{L^2}^2)+\Vert v\Vert _{L^2_{T_1} H^1}^2+\Vert \rho ^{\frac{1}{2}}\partial _t v\Vert _{L^2_{T_1}L^2}+\Vert \overline{\rho }\partial _t v\Vert _{L^2_{T_1}(H^1)^*}^2\nonumber \\&\quad \le 4C_0\Vert \overline{\rho }^{\frac{1}{2}}v_0\Vert _{L^2}^2+4C_0\Vert \nabla v_0\Vert _{L^2}^2+C_0(1+T_1), \end{aligned}$$
(4.10)

and v is a weak solution to the linear \(\mathcal {A}\)-equations (4.1). Moreover, according to (4.10), we may obtain from Aubin–Lions’s lemma [38] that \(v \in C([0, T_1], X^0_{\frac{1}{2}} \cap \, H^1)\).

Step 4: The strong solution. Now, we prove the above weak solution v is a strong one. In fact, for a.e \(t\in [0,T],\) v(t) is a weak solution to the elliptic system in the sense of

$$\begin{aligned} \int _\Omega \mathbb {S}_{\widetilde{\mathcal {A}}}(v):\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}\phi \,dx=\int _\Omega \Big (\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}(\overline{\rho }^2\widetilde{J}^{-2})-\overline{\rho }\partial _t v\Big ) \phi dx \end{aligned}$$
(4.11)

for \(\phi \in H^1. \) Since \(\overline{\rho }^{-\frac{1}{2}}\Big (\nabla _{\widetilde{J}\widetilde{\mathcal {A}}}(\overline{\rho }^2\widetilde{J}^{-2})-\overline{\rho }\partial _t v\Big )\in L^2\) for a.e \(t\in [0,T],\) by elliptic regularity theory, we know this system admires a strong solution v solving (4.1) with \(\rho ^{-\frac{1}{2}+\kappa }\triangle v\in L^2([0,T],L^2)\). The uniqueness comes from energy estimates with zero initial data. \(\square \)

4.2 High regularity of v

In this subsection, we prove when \(\widetilde{v}\in Y_T,\) so does \(v:=\Theta (\widetilde{v}).\) It is mainly based on the priori estimates in Sect. 3.

Lemma 4.2

Assume that v is a strong solution obtained in Lemma 4.1 and \(\widetilde{v}\in Y_{T}\) with initial data \(v_0\in Y_0\), then we have \(v\in Y_{T_1}\) and satisfies

$$\begin{aligned} \Vert v\Vert _{Y_{T_1}}\le CT_1+C_0\Vert v_0\Vert _{Y_0}, \end{aligned}$$

where the constant C depends on \(\Vert \widetilde{v}\Vert _{Y_T}\).

Proof

We take \(\widetilde{\mathcal {A}},\widetilde{J}\) instead of \(\mathcal {A}, J\) respectively in those estimates in Propositions 3.7 and 3.8. System (4.1) is a linear system due to \(\widetilde{\mathcal {A}},\widetilde{J}\) are regarded as known quantities, so for small \(T_1>0\), it is easy to arrive at the following estimate:

$$\begin{aligned} \Vert v^m\Vert _{Y_{T_1}}\le CT_1+C_0\Vert v^m_0\Vert _{Y_0}. \end{aligned}$$

Passing to the limit, we get the desired results. \(\square \)

Remark 4.3

By Lemma 4.2, we know that \(\Theta : Y_{T_1}\rightarrow Y_{T_1} \) is well-defined.

4.3 Contraction

By Lemmas 4.1 and 4.2, we know that if \(\widetilde{v} \in Y_T\) with \(T>0\) sufficiently small , we can find a unique strong solution of equation (4.1) with regular \(v=\Theta (\widetilde{v})\in Y_T.\) In order to construct the solution to (1.18), we need to construct approximate solutions. The approximate solutions \(\{\xi ^{(n)},\,v^{(n)}\}_{n=1}^{\infty }\) we defined are iterated as follows:

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\xi ^{(n)}=v^{(n)}\quad \text{ in }\quad \Omega ,\\&\overline{\rho }\partial _t v^{(n)}+\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}((J^{(n-1)})^{-2}\overline{\rho }^2)-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\cdot \mathbb {S}_{\mathcal {A}^{(n-1)}}(v^{(n)})=0\quad \text{ in }\quad \Omega ,\\&\mathrm {S}_{\mathcal {A}^{(n-1)}}(v^{(n)})~ N^{(n-1)}=0,\quad \text{ on }\quad \Gamma ,\\&v^{(n)}|_{x_3=0}=0,\\&(\xi ^{(n)},\,v^{(n)}|_{t=0}=(\xi _0,\,v_0) \quad \text{ in }\quad \Omega . \end{aligned} \right. \end{aligned}$$
(4.12)

with \(\{\xi ^{(1)},\,v^{(1)}\}\) be the solution of linear equation

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\xi ^{(1)}=v^{(1)}\quad \text{ in }\quad \Omega ,\\&\overline{\rho }\partial _t v^{(1)}+\nabla _{J_0\mathcal {A}_0} (\overline{\rho }^2J_0^{-1})-\nabla _{J_0\mathcal {A}_0}\cdot \mathbb {S}_{\mathcal {A}_0}(v^{(1)})=0\quad \text{ in }\quad \Omega ,\\&\mathrm {S}_{\mathcal {A}_0} (v^{(1)})~ \mathcal {N}_0=0,\quad \text{ on }\quad \Gamma ,\\&v^{(1)}|_{x_3=0}=0,\\&(\xi ^{(1)},\,v^{(1)}|_{t=0}=(\xi _0,\,v_0) \quad \text{ in }\quad \Omega , \end{aligned} \right. \end{aligned}$$
(4.13)

where \(\mathcal {A}_0,J_0\) are given by \(\eta _0(x)= x+\xi _0(x)\) and \(\mathcal {N}_0=\partial _1\eta _0\times \partial _2\eta _0\) on \(\{x_3=1\}.\) Since (4.12) is a decouple linear system in terms of \(\xi ^{(n)}\) and \(v^{(n)}\), we need only to solve first \(v^{(n)}\) then \(\xi ^{(n)}\) according to the first equation in (4.12). Notice that (4.13) is linear, the assumption on initial data \( \Vert v_0\Vert _{Y_0}^2:=\Vert v_0\Vert _{X^{12}_{\frac{1}{2}}}^2+\Vert \nabla v_0\Vert _{L^2}^2 \le \frac{M}{2C_0}\) guarantees that \(v^{(1)}\in Y_T\) with bound \(\Vert v^{(1)}\Vert _{Y_T}^2\le M.\) By Lemma 4.2, we obtain \(\{v^{(n)}\}_{n=1}^{\infty }\subset Y_T\) for any \(n\ge 1.\)

Next, our goal in this subsection is to prove sequence \(\{v^{(n)}\}_{n=1}^{\infty }\) is contracted under norm \(\widetilde{Y}_T\).

First of all, we deduce \(\sigma (v^{(n)})\triangleq v^{(n+1)}-v^{(n)}\) satisfies the following equation

$$\begin{aligned} \left\{ \begin{aligned} \overline{\rho }\partial _t&\sigma (v^{(n)})-\Big (\nabla _{J^{(n)}\mathcal {A}^{(n)}}\cdot \mathbb {S}_{\mathcal {A}^{(n)}}v^{(n+1)}-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\cdot \mathbb {S}_{\mathcal {A}^{(n-1)}}v^{(n)}\Big )\\&+\Big ( \nabla _{J^{(n)}\mathcal {A}^{(n)}}\big ((J^{(n)})^{-2}\overline{\rho }^2\big )-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\big ((J^{(n-1)})^{-2}\overline{\rho }^2\big )\Big ) =0\quad \text{ in }\quad \Omega ,\\&\mathrm {S}_{\mathcal {A}^{(n)}}v^{(n+1)}~ \mathcal {N}^{(n)}-\mathrm {S}_{\mathcal {A}^{(n-1)}}v^{(n)}~ \mathcal {N}^{(n-1)}=0,\quad \text{ on }\quad \Gamma ,\\&\sigma (v^{(n)})|_{x_3=0}=0,\\&\sigma (v^{(n)})|_{t=0}=v^{(n+1)}_0-v^{(n)}_0 =0\quad \text{ in }\quad \Omega . \end{aligned} \right. \end{aligned}$$
(4.14)

Lemma 4.4

Assume that \(\{v^{(n)}\}_{n=1}^{\infty }\) be the solutions of Eq. (4.12) with bound \(\Vert v^{(n)}\Vert _{Y_T}^2\le M\) for each \(n\ge 1.\) It holds that

$$\begin{aligned} \frac{d}{dt}\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)})\Vert _{L^2}^2+\Vert \sigma (v^{(n)})\Vert _{H^1}^2 \le Ct\Vert \sigma (v^{(n-1)})\Vert _{L_t^2L^2}^2(1+\mathfrak { D}(\sigma (v^{(n)}))^2). \end{aligned}$$

Moreover, taking T small enough, the sequence \(v^{(n)}\) is a Cauchy sequence in the space \(\widetilde{Y}_T\).

Proof

Taking \(L^2\) inner product between (4.14) and \(\sigma (v^{(n)})\), we obtain

$$\begin{aligned}&\frac{1}{2}\frac{d}{dt}\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)})\Vert _{L^2}^2-\int _{\Omega }\Big (\nabla _{J^{(n)}\mathcal {A}^{(n)}}\cdot \mathbb {S}_{\mathcal {A}^{(n)}}v^{(n+1)}-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\cdot \mathbb {S}_{\mathcal {A}^{(n-1)}}v^{(n)}\Big )~ \sigma (v^{(n)})dx\\&\quad = -\int _{\Omega }\Big ( \nabla _{J^{(n)}\mathcal {A}^{(n)}}\big ((J^{(n)})^{-2}\overline{\rho }^2\big )-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\big ((J^{(n-1)})^{-2}\overline{\rho }^2\big )\Big ) ~ \sigma (v^{(n)})dx. \end{aligned}$$

Estimate of dissipation term. Since

$$\begin{aligned} e_3J^{(n)}(\mathcal {A}^{(n)})^3_i=\mathcal {N}^{(n)},\quad \,e_3J^{(n-1)}(\mathcal {A}^{(n-1)})^3_i=\mathcal {N}^{(n-1)} \end{aligned}$$

and

$$\begin{aligned} \mathrm {S}_{\mathcal {A}^{(n)}}(v^{(n+1)})~ \mathcal {N}^{(n)}-\mathrm {S}_{\mathcal {A}^{(n-1)}}(v^{(n)})~ \mathcal {N}^{(n-1)}=0 \quad \text{ on }\quad \Gamma , \end{aligned}$$

we get by using integration by parts that

$$\begin{aligned}&-\int _{\Omega }\Big (\nabla _{J^{(n)}\mathcal {A}^{(n)}}\cdot \mathbb {S}_{\mathcal {A}^{(n)}}v^{(n+1)}-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\cdot \mathbb {S}_{\mathcal {A}^{(n-1)}}v^{(n)}\Big )~ \sigma (v^{(n)})dx\\&\quad =\int _{\Omega }\Big (J^{(n)}(\mathcal {A}^{(n)})^k_i(\mathbb {S}_{\mathcal {A}^{(n)}}v^{(n+1)})^i_l-J^{(n-1)}(\mathcal {A}^{(n-1)})^k_i(\mathbb {S}_{\mathcal {A}^{(n-1)}}v^{(n)})^i_l\Big )~ \partial _k\sigma (v^{(n)}_l)dx.\\&\quad =\int _{\Omega } J^{(n)}\mathcal {A}^{(n)}\mathbb {S}_{\mathcal {A}^{(n)}}\sigma (v^{(n)}) \cdot \partial _k\sigma (v^{(n)}_l)dx\\&\qquad +\int _{\Omega } (J^{(n)}\mathcal {A}^{(n)}-J^{(n-1)}\mathcal {A}^{(n-1)}) \mathbb {S}_{\mathcal {A}^{(n)}}(v^{(n)})\cdot \partial _k\sigma (v^{(n)}_l)dx\\&\qquad +\int _{\Omega } J^{(n-1)}\mathcal {A}^{(n-1)}\cdot \mathbb {S}_{\big (\mathcal {A}^{(n)}-\mathcal {A}^{(n-1)}\big )}v^{(n)}\cdot \partial _k\sigma (v^{(n)}_l)dx \end{aligned}$$

Under the assumption \(\Vert v^{(n)}\Vert _{Y_T}^2\le M\), we have

$$\begin{aligned}&-\int _{\Omega }\Big (\nabla _{J^{(n)}\mathcal {A}^{(n)}}\cdot \mathbb {S}_{\mathcal {A}^{(n)}}v^{(n+1)}-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\cdot \mathbb {S}_{\mathcal {A}^{(n-1)}}v^{(n)}\Big )~ \sigma (v^{(n)})dx\\&\quad \ge \,c_0\Vert \sigma (v^{(n)})\Vert _{H^1}^2-C_0\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)})\Vert _{L^2}^2\\&\qquad -\Big |\int _{\Omega }(J^{(n)}\mathcal {A}^{(n)}-J^{(n-1)}\mathcal {A}^{(n-1)}) \mathbb {S}_{\mathcal {A}^{(n)}}(v^{(n)}): \nabla \sigma (v^{(n)})dx\Big |\\&\qquad -\Big |\int _{\Omega }J^{(n-1)}\mathcal {A}^{(n-1)}\cdot \mathbb {S}_{\big (\mathcal {A}^{(n)}-\mathcal {A}^{(n-1)}\big )}v^{(n)}:\nabla \sigma (v^{(n)})dx\Big |\\&\quad \triangleq \,c_0\Vert \sigma (v^{(n)})\Vert _{H^1}^2-C_0\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)})\Vert _{L^2}^2-I_1-I_2, \end{aligned}$$

where we use \(|J^{(n)}|\ge \sigma _0\) and Lemma 3.5 for \(\mathcal {A}^{(n)}.\)

For \(I_1\), owing to

$$\begin{aligned}&J^{(n)}\mathcal {A}^{(n)}-J^{(n-1)}\mathcal {A}^{(n-1)}=\big (\nabla (\eta ^{(n)}-\eta ^{(n-1)})\big )^*\\&\quad =\left( \int _0^t \nabla \sigma (v^{(n-1)})ds \right) ^*\sim \left( \int _0^t \nabla \sigma (v^{(n-1)})ds \right) ^2, \end{aligned}$$

then

$$\begin{aligned}&\Vert J^{(n)}\mathcal {A}^{(n)}-J^{(n-1)}\mathcal {A}^{(n-1)}\Vert _{L^2}\\&\quad \le C t \Vert \nabla \sigma (v^{(n-1)})\Vert _{L^2_t L^2}\Vert \mathfrak { D}(\sigma (v^{(n-1)}))\Vert _{L^2_t}\le C t \Vert \nabla \sigma (v^{(n-1)})\Vert _{L^2_t L^2}. \end{aligned}$$

Applying Holder inequality and Lemma 3.2 to \(\mathcal {A}^{(n)}\), one has

$$\begin{aligned} I_1&\le \Vert J^{(n)}\mathcal {A}^{(n)}-J^{(n-1)}\mathcal {A}^{(n-1)}\Vert _{L^2}\Vert \mathcal {A}^{(n)}\Vert _{L^\infty }\Vert \nabla v^{(n)} \Vert _{L^\infty }\Vert \nabla \sigma (v^{(n)})\Vert _{L^2}\\&\le Ct\Vert \nabla \sigma (v^{(n-1)})\Vert _{L_t^2L^2}\mathfrak { D}(v^{(n)})\Vert \nabla \sigma (v^{(n)})\Vert _{L^2}. \end{aligned}$$

Similarly, we have

$$\begin{aligned} I_2\le Ct\Vert \nabla \sigma (v^{(n-1)})\Vert _{L_t^2L^2}\mathfrak { D}(v^{(n)})\Vert \nabla \sigma (v^{(n)})\Vert _{L^2}. \end{aligned}$$

Combining all above estimates, we obtain

$$\begin{aligned}&-\int _{\Omega }\Big (\nabla _{J^{(n)}\mathcal {A}^{(n)}}\cdot \mathbb {S}_{\mathcal {A}^{(n)}}v^{(n+1)}-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\cdot \mathbb {S}_{\mathcal {A}^{(n-1)}}v^{(n)}\Big )~ \sigma (v^{(n)})dx\\&\quad \ge \frac{3}{4} c_0\Vert \sigma (v^{(n)})\Vert _{H^1}^2-C_0\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)})\Vert _{L^2}^2 -Ct^2\Vert \nabla \sigma (v^{(n-1)})\Vert _{L_t^2L^2}^2\mathfrak { D}(v^{(n)})^2. \end{aligned}$$

Estimate of pressure term. Integrating by parts and using \(\overline{\rho }|_{x_3=1}=0, ~\sigma (v^{(n)})|_{x_3=0}=0\), we prove that

$$\begin{aligned}&-\int _{\Omega }\Big ( \nabla _{J^{(n)}\mathcal {A}^{(n)}}\big ((J^{(n)})^{-2}\overline{\rho }^2\big )-\nabla _{J^{(n-1)}\mathcal {A}^{(n-1)}}\big ((J^{(n-1)})^{-2}\overline{\rho }^2\big )\Big ) ~ \sigma (v^{(n)})dx\\&\quad =\int _{\Omega }\Big ( \mathcal {A}^{(n)}\big ((J^{(n)})^{-1}\overline{\rho }^2\big )-\mathcal {A}^{(n-1)}\big ((J^{(n-1)})^{-1}\overline{\rho }^2\big )\Big ) : \nabla \sigma (v^{(n)})dx\\&\quad =\int _{\Omega }\big (\mathcal {A}^{(n)}-\mathcal {A}^{(n-1)}\big )(J^{(n)})^{-1}\overline{\rho }^2 : \nabla \sigma (v^{(n)})dx\\&\qquad +\int _{\Omega }\mathcal {A}^{(n-1)}\big ((J^{(n)})^{-1}-(J^{(n-1)})^{-1}\big )\overline{\rho }^2 : \nabla \sigma (v^{(n)})dx\\&\quad \le Ct^{\frac{1}{2}}\Vert \nabla \sigma (v^{(n-1)})\Vert _{L^2_tL^2}\Vert \nabla \sigma (v^{(n)})\Vert _{L^2}. \end{aligned}$$

Collecting all above estimates together, we finally obtain

$$\begin{aligned}&\frac{d}{dt}\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)})\Vert _{L^2}^2+\frac{c_0}{2}\Vert \nabla \sigma (v^{(n)})\Vert _{L^2}^2\nonumber \\&\quad \le Ct\Vert \nabla \sigma (v^{(n-1)})\Vert _{L_t^2L^2}^2(1+\mathfrak { D}(v^{(n)})^2)+C_0\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)})\Vert _{L^2}^2. \end{aligned}$$
(4.15)

Integrating (4.15) in \(t \in [0, T]\) and taking T small enough, we have

$$\begin{aligned}&\sup _{t\in [0,T]}\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)}(t))\Vert _{L^2}^2+\frac{c_0}{2}\int _0^T\Vert \sigma (v^{(n)}(t))\Vert _{H^1}^2dt\nonumber \\&\quad \le \Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)}(0))\Vert _{L^2}^2+CT\Vert \nabla \sigma (v^{(n-1)})\Vert _{L_T^2L^2}^2(T+\int _0^T\mathfrak { D}(v^{(n)})^2dt), \end{aligned}$$
(4.16)

and then

$$\begin{aligned}&\sup _{t\in [0,T]}\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)}(t))\Vert _{L^2}^2+\frac{c_0}{2}\int _0^T\Vert \sigma (v^{(n)}(t))\Vert _{H^1}^2dt\\&\quad \le CT(T+M)\Vert \nabla \sigma (v^{(n-1)}\Vert _{L_T^2L^2}^2\le CT\Vert \nabla \sigma (v^{(n-1)})\Vert _{L_T^2L^2}^2. \end{aligned}$$

By now, we get that when T takes small enough, then we get

$$\begin{aligned}&\sup _{t\in [0,T]}\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n)}(t))\Vert _{L^2}^2+\Vert \sigma (v^{(n)}(t))\Vert _{L^2_TH^1}^2\\&\quad \le \frac{1}{2}(\sup _{t\in [0,T]}\Vert \overline{\rho }^{\frac{1}{2}}\sigma (v^{(n-1)}(t))\Vert _{L^2}^2+\Vert \sigma (v^{(n-1)}(t))\Vert _{L^2_TH^1}^2), \end{aligned}$$

which completes this Lemma. \(\square \)

4.4 Proof of Theorem 1.2.

From Lemma 4.4, we know \(\{v^{(n)}\}_{n=1}^{\infty }\) is Cauchy sequence in the space \(\widetilde{Y}_T.\) So as \(n\rightarrow \infty ,\)

$$\begin{aligned} \left\{ \begin{aligned}&\overline{\rho }^{\frac{1}{2}}v^{(n)}\rightarrow \overline{\rho }^{\frac{1}{2}}v\qquad \text{ in }\quad C([0,T],L^2),\\&v^{(n)}\rightarrow v\qquad \text{ in }\quad L^2([0,T],H^1). \end{aligned} \right. \end{aligned}$$
(4.17)

Due to Lemma 4.2 that \(\Vert v^{(n)}\Vert _{Y_T}^2\le M\) uniformly in \(n\ge 1,\) sequence \(\{v^{(n)}\}_{n=1}^{\infty }\) have weakly convergent subsequence. Along with strong convergence (4.17), we infer that as \(n\rightarrow 0\)

$$\begin{aligned} \left\{ \begin{aligned}&v^{(n)}\rightharpoonup ^* v\qquad \text{ in }\quad L^\infty ([0,T],X^{12}_{\frac{1}{2}}),\\&v^{(n)}\rightharpoonup v,\qquad \,\nabla v^{(n)}\rightharpoonup \nabla v\qquad \text{ in }\quad L^2([0,T],X^{12}),\\&\nabla v^{(n)}\rightharpoonup ^* \nabla v\qquad \text{ in }\quad L^\infty ([0,T],L^2),\\&\overline{\rho }^{\frac{1}{2}}\partial _t v^{(n)}\rightharpoonup \overline{\rho }^{\frac{1}{2}}\partial _tv\qquad \text{ in }\quad L^2([0,T],L^2). \end{aligned} \right. \end{aligned}$$

So the function v satisfies equation (1.18) in weak sense. On the other hand, lower semicontinuity gives bound \(\Vert v\Vert _{Y_T}^2\le 2M\), and then (1.25) holds. As a result, thanks to Aubin–Lions’s lemma [38], we get that \((v,\,\eta )\in \,C([0, T]; X^{12}_{\frac{1}{2}}\cap H^1(\Omega )) \times C([0, T]; \mathcal {F}_{\kappa }(\Omega ))\) by using a standard procedure (cf. the proof of Theorem 3.5 in [33]), which is a strong solution to (1.18). The uniqueness comes from \(L^2\) energy estimates with zero initial data. More precise, let \((\xi _1,\,v_1)\) and \((\xi _2,\,v_2)\) are solutions to (1.18) with same initial data. The same process in Lemma 4.4 deduce that

$$\begin{aligned} \Vert v_1-v_2\Vert _{\widetilde{Y}_T}^2\le \frac{1}{2} \Vert v_1-v_2\Vert _{\widetilde{Y}_T}^2, \end{aligned}$$

which implies \(v_1=v_2\) and then \(\xi _1=\xi _2\) on the time interval [0, T]. Furthermore, applying (4.16) to the system (1.18), we may readily prove that the solution \((v,\,\eta )\in \,C([0, T]; X^{12}_{\frac{1}{2}}\cap H^1(\Omega )) \times C([0, T]; \mathcal {F}_{\kappa }(\Omega ))\) depends continuously on the initial data \((v_0,\,\eta _0)\in \,(X^{12}_{\frac{1}{2}}\cap H^1(\Omega )) \times \mathcal {F}_{\kappa }(\Omega )\). This finishes the proof of Theorem 1.2. \(\square \)