This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Asymptotical Behavior of Global Solutions of the Navier–Stokes–Korteweg Equations with Respect to
Capillarity Number at Infinity

Fei Jiang [email protected] Pengfei Li [email protected] Jiawei Wang [email protected] School of Mathematics and Statistics, Fuzhou University, Fuzhou, 350108, China. Key Laboratory of Operations Research and Control of Universities in Fujian, Fuzhou 350108, China. Hua Loo-Keng Center for Mathematical Sciences, Chinese Academy of Sciences, Beijing 100190, China.
Abstract

Vanishing capillarity in the Navier–Stokes–Korteweg (NSK) equations has been widely investigated, in particular, it is well-known that the NSK equations converge to the Navier–Stokes (NS) equations by vanishing capillarity number. To our best knowledge, this paper first investigates the behavior of large capillary number, denoted by κ2\kappa^{2}, for the global(-in-time) strong solutions with small initial perturbations of the three-dimensional (3D) NSK equations in a slab domain with Navier(-slip) boundary condition. Under the well-prepared initial data, we can construct a family of global strong solutions of the 3D incompressible NSK equations with respect to κ>0\kappa>0, where the solutions converge to a unique solution of 2D incompressible NS-like equations as κ\kappa goes to infinity.

keywords:
Navier–Stokes–Korteweg equations; asymptotical behavior; uniform estimates; global-in-time solutions.

1 Introduction

A classical model to describe the dynamics of an inhomogeneous incompressible fluid endowed with internal capillarity (in the diffuse interface setting) is the following general system of incompressible Navier–Stokes–Korteweg (NSK) equations:

{ρt+div(ρv)=0,ρ(vt+vv)div(μ(ρ)𝔻v)+P=divK,divv=0,\displaystyle\begin{cases}\rho_{t}+\mathrm{div}(\rho v)=0,\\ \rho(v_{t}+v\cdot\nabla v)-\mathrm{div}(\mu(\rho)\mathbb{D}v)+\nabla{P}=\mathrm{div}{K},\\ \mathrm{div}v=0,\end{cases} (1.1)

where ρ(x,t)+{\rho(x,t)}\in\mathbb{R}^{+}, v(x,t)3v(x,t)\in{\mathbb{R}^{3}} and P(x,t)P(x,t) denote the density, velocity and kinetic pressure of the fluid resp. at the spacial position x3x\in{\mathbb{R}^{3}} for time t0+:=[0,+)t\in\mathbb{R}^{+}_{0}:=[0,{+\infty}). The differential operator 𝔻\mathbb{D} is defined by 𝔻v=v+v\mathbb{D}v=\nabla v+\nabla v^{\top}, where the superscript \top represents the transposition. The shear viscosity function μ\mu, the capillarity function κ~\tilde{\kappa} are known smooth functions +\mathbb{R}^{+}\to\mathbb{R}, and satisfy μ>0\mu>0, κ~>0\tilde{\kappa}>0. The general capillary tensor is written as

K=(ρdiv(κ~(ρ)ρ)+(κ~(ρ)ρκ~(ρ))|ρ|2/2)𝕀κ~(ρ)ρρ,\displaystyle K=\left({\rho\mathrm{div}(\tilde{\kappa}(\rho)\nabla\rho)+\left({\tilde{\kappa}(\rho)-\rho\tilde{\kappa}^{\prime}(\rho)}\right){{\left|{\nabla\rho}\right|}^{2}}}/2\right)\mathbb{I}-{\tilde{\kappa}(\rho)\nabla\rho\otimes\nabla\rho},

where 𝕀\mathbb{I} is the identity matrix. To conveniently investigate the asymptotical behavior of capillarity number, we considers that μ\mu and κ~\tilde{\kappa} are positive constants [6], thus it holds that

div(μ(ρ)𝔻v)=μΔv and divK=κ2ρΔρ,\displaystyle\mathrm{div}(\mu(\rho)\mathbb{D}v)=\mu\Delta v\mbox{ and }\mathrm{div}K=\kappa^{2}\rho\nabla\Delta\rho, (1.2)

where we have defined that κ:=κ~>0\kappa:=\sqrt{\tilde{\kappa}}>0. We call the parameters μ\mu and κ2\kappa^{2} the viscosity coefficient and the capillarity number, resp..

In classical hydrodynamics, the interface between two immiscible compressible/incompressible fluids is modeled as a free boundary which evolves in time. The equations describing the motion of each fluid are supplemented by boundary conditions at the free surface involving the physical properties of the interface. For instance, in the free-boundary formulation, it is assumed that the interface has an internal surface tension. However, when the interfacial thickness is comparable to the length scale of the phenomena being examined, the free-boundary description breaks down. Diffuse-interface models provide an alternative description where the surface tension is expressed in its simplest form as divK\mathrm{div}K, i.e., the capillary tension which was introduced by Korteweg in 1901 [41]. Later, its modern form was derived by Dunn and Serrin [19].

In the physical view, it can serve as a phase transition model to describe the motion of compressible fluid with capillarity effect. Owing to its importance in mathematics and physics, there has been profuse works for the mathematical theory of the corresponding compressible NSK system, for example, see [2, 7, 24] for the global(-in-time) weak solutions with large initial data, [17, 25, 27, 40, 50] for global strong solutions with small initial data, [12, 26, 42, 43] for local(-in-time) strong/classical solutions with large initial data, [16] for stationary solutions, [20] for highly rotating limit, [15, 28] for the stability of viscous shock wave, [54] for the maximal LpL^{p}LqL^{q} regularity theory, [9, 13, 39] for the decay-in-time of global solutions, [5, 14, 21, 38, 30] for the vanishing capillarity limit, [44, 45, 29] for the nonisentropic case, [48, 55, 37] for the low Mach number limit, [3, 4, 8, 57] for the inviscid case, and so on.

The incompressible NSK system (1.1) has been also widely investigated. The local existence of a unique strong solution was obtained by Tan–Wang [58]. Burtea–Charve also established the existence result of strong solutions with small initial perturbations in Lagrangian coordinates [11], and further presented that the lifespan goes to infinity as the capillarity number goes to zero. Yang–Yao–Zhu proved that as both capillarity number and viscosity coefficient vanish, local solutions of the Cauchy problem for the incompressible NSK system converge to the one of the inhomogeneous incompressible Euler equations [60]. Recently a similar result was also established by Wang–Zhang for the fluid domain being a horizontally periodic slab with Navier-slip boundary conditions [59].

In addition, it also has been investigated that the capillarity under the large capillarity number stabilizes the motion of the incompressible fluid endowed with internal capillarity. Bresch–Desjardins–Gisclon–Sart derived that the capillarity slows down the growth rate of linear Rayleigh–Taylor (RT) instability, which occurs when a denser fluid lies on top of a lighter fluid under gravity [6]. Li–Zhang proved that the capillarity can inhibit RT instability for properly large capillarity number by the linearized motion equations [46] (Interesting readers further refer to [51, 46, 62, 61] for the existence of RT instability solutions in the incompressible NSK system with small capillarity number). Later, motivated by the result of magnetic tension inhibiting the RT instability in the two-dimensional (2D) non-resistive magnetohydrodynamic (MHD) fluid in [33], Jiang–Li–Zhang mathematically proved that the capillarity inhibits RT instability in the 2D NSK equations in the Lagrangian coordinates [34]. Recently Jiang–Zhang–Zhang further verified such inhibition phenomenon for the 3D case in Eulerian coordinates [35]. Jiang–Zhang–Zhang’s result rigorously presents that the capillarity has the stabilizing effect as well as the magnetic tension in the MHD fluid.

It is well-known that the larger the magnetic tension, the stronger the field intensity in the MHD fluid. Moreover the local solutions of the system of the idea MHD fluid with well-prepared initial data tend to a solution of a 2D Euler flow coupled with a linear transport equation as the field intensity goes to infinity, see [36, 22, 10] for relevant results. This naturally motivates us to except a similar result in the incompressible NSK system. In this paper, under the well-prepared initial data, we also construct a family of the global strong solutions of the 3D incompressible NSK equations with respect to κ\kappa, where the solutions converge to a unique solution of a 2D incompressible Navier–Stokes-like equations as κ\kappa goes to infinity, see Theorem 2.2. As Lin pointed out in [49], the stratified fluids with the internal surface tension, the viscoelastic fluids, the non-resistive MHD fluids and the diffuse-interface model (or NSK model) can also be regarded as complex fluids with elasticity. Our mathematical result in Theorem 2.2 supports that the capillarity has the role of elasticity again.

2 Main results

This section is devoted to introducing our main results in details. Let us consider a rest stat of the system (1.1). We choose the equilibrium density ρ¯\bar{\rho} in the rest state to satisfy

ρ¯=ax3+b with given constants a,b>0.\displaystyle\bar{\rho}=ax_{3}+b\mbox{ with given constants }a,\ b>0. (2.1)

Without loss of generality, we consider

a=b=1.\displaystyle a=b=1. (2.2)

Denoting the perturbations around the rest state (ρ¯,0)(\bar{\rho},0) by

ϱ=ρρ¯,u=v0,\varrho=\rho-\bar{\rho},\ \ u=v-0,

and then recalling the relation

Δρρ=(ρΔρ)ρΔρ,\displaystyle\Delta\rho\nabla\rho=\nabla(\rho\Delta\rho)-\rho\nabla\Delta\rho,

we obtain the following perturbation system from (1.1) with (1.2):

{ϱt+uϱ+u3=0,ρ(ut+uu)+β=μΔuκ2Δϱ(𝐞3+ϱ),divu=0,\begin{cases}\varrho_{t}+u\cdot\nabla\varrho+u_{3}=0,\\ \rho(u_{t}+u\cdot\nabla u)+\nabla\beta=\mu\Delta u-\kappa^{2}\Delta{\varrho}(\mathbf{e}^{3}+\nabla{\varrho}),\\ \mathrm{div}u=0,\end{cases} (2.3)

where 𝐞3\mathbf{e}^{3} represents the unit vector with the third component being 11 and we have defined that β:=Pκ2ρΔρ\beta:=P-\kappa^{2}\rho\Delta\rho. The initial condition reads as follows

(u,ϱ)|t=0=(u0,ϱ0).\displaystyle(u,\varrho)|_{t=0}=(u^{0},\varrho^{0}). (2.4)

Here and in what follows, we always use the right superscript 0 to emphasize the initial data.

We consider that the fluid domain denoted by Ω\Omega is a slab, i.e.

Ω:=2×(0,h),\displaystyle\Omega:=\mathbb{R}^{2}\times(0,h),

The 2D domain 2×{0,h}\mathbb{R}^{2}\times\{0,h\}, denoted by Ω\partial\Omega, is the boundary of Ω\Omega. We focus on the following Navier(-slip) boundary conditions on the velocity on Ω\partial\Omega:

u|Ω𝐧=0,((𝔻u|Ω)𝐧)tan=0,\displaystyle u|_{\partial\Omega}\cdot{\mathbf{n}}=0,\ ((\mathbb{D}u|_{\partial\Omega}){\mathbf{n}})_{\mathrm{tan}}=0,

where 𝐧\mathbf{n} denotes the unit outward normal to Ω\Omega and the subscript “tan” means the tangential component of a vector (for example, utan=u(u𝐧)𝐧u_{\mathrm{tan}}=u-(u\cdot\mathbf{n})\mathbf{n}) [18, 47]. The Navier slip boundary conditions describe an interaction between a viscous fluid and a solid wall. Since Ω\Omega is a slab domain, the Navier boundary conditions are equivalent to the boundary condition

(u3,3u1,3u2)|Ω=0.\displaystyle(u_{3},\partial_{3}u_{1},\partial_{3}u_{2})|_{\partial\Omega}=0. (2.5)

We mention that it is not clear to us whether our main results in Theorems 2.1 and 2.2 can be extended to the non-slip boundary condition of velocity.

However, to investigate the asymptotical behavior of solutions with respect to capillarity number at infinity and to except the vanishing of ϱ\varrho, we set

ϱ=κ1σ\displaystyle\varrho=\kappa^{-1}\sigma

and then transform (2.3) into

{σt+uσ+κu3=0,ρ(ut+uu)+β=μΔuΔσ(κ𝐞3+σ),divu=0.\displaystyle\begin{cases}\sigma_{t}+u\cdot\nabla\sigma+\kappa u_{3}=0,\\ \rho(u_{t}+u\cdot\nabla u)+\nabla\beta=\mu\Delta u-\Delta\sigma(\kappa\mathbf{e}^{3}+\nabla\sigma),\\ \mathrm{div}u=0.\end{cases} (2.6)

The corresponding initial condition reads as follows

(u,σ)|t=0=(u0,σ0).\displaystyle(u,\sigma)|_{t=0}=(u^{0},\sigma^{0}). (2.7)

In addition, to avoid the vacuum of density, we assume that

dinfxΩ{ρ0(x)}, where ρ0:=ρ¯+κ1σ0 and d+.\displaystyle d\leqslant\inf\limits_{x\in\Omega}\big{\{}{\rho}^{0}(x)\big{\}},\mbox{ where }{\rho}^{0}:=\bar{\rho}+\kappa^{-1}\sigma^{0}\mbox{ and }d\in\mathbb{R}^{+}. (2.8)

Before stating our main results, we shall introduce simplified notations, which will be used throughout this paper.

(1) Simplified basic notations: Ia:=(0,a)I_{a}:=(0,a) denotes an interval, in particular, I=+=(0,)I_{\infty}=\mathbb{R}^{+}=(0,\infty). S¯\overline{S} denotes the closure of a set SnS\subset\mathbb{R}^{n} with n1n\geqslant 1, in particular, IT¯=[0,T]\overline{I_{T}}=[0,T] and I¯=0+\overline{I_{\infty}}=\mathbb{R}^{+}_{0}. aba\lesssim b means that acba\leqslant cb for some constant c>0c>0, where c>0c>0 at most depends on the domain Ω\Omega, ρ¯\bar{\rho}, μ\mu, dd in (2.8), and may vary from line to line. Sometimes we also denote cc by cic_{i} for emphasizing that cic_{i} is fixed, where i=1i=1, 22. i:=xi\partial_{i}:=\partial_{x_{i}}, where 1i31\leqslant i\leqslant 3. Let f:=(f1,f2,f3)f:=(f_{1},f_{2},f_{3})^{\top} be a vector function defined in a 3D domain, we define that fh:=(f1,f2)f_{\mathrm{h}}:=(f_{1},f_{2})^{\top} and curlf:=(2f33f2,3f11f3,1f22f1)\mathrm{curl}{f}:=(\partial_{2}f_{3}-\partial_{3}f_{2},\partial_{3}f_{1}-\partial_{1}f_{3},\partial_{1}f_{2}-\partial_{2}f_{1})^{\top}. =(2,1,0)\nabla^{\bot}=(-\partial_{2},\partial_{1},0), h:=(1,2)\nabla_{\mathrm{h}}:=(\partial_{1},\partial_{2})^{\top}, h:=(2,1)\nabla^{\perp}_{\mathrm{h}}:=(-\partial_{2},\partial_{1})^{\top}, divh:=1+2\mathrm{div}_{\mathrm{h}}\cdot:=\partial_{1}\cdot+\partial_{2}\cdot and Δh:=12+22\Delta_{\mathrm{h}}:=\partial^{2}_{1}+\partial^{2}_{2}. For the simplicity, we denote i=1nwiX2\sqrt{\sum_{i=1}^{n}\|w_{i}\|_{X}^{2}} by (w1,,wn)X\|(w_{1},\cdots,w_{n})\|_{X}, where X\|\cdot\|_{X} represents a norm, and wiw_{i} is scalar function or vector function for 1in1\leqslant i\leqslant n.

(2) Simplified Banach spaces, norms and semi-norms:

Lp:=Lp(Ω)=W0,p(Ω),Hi:=Wi,2(Ω),Hloci:=Wloci,2(Ω),H0j:={ϕHj|ϕ|Ω=0},\displaystyle L^{p}:=L^{p}(\Omega)=W^{0,p}(\Omega),\ {H}^{i}:=W^{i,2}(\Omega),\ {H}^{i}_{\mathrm{loc}}:=W^{i,2}_{\mathrm{loc}}(\Omega),\ {H}^{j}_{0}:=\{\phi\in H^{j}~{}|~{}\phi|_{\partial\Omega}=0\},
Hρ¯3:={ϕH03|32ϕ|Ω=0,3ϕ|Ω=κρ¯|Ω},i:={uHi|divu=0,u3|Ω=0},\displaystyle{{H}}^{3}_{\bar{\rho}}:=\{\phi\in H^{3}_{0}~{}|~{}\partial^{2}_{3}\phi|_{\partial\Omega}=0,\partial_{3}\phi|_{\partial\Omega}=-\kappa\bar{\rho}^{\prime}|_{\partial\Omega}\},\ \mathcal{H}^{i}:=\{u\in H^{i}~{}|~{}\mathrm{div}u=0,\ u_{3}|_{\partial\Omega}=0\},
sk:={wk|3w1=3w2=0 on Ω},i:=Hi,i,l:=α1+α2=i1α12α2l,\displaystyle\mathcal{H}^{k}_{\mathrm{s}}:=\{w\in\mathcal{H}^{k}~{}|~{}\partial_{3}w_{1}=\partial_{3}w_{2}=0\mbox{ on }\partial\Omega\},\ \|\cdot\|_{i}:=\|\cdot\|_{H^{i}},\ \|\cdot\|_{{i},l}:=\sum_{\alpha_{1}+\alpha_{2}=i}\|\partial_{1}^{\alpha_{1}}\partial_{2}^{\alpha_{2}}\cdot\|_{l},

where 1p1\leqslant p\leqslant\infty is a real number, and ii, l0l\geqslant 0, j1j\geqslant 1, k2k\geqslant 2 are integers.

(3) Simplified spaces of functions with values in a Banach space:

𝔓T:={σC0(IT¯,Hρ¯3)|σtC0(IT¯,H1)L2(IT,H2)},\displaystyle{\mathfrak{P}}_{T}:=\{\sigma\in C^{0}(\overline{I_{T}},{{H}}^{3}_{\bar{\rho}})~{}|~{}\sigma_{t}\in C^{0}(\overline{I_{T}},H^{1})\cap L^{2}(I_{T},{H}^{2})\},
𝒱T:={uC0(IT¯,s2)L2(IT,H3)|utC0(IT¯,L2)L2(IT,H1)}.\displaystyle{\mathcal{V}}_{T}:=\{u\in C^{0}(\overline{I_{T}},{\mathcal{H}^{2}_{\mathrm{s}}})\cap L^{2}(I_{T},{H}^{3})~{}|~{}u_{t}\in C^{0}(\overline{I_{T}},L^{2})\cap L^{2}(I_{T},{H}^{1})\}.

(4) Energy/dissipation functionals (generalized):

E(t):=(σ32+u22+κu312+κ(Δhσ,h3σ)02,E0:=E(0),\displaystyle E(t):=\|(\sigma\|_{3}^{2}+\|u\|_{2}^{2}+\|\kappa u_{3}\|_{1}^{2}+\|\kappa(\Delta_{\mathrm{h}}\sigma,\nabla_{\mathrm{h}}\partial_{3}\sigma)\|_{0}^{2},\ E^{0}:=E(0),
(t):=σt12+ut02+E(t),𝒟(t):=σ1,22+(σt,u)22+ut12.\displaystyle\mathcal{E}(t):=\|\sigma_{t}\|_{1}^{2}+\|u_{t}\|_{0}^{2}+E(t),\ \mathcal{D}(t):=\|\sigma\|^{2}_{1,2}+\|(\sigma_{t},\nabla u)\|^{2}_{2}+\|u_{t}\|^{2}_{1}.

Now we state the first result, which presents that the initial-boundary value problem (2.5)–(2.7) with small initial data admits a unique global strong solution with uniform estimates with respect to κ\kappa.

Theorem 2.1.

Let μ\mu, κ\kappa, d>0d>0 be given, and ρ¯\bar{\rho} be defined by (2.1). There exist constants c1c_{1}, c2c_{2} and χ\chi, where

c1,c21 and χ:=max{κ1,c2},\displaystyle c_{1},\ c_{2}\geqslant 1\mbox{ and }\chi:=\max\{\kappa^{-1},c_{2}\}, (2.9)

such that, for any (σ0,u0)(\sigma^{0},u^{0}) satisfying (2.8),

(σ0,u0)Hρ¯3×s2 and E0(3c1χ9)3,\displaystyle(\sigma^{0},u^{0})\in H^{3}_{\bar{\rho}}\times{\mathcal{H}^{2}_{\mathrm{s}}}\mbox{ and }E^{0}\leqslant(3c_{1}\chi^{9})^{-3}, (2.10)

the initial-boundary value problem (2.5)–(2.7) has a unique global strong solution

(σ,u,β)𝔓×𝒱×C0(0+,L2).(\sigma,u,\nabla\beta)\in{\mathfrak{P}}_{\infty}\times{\mathcal{V}}_{\infty}\times C^{0}(\mathbb{R}_{0}^{+},L^{2}).

Moreover the solution satisfies the following estimates

(t)+0t𝒟(τ)𝑑τ2c1χ9E0<1,\displaystyle\mathcal{E}(t)+\int_{0}^{t}\mathcal{D}(\tau)d\tau\leqslant 2c_{1}\chi^{9}E^{0}<1, (2.11)
𝒫(t)02c1χ9(1+χ2)E0 for any t>0\displaystyle\|\nabla\mathcal{P}(t)\|_{0}^{2}\lesssim c_{1}\chi^{9}(1+\chi^{2}){E^{0}}\mbox{ for any }t>0 (2.12)

and

(κ(Δhσ,h3σ),𝒫)L2(IT,H1)2c1χ9(1+χ2+T)E0 for any T>0,\displaystyle\|(\kappa(\Delta_{\mathrm{h}}\sigma,\nabla_{\mathrm{h}}\partial_{3}\sigma),\nabla\mathcal{P})\|_{L^{2}(I_{T},H^{1})}^{2}\lesssim c_{1}\chi^{9}(1+\chi^{2}+T){E^{0}}\mbox{ for any }T>0, (2.13)

where 𝒫:=β+κ3σ\mathcal{P}:=\beta+\kappa\partial_{3}\sigma.

The existence of strong solutions with small initial perturbations to the initial-boundary value problem (2.3)–(2.5) in a horizontally periodic domain Ωp:=𝕋2×(0,h)\Omega_{\mathrm{p}}:=\mathbb{T}^{2}\times(0,h) has been established by Jiang–Zhang–Zhang, where 𝕋:=/\mathbb{T}:=\mathbb{R}/\mathbb{Z}. Let us first roughly recall some key ideas in Jiang–Zhang–Zhang’s proof in [35].

The standing point of Jiang–Zhang–Zhang’s proof is the basic energy identity in differential version for the problem (2.3)–(2.5):

12ddt(κϱL2(Ωp)2+ρuL2(Ωp)2)+μuL2(Ωp)2=0.\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}(\|\kappa\nabla\varrho\|_{L^{2}(\Omega_{\mathrm{p}})}^{2}+\|\sqrt{\rho}u\|_{L^{2}(\Omega_{\mathrm{p}})}^{2})+\mu\|\nabla u\|_{L^{2}(\Omega_{\mathrm{p}})}^{2}=0. (2.14)

We call (2.14) the zero-order energy estimate of (ϱ,u)(\varrho,u). Then they further derived energy estimates for the high-order spacial derivatives and the temporal derivatives of (ϱ,u)(\varrho,u). However, the integrals related to the nonlinear terms in the system (2.3) appear in the high-order energy estimates. We call such integrals the nonlinear integrals for simplicity. In particular, there exist a troublesome nonlinear integral (33ϱ)23u3dx\int(\partial_{3}^{3}\varrho)^{2}\partial_{3}u_{3}\mathrm{d}x, which is equal to (33ϱ)2divhuhdx-\int(\partial_{3}^{3}\varrho)^{2}\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}}\mathrm{d}x due to the incompressible condition (2.3)3. Jiang–Zhang–Zhang naturally expect that hu\nabla_{\mathrm{h}}u enjoy a fine decay-in-time, which contributes to close the troublesome nonlinear integral. However one can not directly derive the decay-in-time of hu\nabla_{\mathrm{h}}u from high-order energy estimates due to the absence of the dissipation of hϱ\nabla_{\mathrm{h}}\varrho. Fortunately, they can capture the dissipation of hϱ\nabla_{\mathrm{h}}\varrho from the vortex equations, which can be obtained by applying curl\mathrm{curl} to (2.3)2. The dissipation of (hϱ,hu)(\nabla_{\mathrm{h}}\varrho,\nabla_{\mathrm{h}}u), together with the horizontal periodicity of Ωp\Omega_{\mathrm{p}}, results in the decay-in-time of hϱ\nabla_{\mathrm{h}}\varrho and hu\nabla_{\mathrm{h}}u by energy method with extremely fine estimates.

Following Jiang–Zhang–Zhang’s ideas in [35] for the initial-boundary vaule problem (2.5)–(2.7), however it seems to be difficult to establish the desired decay-in-time of hu\nabla_{\mathrm{h}}u due to the unboundedness of the fluid domain Ω\Omega. This results in that we shall develop a new alternative method to estimate for the troublesome nonlinear integral (33σ)23u3dx\int(\partial_{3}^{3}\sigma)^{2}\partial_{3}u_{3}\mathrm{d}x (recalling the linear relation σ=κϱ\sigma=\kappa\varrho). More precisely, we use the transport equation (2.6)1 twice and the anisotropic Gagliardo–Nirenberg–Sobolev type estimate in A.6 to estimate (33σ)23u3dx\int(\partial_{3}^{3}\sigma)^{2}\partial_{3}u_{3}\mathrm{d}x, see (3.20) for the detailed derivation. Based on this new idea, we can refine the energy method in [35] to establish Theorem 2.1.

It is easy see from (2.9) that χ\chi for κ1\kappa\geqslant 1 reduces to χ=c2\chi=c_{2}. Consequently, we can make use of the uniform-in-κ\kappa estimates in (2.11)–(2.13) with κ1\kappa\geqslant 1 to establish the asymptotical behavior with respect to capillarity number at infinity. More precisely, for any given κ>0\kappa>0, we choose initial data σκ0\sigma^{0}_{\kappa} and uκ0u^{0}_{\kappa}, which satisfy (2.8) and (2.10) with (σκ0,uκ0)(\sigma^{0}_{\kappa},u^{0}_{\kappa}) in place of (σ0,u0)(\sigma^{0},u^{0}). In view of Theorem 2.1, the initial-boundary value problem (2.3)–(2.5) with (ϱκ0,uκ0)(\varrho^{0}_{\kappa},u^{0}_{\kappa}) in place of (ϱ0,u0)(\varrho^{0},u^{0}) admits a global strong solution denoted by (ϱκ,uκ)(\varrho^{\kappa},u^{\kappa}), where we have defined that ϱκ0:=κ1σκ0\varrho^{0}_{\kappa}:=\kappa^{-1}\sigma^{0}_{\kappa}. Moreover the solutions have the following asymptotical behavior with respect to κ\kappa at infinity

Theorem 2.2.

Let p1p\geqslant 1. We additionally assume that (uκ0)hwh0(u^{0}_{\kappa})_{\mathrm{h}}\to w^{0}_{\mathrm{h}} in H2H^{2} as κ\kappa\to\infty. There exist functions u~{\tilde{u}}, 𝒬\mathcal{Q}, \mathcal{M} and 𝒩3\mathcal{N}_{3} such that, for any given T>0T>0,

uκu~ weakly* in L(IT,s2) and weakly in L2(IT,H3) wit u~3=0,\displaystyle u^{\kappa}\rightarrow{\tilde{u}}\mbox{ weakly* in }L^{\infty}(I_{T},{\mathcal{H}^{2}_{\mathrm{s}}})\mbox{ and weakly in }L^{2}(I_{T},H^{3})\mbox{ wit }\tilde{u}_{3}=0, (2.15)
uκu~ strongly in Lp(IT,Hloc2)C0(IT¯,Hloc1),\displaystyle u^{\kappa}\rightarrow{\tilde{u}}\mbox{ strongly in }L^{p}(I_{T},H^{2}_{\mathrm{loc}})\cap C^{0}(\overline{I_{T}},H^{1}_{\mathrm{loc}}), (2.16)
utκu~t weakly* in L(IT,L2) and weakly in L2(IT,1),\displaystyle u_{t}^{\kappa}\rightarrow\tilde{u}_{t}\mbox{ weakly* in }L^{\infty}(I_{T},L^{2})\mbox{ and weakly in }L^{2}(I_{T},{\mathcal{H}^{1}}), (2.17)
(ϱκ,ϱtκ)(0,0), strongly in L(IT,H3)×L2(IT,H2),\displaystyle(\varrho^{\kappa},\varrho^{\kappa}_{t})\to(0,0),\mbox{ strongly in }L^{\infty}(I_{T},H^{3})\times L^{2}(I_{T},H^{2}), (2.18)
𝒫κ𝒬 weakly* in L(IT,L2) and weakly in L2(IT,H1),\displaystyle\nabla\mathcal{P}^{\kappa}\rightarrow\nabla\mathcal{Q}\mbox{ weakly* in }L^{\infty}(I_{T},L^{2})\mbox{ and weakly in }L^{2}(I_{T},H^{1}), (2.19)
κ((h3σκ),Δhσκ)(h,𝒩3)\displaystyle\kappa((\nabla_{\mathrm{h}}\partial_{3}\sigma^{\kappa})^{\top},-\Delta_{\rm h}\sigma^{\kappa})^{\top}\rightarrow(\nabla_{\mathrm{h}}\mathcal{M}^{\top},\mathcal{N}_{3})^{\top}
weakly* in L(IT,L2) and weakly in L2(IT,H01),\displaystyle\mbox{weakly* in }L^{\infty}(I_{T},L^{2})\mbox{ and weakly in }L^{2}(I_{T},H^{1}_{0}), (2.20)

as κ\kappa\rightarrow\infty. Moreover, u~h{\tilde{u}}_{\mathrm{h}}, 𝒬\mathcal{Q}, \mathcal{M} and 𝒩3\mathcal{N}_{3} satisfy

{ρ¯(tu~h+u~hhu~h)+h(𝒬)μΔhu~h=μ32u~h,divhu~h=0,u~h|t=0=wh0 with divhwh0=0\displaystyle\begin{cases}\bar{\rho}(\partial_{t}{\tilde{u}}_{\mathrm{h}}+{\tilde{u}}_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}{\tilde{u}}_{\mathrm{h}})+\nabla_{\rm h}(\mathcal{Q}-\mathcal{M})-\mu\Delta_{\mathrm{h}}{\tilde{u}}_{\mathrm{h}}=\mu\partial_{3}^{2}{\tilde{u}}_{\mathrm{h}},\\ \mathrm{div}_{\mathrm{h}}{\tilde{u}}_{\mathrm{h}}=0,\\ {\tilde{u}}_{\mathrm{h}}|_{t=0}=w_{\mathrm{h}}^{0}\mbox{ with }\mathrm{div}_{\mathrm{h}}w_{\mathrm{h}}^{0}=0\end{cases} (2.21)

and

{Δ𝒬=ρ¯u~h:hu~h,3𝒬|Ω=0,Δh=32𝒬,𝒩3=3𝒬.\displaystyle\begin{cases}-\Delta\mathcal{Q}=\bar{\rho}\nabla{\tilde{u}}_{\mathrm{h}}:\nabla_{\mathrm{h}}\tilde{u}_{\mathrm{h}},\ \partial_{3}\mathcal{Q}|_{\partial\Omega}=0,\\ -\Delta_{\mathrm{h}}\mathcal{M}=\partial_{3}^{2}\mathcal{Q},\\ \mathcal{N}_{3}=\partial_{3}\mathcal{Q}.\end{cases} (2.22)
Remark 2.1.

Modifying the proofs of Theorems 2.1 and 2.2 by further using the energy method with time-weight in [35], we also obtain the correspond results for the case of the horizontally periodic domain with finite height, i.e. 2πL1𝕋×2πL2𝕋×(0,h)2\pi L_{1}\mathbb{T}\times 2\pi L_{2}\mathbb{T}\times(0,h). In addition, since χ\chi\to\infty for κ0\kappa\to 0, we can not except the vanishing capillarity limit from Theorem 2.1.

We will prove the asymptotical behavior in Theorem 2.2 by exploiting a compactness argument, the details of which will be presented in Section 5.

3 Proof of Theorem 2.1

This section is devoted to the proof of the global(-in-time) solvability with uniform-in-κ\kappa estimates for the initial-boundary value problem (2.5)–(2.7). The key point is to a priori derive the uniform-in-κ\kappa estimate in (2.11). For this purpose, let T>0T>0 be a fixed time and (σ,u)(\sigma,u) a solution to (2.5)–(2.7) on Ω×IT\Omega\times I_{T} with initial data (σ0,u0)Hρ¯3×s2(\sigma^{0},u^{0})\in{{H}}^{3}_{\bar{\rho}}\times{\mathcal{H}^{2}_{\mathrm{s}}}. Moreover, we assume that σ0\sigma^{0} satisfies (2.8) and the solution (σ,u)(\sigma,u) is sufficiently regular so that the procedure of formal deduction makes sense.

3.1 Basic estimates

This section is devoted to deriving some basic estimates of (σ,u)(\sigma,u) from the initial-boundary value problem (2.5)–(2.7). Let us first recall the energy identity of (σ,u)(\sigma,u).

Lemma 3.1.

It holds that

12ddt(σ02+ρu02)+μu02=0,\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}(\|\nabla\sigma\|_{0}^{2}+\|\sqrt{\rho}u\|_{0}^{2})+\mu\|\nabla u\|_{0}^{2}=0, (3.1)

where ρ=ρ¯+σ/κ\rho=\bar{\rho}+\sigma/\kappa and ρ¯\bar{\rho} is defined by (2.1) with a=b=1a=b=1 in (2.2).

Proof.

Taking the inner product of (2.6)1 and Δσ-\Delta\sigma in L2L^{2}, and then exploiting the integration by parts and the boundary condition of σ\sigma in (A.16), we have

12ddt|σ|2dx=Δσ(κu3+uσ)dx.\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\int|\nabla\sigma|^{2}\mathrm{d}x=\int\Delta\sigma(\kappa u_{3}+u\cdot\nabla\sigma)\mathrm{d}x.

Similarly, taking the inner product of (2.6)2 and uu in L2L^{2}, and then making using the integration by parts, the mass equation (1.1)1\eqref{1.1}_{1}, the boundary condition (2.5) and the incompressible condition (2.6)3, we obtain

12ddtρ|u|2dx+μ|u|2dx\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\int\rho|u|^{2}\mathrm{d}x+\mu\int|\nabla u|^{2}\mathrm{d}x
=\displaystyle= 12ρt|u|2dx(Δσ(κ𝐞3+σ)+ρuu)udx\displaystyle\frac{1}{2}\int\rho_{t}|u|^{2}\mathrm{d}x-\int(\Delta{\sigma}(\kappa\mathbf{e}^{3}+\nabla{\sigma})+{\rho}u\cdot\nabla u)\cdot u\mathrm{d}x
=\displaystyle= Δσ(κu3+σu)dx.\displaystyle-\int\Delta{\sigma}(\kappa u_{3}+\nabla{\sigma}\cdot u)\mathrm{d}x.

Summing up the above two identities yields (3.1). ∎

Next we further extend (3.1) to both the cases satisfied by the highest-order spacial derivatives and the temporal derivatives of (σ,u)(\sigma,u) resp..

Lemma 3.2.

It holds that

12ddt(Δσ02+ρΔu023(κ172κ23σ)3σ(33σ)2dx)+μΔu02\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\left(\|\nabla\Delta\sigma\|_{0}^{2}+\|\sqrt{\rho}\Delta u\|^{2}_{0}-3\int\left(\kappa^{-1}-\frac{7}{2\kappa^{2}}\partial_{3}\sigma\right)\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x\right)+\mu\|\nabla\Delta u\|^{2}_{0}
(1+κ1σ3)u1(ut1+u2u2)+σ1,2(σ3+κ2σ33)u2\displaystyle\lesssim(1+\kappa^{-1}\|\sigma\|_{3})\|\nabla u\|_{1}(\|u_{t}\|_{1}+\|u\|_{2}\|\nabla u\|_{2})+\|\sigma\|_{1,2}(\|\sigma\|_{3}+\kappa^{-2}\|\sigma\|_{3}^{3})\|\nabla u\|_{2} (3.2)

and

ddt(σt02+ρut02)+μut02\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}(\|\nabla\sigma_{t}\|_{0}^{2}+\|\sqrt{{\rho}}u_{t}\|^{2}_{0})+\mu\|\nabla u_{t}\|^{2}_{0}
σt2(σt1u2+(σ3+κ1u22)ut0)\displaystyle\lesssim\|\sigma_{t}\|_{2}\left(\|\sigma_{t}\|_{1}\|u\|_{2}+(\|\sigma\|_{3}+\kappa^{-1}\|u\|_{2}^{2})\|u_{t}\|_{0}\right)
+(κ1σt1+(1+κ1σ2)u2)ut0ut1.\displaystyle\quad+(\kappa^{-1}\|\sigma_{t}\|_{1}+(1+\kappa^{-1}\|\sigma\|_{2})\|u\|_{2})\|u_{t}\|_{0}\|u_{t}\|_{1}. (3.3)
Proof.

(1) Let 1i31\leqslant i\leqslant 3. Applying Δ\Delta and i\partial_{i} to (2.6)1 and (2.6)2 resp., we obtain

{Δ(σt+uσ+κu3)=0,ρ(iut+uiu)+iβ=i(μΔuΔσ(κ𝐞3+σ))iρuti(ρu)u.\displaystyle\begin{cases}\Delta(\sigma_{t}+u\cdot\nabla\sigma+{\kappa}u_{3})=0,\\ \rho(\partial_{i}u_{t}+u\cdot\nabla\partial_{i}u)+\nabla\partial_{i}\beta\\ =\partial_{i}(\mu\Delta u-\Delta{\sigma}(\kappa\mathbf{e}^{3}+\nabla\sigma))-\partial_{i}\rho u_{t}-\partial_{i}(\rho u)\cdot\nabla u.\end{cases} (3.4)

Taking the inner product of (3.4)1 and Δ2σ-\Delta^{2}\sigma in L2L^{2}, and then using the integration by parts and the boundary conditions of (σ,32σ,u3)(\sigma,\partial_{3}^{2}\sigma,u_{3}) in (2.5), (A.16) yields

12ddtΔσ02=(Δ2σΔ(uσ)κΔσΔu3)dx.\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\|\nabla\Delta\sigma\|^{2}_{0}=\int(\Delta^{2}\sigma\Delta(u\cdot\nabla\sigma)-\kappa\nabla\Delta\sigma\cdot\nabla\Delta u_{3})\mathrm{d}x. (3.5)

Exploiting the integration by parts, (1.1)1 and the boundary condition of u3u_{3} in (2.5), we can derive that

ρuΔuΔudx=12div(ρu)|Δu|2dx=12ρt|Δu|2dx.\displaystyle\int\rho u\cdot\nabla\Delta u\cdot\Delta u\mathrm{d}x=-\frac{1}{2}\int\mathrm{div}(\rho u)|\Delta u|^{2}\mathrm{d}x=\frac{1}{2}\int\rho_{t}|\Delta u|^{2}\mathrm{d}x. (3.6)

In view of the incompressible condition, the boundary condition of (ϱ,32ϱ,u3,3uh)(\varrho,\partial_{3}^{2}\varrho,u_{3},\partial_{3}u_{\mathrm{h}}) in (2.5), (A.16) and the equation (2.6)2, it is to see that

(32u3,3β)|Ω=0.\displaystyle(\partial_{3}^{2}u_{3},\partial_{3}\beta)|_{\partial\Omega}=0. (3.7)

Taking the inner product of (3.4)2 and iΔu-\partial_{i}\Delta u in L2L^{2}, and then making use of the incompressible condition, the integration by parts, the boundary conditions of (σ,3ρ,u3,3uh,32u3,3β)(\sigma,\partial_{3}\rho,u_{3},\partial_{3}u_{\mathrm{h}},\partial_{3}^{2}u_{3},\partial_{3}\beta) in (2.5), (3.7), (A.16), and the identity (3.6), we deduce that

12ddtρΔu02+μΔu02\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\|\sqrt{\rho}\Delta u\|^{2}_{0}+\mu\|\nabla\Delta u\|^{2}_{0}
=\displaystyle= (i(Δσ(κ𝐞3+σ))Δiu(2ρut\displaystyle\int\bigg{(}\partial_{i}(\Delta\sigma(\kappa\mathbf{e}^{3}+\nabla\sigma))\cdot\Delta\partial_{i}u-(2\nabla\rho\cdot\nabla u_{t}
+Δρut+2i(ρu)iu+Δ(ρu)u)Δu)dx.\displaystyle+\Delta\rho u_{t}+2\partial_{i}(\rho u)\cdot\partial_{i}\nabla u+\Delta(\rho u)\cdot\nabla u)\cdot\Delta u\bigg{)}\mathrm{d}x. (3.8)

Putting (3.5) and (3.8) together yields

12ddt(Δσ02+ρΔu02)+μΔu02\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}(\|\nabla\Delta\sigma\|_{0}^{2}+\|\sqrt{\rho}\Delta u\|^{2}_{0})+\mu\|\nabla\Delta u\|^{2}_{0}
=\displaystyle= (iΔui(Δσσ)+Δ2σΔ(uσ))dx(2ρut\displaystyle\int(\partial_{i}\Delta u\cdot\partial_{i}(\Delta\sigma\nabla\sigma)+\Delta^{2}\sigma\Delta(u\cdot\nabla\sigma))\mathrm{d}x-\int(2\nabla\rho\cdot\nabla u_{t}
+Δρut+2i(ρu)iu+Δ(ρu)u)Δudx=:J2J1,\displaystyle+\Delta\rho u_{t}+2\partial_{i}(\rho u)\cdot\nabla\partial_{i}u+\Delta(\rho u)\cdot\nabla u)\cdot\Delta u\mathrm{d}x=:J_{2}-J_{1}, (3.9)

where we have used the Einstein convention of summation over repeated indices. Next we shall estimate for J1J_{1} and J2J_{2} in sequence.

Using the definition of ρ\rho and the imbedding inequality (A.1), we get

iρL=i(ρ¯+κ1σ)L1+κ1iσ2 for i=0, 1.\displaystyle\|\nabla^{i}\rho\|_{L^{\infty}}=\|\nabla^{i}(\bar{\rho}+\kappa^{-1}\sigma)\|_{L^{\infty}}\lesssim 1+\kappa^{-1}\|\nabla^{i}\sigma\|_{2}\mbox{ for }i=0,\ 1. (3.10)

Making use of Hölder inequality, the product estimates in (A.12) and (3.10), we can estimate that

J1(1+κ1σ3)u1(ut1+u2u2).\displaystyle J_{1}\lesssim(1+\kappa^{-1}\|\sigma\|_{3})\|\nabla u\|_{1}(\|u_{t}\|_{1}+\|u\|_{2}\|\nabla u\|_{2}). (3.11)

Now we turn to estimating for J2J_{2}. Using the incompressible condition, the integration by parts and the boundary conditions of (σ,32σ,u3,3uh,32u3)(\sigma,\partial_{3}^{2}\sigma,u_{3},\partial_{3}u_{\mathrm{h}},\partial_{3}^{2}u_{3}) in (2.5), (3.7), (A.16), it holds that

Δ2σuΔσdx=(Δσ)uΔσdx,\displaystyle\int\Delta^{2}\sigma u\cdot\nabla\Delta\sigma\mathrm{d}x=-\int(\nabla\Delta\sigma\cdot\nabla)u\cdot\nabla\Delta\sigma\mathrm{d}x,
Δ2σjujσdx=iΔσ(juijσ+ijujσ)dx.\displaystyle\int\Delta^{2}\sigma\partial_{j}u\cdot\nabla\partial_{j}\sigma\mathrm{d}x=-\int\partial_{i}\Delta\sigma(\partial_{j}u\cdot\nabla\partial_{i}\partial_{j}\sigma+\partial_{i}\partial_{j}u\cdot\nabla\partial_{j}\sigma)\mathrm{d}x.

and

Δ2σΔuσdx=Δσ(Δuσ)dx.\displaystyle\int\Delta^{2}\sigma\Delta u\cdot\nabla\sigma\mathrm{d}x=-\int\nabla\Delta\sigma\cdot\nabla(\Delta u\cdot\nabla\sigma)\mathrm{d}x.

We can use the above three identities and the integration by parts to rewrite J2J_{2} as follows:

J2=\displaystyle J_{2}= (ΔσiσiΔuiΔσ(iuΔσ+2(juijσ\displaystyle\int(\Delta\sigma\nabla\partial_{i}\sigma\cdot\partial_{i}\Delta u-\partial_{i}\Delta\sigma(\partial_{i}u\cdot\nabla\Delta\sigma+2(\partial_{j}u\cdot\nabla\partial_{i}\partial_{j}\sigma
+ijujσ)+Δuiσ))dx\displaystyle+\partial_{i}\partial_{j}u\cdot\nabla\partial_{j}\sigma)+\Delta u\cdot\nabla\partial_{i}\sigma))\mathrm{d}x
=\displaystyle= (iΔσ(iuΔσ+2(juijσ+Δuiσ\displaystyle-\int(\partial_{i}\Delta\sigma(\partial_{i}u\cdot\nabla\Delta\sigma+2(\partial_{j}u\cdot\nabla\partial_{i}\partial_{j}\sigma+\Delta u\cdot\nabla\partial_{i}\sigma
+ijujσ))+ΔσΔuΔσ)dx=J2,1+J2,2+J2,3,\displaystyle+\partial_{i}\partial_{j}u\cdot\nabla\partial_{j}\sigma))+\Delta\sigma\Delta u\cdot\nabla\Delta\sigma)\mathrm{d}x=J_{2,1}+J_{2,2}+J_{2,3}, (3.12)

where we have defined that

J2,1:=\displaystyle J_{2,1}:= (i=12iΔσ(iuΔσ+2(juijσ+Δuiσ\displaystyle-\int\bigg{(}\sum_{i=1}^{2}\partial_{i}\Delta\sigma(\partial_{i}u\cdot\nabla\Delta\sigma+2(\partial_{j}u\cdot\nabla\partial_{i}\partial_{j}\sigma+\Delta u\cdot\nabla\partial_{i}\sigma
+ijujσ))+3Δhσ(3uΔσ+2(ju3jσ+3jujσ\displaystyle+\partial_{i}\partial_{j}u\cdot\nabla\partial_{j}\sigma))+\partial_{3}\Delta_{\mathrm{h}}\sigma(\partial_{3}u\cdot\nabla\Delta\sigma+2(\partial_{j}u\cdot\nabla\partial_{3}\partial_{j}\sigma+\partial_{3}\partial_{j}u\cdot\nabla\partial_{j}\sigma
+Δu3σ))+33σ(3uhhΔσ+3u33Δhσ+2(juhh3jσ\displaystyle+\Delta u\cdot\nabla\partial_{3}\sigma))+\partial_{3}^{3}\sigma(\partial_{3}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\Delta\sigma+\partial_{3}u_{3}\partial_{3}\Delta_{\mathrm{h}}\sigma+2(\partial_{j}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\partial_{3}\partial_{j}\sigma
+hu3h32σ+3juhhjσ+3hu33hσ+Δuhh3σ))\displaystyle+\nabla_{\mathrm{h}}u_{3}\cdot\nabla_{\mathrm{h}}\partial_{3}^{2}\sigma+\partial_{3}\partial_{j}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\partial_{j}\sigma+\partial_{3}\nabla_{\mathrm{h}}u_{3}\cdot\partial_{3}\nabla_{\mathrm{h}}\sigma+\Delta u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\partial_{3}\sigma))
+ΔhσΔuΔσ+32σ(ΔuhhΔσ+Δu3Δh3σ))dx,\displaystyle+\Delta_{\mathrm{h}}\sigma\Delta u\cdot\nabla\Delta\sigma+\partial_{3}^{2}\sigma(\Delta u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\Delta\sigma+\Delta u_{3}\Delta_{\mathrm{h}}\partial_{3}\sigma)\bigg{)}\mathrm{d}x,
J2,2:=\displaystyle J_{2,2}:= 32σ33σ(232u3+3Δu3)dx and J2,3:=3(33σ)23u3dx.\displaystyle-\int\partial_{3}^{2}\sigma\partial_{3}^{3}\sigma(2\partial_{3}^{2}u_{3}+3\Delta u_{3})\mathrm{d}x\mbox{ and }J_{2,3}:=-3\int(\partial_{3}^{3}\sigma)^{2}\partial_{3}u_{3}\mathrm{d}x.

Take a similar procedure as (3.11), we can easily obtain

J2,1σ1,2σ3u2.\displaystyle J_{2,1}\lesssim\|\sigma\|_{1,2}\|\sigma\|_{3}\|\nabla u\|_{2}. (3.13)

Making use of the incompressible condition, the integration by parts, the product estimate, and the boundary condition of 32σ\partial_{3}^{2}\sigma in (A.16), we deduce that

J2,2=\displaystyle J_{2,2}= 12(32σ)23(232u3+3Δu3)dx=12(32σ)2(232divhuh+3Δdivhuh)dx\displaystyle\frac{1}{2}\int(\partial_{3}^{2}\sigma)^{2}\partial_{3}(2\partial_{3}^{2}u_{3}+3\Delta u_{3})\mathrm{d}x=-\frac{1}{2}\int(\partial_{3}^{2}\sigma)^{2}(2\partial_{3}^{2}\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}}+3\Delta\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}})\mathrm{d}x
=\displaystyle= 12(232uh+3Δuh)h(32σ)2dxσ1,2σ3u2.\displaystyle\frac{1}{2}\int(2\partial_{3}^{2}u_{\mathrm{h}}+3\Delta u_{\mathrm{h}})\cdot\nabla_{\mathrm{h}}(\partial_{3}^{2}\sigma)^{2}\mathrm{d}x\lesssim\|\sigma\|_{1,2}\|\sigma\|_{3}\|\nabla u\|_{2}. (3.14)

Due to the absence of the dissipation of 33σ\partial_{3}^{3}\sigma, we can not directly estimate for J2,3J_{2,3}. To overcome this difficulty, we shall use equation (2.6)1\eqref{1.7}_{1} twice as follows.

J2,3/3\displaystyle J_{2,3}/3
=\displaystyle= κ1(33σ)23(σt+uσ)dx\displaystyle{\kappa}^{-1}\int(\partial_{3}^{3}\sigma)^{2}\partial_{3}(\sigma_{t}+u\cdot\nabla\sigma)\mathrm{d}x
=\displaystyle= κ1((33σ)23(uσ)23σ33σ33σt)dx+κ1ddt3σ(33σ)2dx\displaystyle{\kappa}^{-1}\int((\partial_{3}^{3}\sigma)^{2}\partial_{3}(u\cdot\nabla\sigma)-{2}\partial_{3}\sigma\partial_{3}^{3}\sigma\partial_{3}^{3}\sigma_{t})\mathrm{d}x+{\kappa}^{-1}\frac{\mathrm{d}}{\mathrm{d}t}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
=\displaystyle= κ1(23σ33σ33(κu3+uσ)+(33σ)23(uσ))dx+κ1ddt3σ(33σ)2dx\displaystyle{\kappa}^{-1}\int({2}\partial_{3}\sigma\partial_{3}^{3}\sigma\partial_{3}^{3}({\kappa}u_{3}+u\cdot\nabla\sigma)+(\partial_{3}^{3}\sigma)^{2}\partial_{3}(u\cdot\nabla\sigma))\mathrm{d}x+{\kappa}^{-1}\frac{\mathrm{d}}{\mathrm{d}t}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
=\displaystyle= κ1(23σ33σ(κ33u3+33uσ+332u3σ\displaystyle{\kappa}^{-1}\int({2}\partial_{3}\sigma\partial_{3}^{3}\sigma({\kappa}\partial_{3}^{3}u_{3}+\partial_{3}^{3}u\cdot\nabla\sigma+3\partial_{3}^{2}u\cdot\nabla\partial_{3}\sigma
+33u32σ)+(33σ)23uσ)dx+κ1ddt3σ(33σ)2dx\displaystyle+3\partial_{3}u\cdot\nabla\partial_{3}^{2}\sigma)+(\partial_{3}^{3}\sigma)^{2}\partial_{3}u\cdot\nabla\sigma)\mathrm{d}x+{\kappa}^{-1}\frac{\mathrm{d}}{\mathrm{d}t}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
=\displaystyle= κ133σ(23σ(33uhhσ+332uhh3σ+33uhh32σ\displaystyle{\kappa}^{-1}\int\partial_{3}^{3}\sigma(2\partial_{3}\sigma(\partial_{3}^{3}u_{\rm h}\cdot\nabla_{\rm h}\sigma+3\partial_{3}^{2}u_{\rm h}\cdot\nabla_{\rm h}\partial_{3}\sigma+3\partial_{3}u_{\rm h}\cdot\nabla_{\rm h}\partial_{3}^{2}\sigma
+33σ3uhhσ))dx+23σ33σ(33u3+κ1(3σ33u3+332σ32u3))dx\displaystyle+\partial_{3}^{3}\sigma\partial_{3}u_{\rm h}\cdot\nabla_{\rm h}\sigma))\mathrm{d}x+2\int\partial_{3}\sigma\partial_{3}^{3}\sigma(\partial_{3}^{3}u_{3}+{\kappa}^{-1}(\partial_{3}\sigma\partial_{3}^{3}u_{3}+3\partial_{3}^{2}\sigma\partial_{3}^{2}u_{3}))\mathrm{d}x
+7κ13σ(33σ)23u3dx+κ1ddt3σ(33σ)2dx\displaystyle+{7}{\kappa}^{-1}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\partial_{3}u_{3}\mathrm{d}x+{\kappa}^{-1}\frac{\mathrm{d}}{\mathrm{d}t}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
=:\displaystyle=: J12,3+2J22,3+7J32,3+κ1ddt3σ(33σ)2dx,\displaystyle J^{2,3}_{1}+2J^{2,3}_{2}+7J^{2,3}_{3}+{\kappa}^{-1}\frac{\mathrm{d}}{\mathrm{d}t}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x, (3.15)

where in the fourth equality we have used the identity

33σ(33σu3σ+23σu33σ)dx=u(3σ(33σ)2)dx=0.\displaystyle\int\partial_{3}^{3}\sigma(\partial_{3}^{3}\sigma u\cdot\nabla\partial_{3}\sigma+2\partial_{3}\sigma u\cdot\nabla\partial_{3}^{3}\sigma)\mathrm{d}x=\int u\cdot\nabla(\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2})\mathrm{d}x=0.

Next we shall estimate the three terms J12,3J^{2,3}_{1}J32,3J^{2,3}_{3} in sequence.

It follows from the product estimates that

J12,3κ1σ1,2σ32u2.\displaystyle J^{2,3}_{1}\lesssim\kappa^{-1}\|\sigma\|_{1,2}\|\sigma\|_{3}^{2}\|\nabla u\|_{2}. (3.16)

Exploiting the incompressible condition, the integration by parts, the product estimates and the boundary condition of 32σ\partial_{3}^{2}\sigma in (A.16), we have

J22,3=\displaystyle J^{2,3}_{2}= 33σ(3σ32divhuh+κ13σ(3σ32divhuh+332σ3divhuh))dx\displaystyle-\int\partial_{3}^{3}\sigma(\partial_{3}\sigma\partial_{3}^{2}\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}}+{\kappa}^{-1}\partial_{3}\sigma(\partial_{3}\sigma\partial_{3}^{2}\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}}+3\partial_{3}^{2}\sigma\partial_{3}\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}}))\mathrm{d}x
=\displaystyle= 33σ(32uhh3σ+κ1(32uhh(3σ)2+33uhh(3σ32σ)))dx\displaystyle\int\partial_{3}^{3}\sigma(\partial_{3}^{2}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\partial_{3}\sigma+{\kappa}^{-1}(\partial_{3}^{2}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}(\partial_{3}\sigma)^{2}+3\partial_{3}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}(\partial_{3}\sigma\partial_{3}^{2}\sigma)))\mathrm{d}x
3(3σ32uh+κ13σ(3σ32uh+332σ3uh))h32σdx\displaystyle-\int\partial_{3}(\partial_{3}\sigma\partial_{3}^{2}u_{\mathrm{h}}+{\kappa}^{-1}\partial_{3}\sigma(\partial_{3}\sigma\partial_{3}^{2}u_{\mathrm{h}}+3\partial_{3}^{2}\sigma\partial_{3}u_{\mathrm{h}}))\cdot\nabla_{\mathrm{h}}\partial_{3}^{2}\sigma\mathrm{d}x
\displaystyle\lesssim (1+κ1σ3)σ1,2σ3u2.\displaystyle(1+\kappa^{-1}\|\sigma\|_{3})\|\sigma\|_{1,2}\|\sigma\|_{3}\|\nabla u\|_{2}. (3.17)

Similar to J2,3J_{2,3}, we shall use the equation (2.6)1\eqref{1.7}_{1} twice to rewrite J32,3J_{3}^{2,3} as follows:

J32,3=\displaystyle J_{3}^{2,3}= κ23σ(33σ)23(σt+uσ)dx\displaystyle-{\kappa}^{-2}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\partial_{3}(\sigma_{t}+u\cdot\nabla\sigma)\mathrm{d}x
=\displaystyle= κ2((3σ)233σ33σt3σ(33σ)23(uσ))dx12κ2ddt(3σ33σ)2dx\displaystyle{\kappa}^{-2}\int\left({(\partial_{3}\sigma)^{2}}\partial_{3}^{3}\sigma\partial_{3}^{3}\sigma_{t}-\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\partial_{3}(u\cdot\nabla\sigma)\right)\mathrm{d}x-\frac{1}{2{\kappa}^{2}}\frac{\mathrm{d}}{\mathrm{d}t}\int(\partial_{3}\sigma\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
=\displaystyle= κ23σ33σ(3σ33(κu3+uσ)+33σ3(uσ))dx\displaystyle-{\kappa}^{-2}\int\partial_{3}\sigma\partial_{3}^{3}\sigma\left({\partial_{3}\sigma}\partial_{3}^{3}({\kappa}u_{3}+u\cdot\nabla\sigma)+\partial_{3}^{3}\sigma\partial_{3}(u\cdot\nabla\sigma)\right)\mathrm{d}x
12κ2ddt(3σ33σ)2dx\displaystyle-\frac{1}{2{\kappa}^{2}}\frac{\mathrm{d}}{\mathrm{d}t}\int(\partial_{3}\sigma\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
=\displaystyle= κ1(3σ)233σ33u3dxκ23σ(33σ)23uσdxκ2(3σ)233σ(33uσ\displaystyle-\kappa^{-1}\int(\partial_{3}\sigma)^{2}\partial_{3}^{3}\sigma\partial_{3}^{3}u_{3}\mathrm{d}x-{\kappa}^{-2}\int\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\partial_{3}u\cdot\nabla\sigma\mathrm{d}x-{\kappa}^{-2}\int(\partial_{3}\sigma)^{2}\partial_{3}^{3}\sigma(\partial_{3}^{3}u\cdot\nabla\sigma
+332u3σ+33u32σ)dx12κ2ddt(3σ33σ)2dx\displaystyle+3\partial_{3}^{2}u\cdot\nabla\partial_{3}\sigma+3\partial_{3}u\cdot\nabla\partial_{3}^{2}\sigma)\mathrm{d}x-\frac{1}{2{\kappa}^{2}}\frac{\mathrm{d}}{\mathrm{d}t}\int(\partial_{3}\sigma\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
=:\displaystyle=: J3,12,3+J3,22,3+J3,32,312κ2ddt(3σ33σ)2dx.\displaystyle J^{2,3}_{3,1}+J^{2,3}_{3,2}+J^{2,3}_{3,3}-\frac{1}{2{\kappa}^{2}}\frac{\mathrm{d}}{\mathrm{d}t}\int(\partial_{3}\sigma\partial_{3}^{3}\sigma)^{2}\mathrm{d}x. (3.18)

By the incompressible condition, the integration by parts, the product estimate, and the boundary condition of 32σ\partial_{3}^{2}\sigma in  (A.16), we have

J3,12,3=\displaystyle J^{2,3}_{3,1}= κ1(3σ)233σ32divhuhdx\displaystyle\kappa^{-1}\int(\partial_{3}\sigma)^{2}\partial_{3}^{3}\sigma\partial_{3}^{2}\mathrm{div}_{\rm h}u_{\rm h}\mathrm{d}x
=\displaystyle= 12κ(h32σ3((3σ)232uh)33σ32uhh(3σ)2)dx\displaystyle\frac{1}{2\kappa}\int\left(\nabla_{\rm h}\partial_{3}^{2}\sigma\cdot\partial_{3}((\partial_{3}\sigma)^{2}\partial_{3}^{2}u_{\rm h})-\partial_{3}^{3}\sigma\partial_{3}^{2}u_{\rm h}\cdot\nabla_{\rm h}(\partial_{3}\sigma)^{2}\right)\mathrm{d}x
\displaystyle\lesssim κ1σ1,2σ32u2.\displaystyle\kappa^{-1}\|\sigma\|_{1,2}\|\sigma\|_{3}^{2}\|\nabla u\|_{2}.

In addition, we can use the product estimate and the anisotropic interpolation inequality (A.6) to obtain

J3,22,3\displaystyle J^{2,3}_{3,2}\lesssim κ2σL233σ023u2\displaystyle\kappa^{-2}\|\nabla\sigma\|_{L^{\infty}}^{2}\|\partial_{3}^{3}\sigma\|_{0}^{2}\|\partial_{3}u\|_{2}
\displaystyle\lesssim κ233σ02hσ1σ23u2κ2σ1,2σ33u2.\displaystyle\kappa^{-2}\|\partial_{3}^{3}\sigma\|_{0}^{2}\|\nabla\nabla_{\rm h}\sigma\|_{1}\|\nabla\sigma\|_{2}\|\partial_{3}u\|_{2}\lesssim\kappa^{-2}\|\sigma\|_{1,2}\|\sigma\|_{3}^{3}\|\nabla u\|_{2}.

Similarly, we also get

J3,32,3\displaystyle J^{2,3}_{3,3}\lesssim κ2h3σ13σ22(32σ03u2+3σ132u1+σ233u0)\displaystyle\kappa^{-2}\|\nabla_{\rm h}\partial_{3}\sigma\|_{1}\|\partial_{3}\sigma\|_{2}^{2}(\|\nabla\partial_{3}^{2}\sigma\|_{0}\|\partial_{3}u\|_{2}+\|\nabla\partial_{3}\sigma\|_{1}\|\partial_{3}^{2}u\|_{1}+\|\nabla\sigma\|_{2}\|\partial_{3}^{3}u\|_{0})
\displaystyle\lesssim κ2σ1,2σ33u2.\displaystyle\kappa^{-2}\|\sigma\|_{1,2}\|\sigma\|_{3}^{3}\|\nabla u\|_{2}.

Thus substituting the above three estimates into (3.18) yields

J32,3σ1,2(κ1σ32+κ2σ33)u212κ2ddt(3σ33σ)2dx.\displaystyle J^{2,3}_{3}\leqslant\|\sigma\|_{1,2}(\kappa^{-1}\|\sigma\|_{3}^{2}+\kappa^{-2}\|\sigma\|_{3}^{3})\|\nabla u\|_{2}-\frac{1}{2{\kappa}^{2}}\frac{\mathrm{d}}{\mathrm{d}t}\int(\partial_{3}\sigma\partial_{3}^{3}\sigma)^{2}\mathrm{d}x. (3.19)

Now inserting (3.16), (3.17) and (3.19) into (3.15) and then using Young’s inequality, we obtain

J2,33ddt(κ172κ23σ)3σ(33σ)2dx+cσ1,2(σ3+κ2σ33)u2.\displaystyle J_{2,3}\leqslant{3}\frac{\mathrm{d}}{\mathrm{d}t}\int\left({\kappa^{-1}}-\frac{7}{2{\kappa^{2}}}\partial_{3}\sigma\right)\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x+c\|\sigma\|_{1,2}(\|\sigma\|_{3}+\kappa^{-2}\|\sigma\|_{3}^{3})\|\nabla u\|_{2}. (3.20)

Thanks to the three estimates (3.13), (3.1) and (3.20), we derive from (3.12) that

J23ddt(κ172κ23σ)3σ(33σ)2dx+cσ1,2(σ3+κ2σ33)u2.\displaystyle J_{2}\leqslant{3}\frac{\mathrm{d}}{\mathrm{d}t}\int\left({\kappa^{-1}}-\frac{7}{2{\kappa^{2}}}\partial_{3}\sigma\right)\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x+c\|\sigma\|_{1,2}(\|\sigma\|_{3}+\kappa^{-2}\|\sigma\|_{3}^{3})\|\nabla u\|_{2}. (3.21)

Finally, putting (3.11) and (3.21) into (3.1), we arrive at the desired estimate (3.2).

(2) Applying t\partial_{t} to (2.6)1 and (2.6)2, we get

{t(σt+uσ+κu3)=0,t(ρ(ut+uu)+βμΔu+Δσ(κ𝐞3+σ))=0.\begin{cases}\partial_{t}(\sigma_{t}+u\cdot\nabla\sigma+{\kappa}u_{3})=0,\\ \partial_{t}({\rho}(u_{t}+u\cdot\nabla u)+\nabla\beta-\mu\Delta u+\Delta{\sigma}(\kappa\mathbf{e}^{3}+\nabla{\sigma}))=0.\end{cases} (3.22)

Taking the inner products of (3.22)1 resp. (3.22)2 and Δσt-\Delta\sigma_{t}, resp. utu_{t} in L2L^{2}, then following the argument of (3.1), and finally making use of the product estimates, the relation ρt=κ1σt\rho_{t}=\kappa^{-1}\sigma_{t} and (3.10), we have

12ddt(σt02+ρut02)+μut02\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}(\|\nabla\sigma_{t}\|^{2}_{0}+\|\sqrt{\rho}u_{t}\|^{2}_{0})+\mu\|\nabla u_{t}\|^{2}_{0}
=\displaystyle= (Δσtuσt(ρtut+t(ρu)u+Δσσt)ut)dx\displaystyle\int\left(\Delta{\sigma_{t}}u\cdot\nabla{\sigma_{t}}-(\rho_{t}u_{t}+\partial_{t}({\rho}u)\cdot\nabla u+\Delta{\sigma}\nabla{\sigma_{t}})\cdot u_{t}\right)\mathrm{d}x
\displaystyle\lesssim σt2(σt1u2+(σ3+κ1u22)ut0)\displaystyle\|\sigma_{t}\|_{2}\left(\|\sigma_{t}\|_{1}\|u\|_{2}+(\|\sigma\|_{3}+\kappa^{-1}\|u\|_{2}^{2})\|u_{t}\|_{0}\right)
+((1+κ1σ2)u2+κ1σt1)ut0ut1,\displaystyle+((1+\kappa^{-1}\|\sigma\|_{2})\|u\|_{2}+\kappa^{-1}\|\sigma_{t}\|_{1})\|u_{t}\|_{0}\|u_{t}\|_{1},

which yields (3.2). This completes the proof. ∎

Next we shall derive the dissipative estimates for Δσt\Delta\sigma_{t} and utu_{t}.

Lemma 3.3.

It holds that

ddt(μ2u02Δσσtdx)+ρut02\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{\mu}{2}\|\nabla u\|^{2}_{0}-\int\Delta\sigma\sigma_{t}\mathrm{d}x\right)+\|\sqrt{\rho}u_{t}\|^{2}_{0}
\displaystyle\lesssim σt02+σt1(σ1,2u1+σ3u1)+(1+κ1σ2)u0u2ut0\displaystyle\|\nabla\sigma_{t}\|_{0}^{2}+\|\sigma_{t}\|_{1}(\|\sigma\|_{1,2}\|u\|_{1}+\|\sigma\|_{3}\|\nabla u\|_{1})+(1+\kappa^{-1}\|\sigma\|_{2})\|\nabla u\|_{0}\|u\|_{2}\|u_{t}\|_{0} (3.23)

and

ddt(μ2Δu02ρΔuutdx)+Δσt02\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{\mu}{2}\|\Delta u\|_{0}^{2}-\int\rho\Delta u\cdot u_{t}\mathrm{d}x\right)+\|\Delta\sigma_{t}\|_{0}^{2}
\displaystyle\lesssim κ1σt0u22u2+(1+κ1σ2)ut1(ut0+u2u2)\displaystyle\kappa^{-1}\|\sigma_{t}\|_{0}\|u\|_{2}^{2}\|\nabla u\|_{2}+(1+\kappa^{-1}\|\sigma\|_{2})\|u_{t}\|_{1}(\|\nabla u_{t}\|_{0}+\|u\|_{2}\|\nabla u\|_{2})
+σt2(σ1,2u2+σ3u2).\displaystyle+\|\sigma_{t}\|_{2}(\|\sigma\|_{1,2}\|u\|_{2}+\|\sigma\|_{3}\|\nabla u\|_{2}). (3.24)
Proof.

(1) Taking the inner product of (2.6)2 and utu_{t} in L2L^{2}, and then using the integration by parts and the boundary condition (2.5), we get

μ2ddtu02+ρut02=(Δσ(κ𝐞3+σ)+ρuu)utdx.\displaystyle\frac{\mu}{2}\frac{\mathrm{d}}{\mathrm{d}t}\|\nabla u\|^{2}_{0}+\|\sqrt{\rho}u_{t}\|^{2}_{0}=-\int(\Delta\sigma(\kappa\mathbf{e}^{3}+\nabla{\sigma})+{\rho}u\cdot\nabla u)\cdot u_{t}\mathrm{d}x. (3.25)

Using (2.6)1\eqref{1.7}_{1} and the boundary condition of σ\sigma in (A.16), we find that

κΔσtu3dx=\displaystyle-\kappa\int\Delta\sigma\partial_{t}u_{3}\mathrm{d}x= Δσt(σt+uσ)dx\displaystyle\int\Delta\sigma\partial_{t}(\sigma_{t}+u\cdot\nabla\sigma)\mathrm{d}x
=\displaystyle= ddtΔσσtdx+σt02+Δσt(uσ)dx.\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\int\Delta\sigma\sigma_{t}\mathrm{d}x+\|\nabla\sigma_{t}\|_{0}^{2}+\int\Delta\sigma\partial_{t}(u\cdot\nabla\sigma)\mathrm{d}x.

Putting the above identity into (3.25) yields

ddt(μ2u02Δσσtdx)+ρut02\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{\mu}{2}\|\nabla u\|^{2}_{0}-\int\Delta\sigma\sigma_{t}\mathrm{d}x\right)+\|\sqrt{\rho}u_{t}\|^{2}_{0}
=\displaystyle= σt02+(Δσuσtρuuut)dx.\displaystyle\|\nabla\sigma_{t}\|_{0}^{2}+\int(\Delta{\sigma}u\cdot\nabla{\sigma_{t}}-{\rho}u\cdot\nabla u\cdot u_{t})\mathrm{d}x. (3.26)

Exploiting Hölder inequality, the incompressible condition, the integration by parts, the product estimate, the boundary condition of u3u_{3} in (2.5), (3.10), Poincaré-type inequality (A.3), we obtain

Δσuσtdx=\displaystyle\int\Delta{\sigma}u\cdot\nabla{\sigma_{t}}\mathrm{d}x= (uhhΔσ+u33Δσ)σtdx\displaystyle-\int(u_{\rm h}\cdot\nabla_{\rm h}\Delta{\sigma}+u_{3}\partial_{3}\Delta{\sigma}){\sigma}_{t}\mathrm{d}x
\displaystyle\lesssim σt1(σ1,2u1+σ3u1).\displaystyle\|\sigma_{t}\|_{1}(\|\sigma\|_{1,2}\|u\|_{1}+\|\sigma\|_{3}\|\nabla u\|_{1}).

Similarly, we also have

ρuuutdx(1+κ1σ2)u0u2ut0.\displaystyle-\int{\rho}u\cdot\nabla u\cdot u_{t}\mathrm{d}x\lesssim(1+\kappa^{-1}\|\sigma\|_{2})\|\nabla u\|_{0}\|u\|_{2}\|u_{t}\|_{0}.

Putting the above two estimates into (3.1) yields (3.3).

(2) Applying Δ\Delta to the mass equation (2.6)1 yields

Δ(σt+uσ+κu3)=0.\Delta(\sigma_{t}+u\cdot\nabla\sigma+{\kappa}u_{3})=0.

Taking the inner product of the above identity and Δσt\Delta\sigma_{t} in L2L^{2}, we have

Δσt02=ΔσtΔ(κu3+uσ)dx.\displaystyle\|\Delta\sigma_{t}\|_{0}^{2}=-\int\Delta\sigma_{t}\Delta(\kappa u_{3}+u\cdot\nabla\sigma)\mathrm{d}x.

Taking the inner product of (3.22)2 and Δu-\Delta u in L2L^{2}, and then using the integration by parts, the boundary conditions of (u3,32u3)(u_{3},\partial_{3}^{2}u_{3}) in (2.5), (3.7), we arrive at

ddt(μ2Δu02ρΔuutdx)\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{\mu}{2}\|\Delta u\|_{0}^{2}-\int\rho\Delta u\cdot u_{t}\mathrm{d}x\right)
=\displaystyle= (t(Δσ(κ𝐞3+σ)+ρuu)ΔuρutΔut)dx.\displaystyle\int(\partial_{t}(\Delta\sigma(\kappa\mathbf{e}^{3}+\nabla\sigma)+\rho u\cdot\nabla u)\cdot\Delta u-\rho u_{t}\cdot\Delta u_{t})\mathrm{d}x.

Summing up the above two identities, we obtain

ddt(μ2Δu02ρΔuutdx)+Δσt02\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{\mu}{2}\|\Delta u\|_{0}^{2}-\int\rho\Delta u\cdot u_{t}\mathrm{d}x\right)+\|\Delta\sigma_{t}\|_{0}^{2}
=\displaystyle= ((Δσσt+Δσtσ+t(ρu)u+ρuut)Δu\displaystyle\int\left((\Delta\sigma\nabla\sigma_{t}+\Delta\sigma_{t}\nabla\sigma+\partial_{t}(\rho u)\cdot\nabla u+\rho u\cdot\nabla u_{t})\cdot\Delta u\right.
ΔσtΔ(uσ)ρutΔut)dx=J3+J4,\displaystyle\left.-\Delta\sigma_{t}\Delta(u\cdot\nabla\sigma)-\rho u_{t}\cdot\Delta u_{t}\right)\mathrm{d}x=J_{3}+J_{4}, (3.27)

where we have defined that

J3:=\displaystyle J_{3}:= ((Δσσt+Δσtσ+(ρut+ρtu)u)Δu\displaystyle\int\Big{((}\Delta\sigma\nabla\sigma_{t}+\Delta\sigma_{t}\nabla\sigma+(\rho u_{t}+\rho_{t}u)\cdot\nabla u)\cdot\Delta u
Δσt(2iuiσ+Δuσ))dx,\displaystyle\ \ \ \ \ \ -\Delta\sigma_{t}(2\partial_{i}u\cdot\nabla\partial_{i}\sigma+\Delta u\cdot\nabla\sigma)\Big{)}\mathrm{d}x,
J4:=\displaystyle J_{4}:= (ρuutΔuρutΔutΔσtuΔσ)dx.\displaystyle\int(\rho u\cdot\nabla u_{t}\cdot\Delta u-\rho u_{t}\cdot\Delta u_{t}-\Delta\sigma_{t}u\cdot\nabla\Delta\sigma)\mathrm{d}x.

Making use of the product estimates, (3.10) and the relation ρt=κ1σt\rho_{t}=\kappa^{-1}\sigma_{t}, it is easy to have

J3(σ2σt2+(κ1σt0u2+(1+κ1σ2)ut0)u2)u2.\displaystyle J_{3}\lesssim(\|\sigma\|_{2}\|\sigma_{t}\|_{2}+(\kappa^{-1}\|\sigma_{t}\|_{0}\|u\|_{2}+(1+\kappa^{-1}\|\sigma\|_{2})\|u_{t}\|_{0})\|u\|_{2})\|\nabla u\|_{2}. (3.28)

Exploiting the integration by parts, the product estimate, the boundary condition (2.5), and Poincaré-type inequality (A.3), we arrive at

J4=\displaystyle J_{4}= (i(ρut)iut+ρuutΔuΔσt(uhhΔσ+u33Δσ))dx\displaystyle\int(\partial_{i}(\rho u_{t})\cdot\partial_{i}u_{t}+\rho u\cdot\nabla u_{t}\cdot\Delta u-\Delta\sigma_{t}(u_{\rm h}\cdot\nabla_{\rm h}\Delta\sigma+u_{3}\partial_{3}\Delta\sigma))\mathrm{d}x
\displaystyle\lesssim (1+κ1σ2)ut1(ut0+u1u2)\displaystyle(1+\kappa^{-1}\|\sigma\|_{2})\|u_{t}\|_{1}(\|\nabla u_{t}\|_{0}+\|\nabla u\|_{1}\|u\|_{2})
+σt2(σ1,2u2+σ3u2).\displaystyle+\|\sigma_{t}\|_{2}(\|\sigma\|_{1,2}\|u\|_{2}+\|\sigma\|_{3}\|\nabla u\|_{2}). (3.29)

Consequently, inserting (3.28) and (3.1) into (3.1) yields (3.3). ∎

Next we shall establish the energy estimate of κ(h3σ,Δhσ)0\kappa\|(\nabla_{\mathrm{h}}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)\|_{0} and the dissipation estimate of σ1,2\|\sigma\|_{1,2}.

Lemma 3.4.

It holds that

κ(Δhσ,h3σ)0\displaystyle\kappa\|(\Delta_{\mathrm{h}}\sigma,\nabla_{\mathrm{h}}\partial_{3}\sigma)\|_{0}
u2+(1+κ1σ2)(ut0+u0u2)+σ2σ3.\displaystyle\lesssim\|u\|_{2}+(1+\kappa^{-1}\|\sigma\|_{2})(\|u_{t}\|_{0}+\|\nabla u\|_{0}\|u\|_{2})+\|\sigma\|_{2}\|\sigma\|_{3}. (3.30)

and

σ1,2\displaystyle\|\sigma\|_{1,2}\lesssim κ1(Δu0+σ1,2σ3)+(κ1+κ2σ2)(ut1+u2u1)\displaystyle\kappa^{-1}(\|\nabla\Delta u\|_{0}+\|\sigma\|_{1,2}\|\sigma\|_{3})+(\kappa^{-1}+\kappa^{-2}\|\sigma\|_{2})\left(\|u_{t}\|_{1}+\|u\|_{2}\|\nabla u\|_{1}\right) (3.31)
Proof.

(1) We can rewrite (2.6)2(\ref{1.7})_{2} as follows

κ(13σ,23σ,Δhσ)=𝒫μΔu+ρ(ut+uu)+σΔσ,\displaystyle\kappa(\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)^{\top}=\nabla\mathcal{P}-\mu\Delta u+\rho(u_{t}+u\cdot\nabla u)+\nabla\sigma\Delta\sigma, (3.32)

where 𝒫=β+κ3σ\mathcal{P}=\beta+\kappa\partial_{3}\sigma. Taking the inner product of the above identity and (13σ,23σ,Δhσ)(\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)^{\top} in L2L^{2}, and then using the integration by parts and the boundary conditions of (σ,3ρ)(\sigma,\partial_{3}\rho) in (A.16), we have

κ|13σ,23σ,Δhσ|2dx\displaystyle\kappa\int|\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma|^{2}{\rm d}x
=\displaystyle= (σΔσμΔu+ρ(ut+uu))(13σ,23σ,Δhσ)dx.\displaystyle\int(\nabla\sigma\Delta\sigma-\mu\Delta u+\rho(u_{t}+u\cdot\nabla u))\cdot(\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)^{\top}{\rm d}x.

We easily deduce from the above identity that

κ(h3σ,Δhσ)0\displaystyle\kappa\|(\nabla_{\mathrm{h}}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)\|_{0}
u2+(1+κ1σ2)(ut0+u0u2)+σ2σ3,\displaystyle\lesssim\|u\|_{2}+(1+\kappa^{-1}\|\sigma\|_{2})(\|u_{t}\|_{0}+\|\nabla u\|_{0}\|u\|_{2})+\|\sigma\|_{2}\|\sigma\|_{3},

which yields (3.30).

(2) Applying curl\mathrm{curl} to the momentum equation (2.6)2, we can obtain the vortex equation

κΔσ=ρ(ωt+uω)μΔω+𝐌+𝐍,\displaystyle\kappa\nabla^{\bot}\Delta\sigma=\rho(\omega_{t}+u\cdot\nabla\omega)-\mu\Delta\omega+{\mathbf{M}}+\mathbf{N}, (3.33)

where we have defined that

{𝐌:=(tu2,tu1,0),𝐍:=𝐍m+𝐍c+𝐍k,𝐍m:=(2ρtu33ρtu2,3ρtu11ρtu3,1ρtu22ρtu1),𝐍c:=(2(ρu)u33(ρu)u2,3(ρu)u11(ρu)u3,1(ρu)u22(ρu)u1),𝐍k:=(3σ2Δσ2σ3Δσ,1σ3Δσ3σ1Δσ,2σ1Δσ1σ2Δσ).\displaystyle\begin{cases}\mathbf{M}:=(-\partial_{t}u_{2},\partial_{t}u_{1},0)^{\top},\ \mathbf{N}:=\mathbf{N}^{\mathrm{m}}+\mathbf{N}^{\mathrm{c}}+\mathbf{N}^{\mathrm{k}},\\ \mathbf{N}^{\mathrm{m}}:=(\partial_{2}\rho\partial_{t}u_{3}-\partial_{3}\rho\partial_{t}u_{2},\partial_{3}\rho\partial_{t}u_{1}-\partial_{1}\rho\partial_{t}u_{3},\partial_{1}\rho\partial_{t}u_{2}-\partial_{2}\rho\partial_{t}u_{1})^{\top},\\ \mathbf{N}^{\mathrm{c}}:=\big{(}\partial_{2}({\rho}u)\cdot\nabla u_{3}-\partial_{3}({\rho}u)\cdot\nabla u_{2},\partial_{3}({\rho}u)\cdot\nabla u_{1}-\partial_{1}({\rho}u)\cdot\nabla u_{3},\\ \qquad\quad\partial_{1}({\rho}u)\cdot\nabla u_{2}-\partial_{2}({\rho}u)\cdot\nabla u_{1}\big{)}^{\top},\\ \mathbf{N}^{\mathrm{k}}:=(\partial_{3}\sigma\partial_{2}\Delta\sigma-\partial_{2}\sigma\partial_{3}\Delta\sigma,\partial_{1}\sigma\partial_{3}\Delta\sigma-\partial_{3}\sigma\partial_{1}\Delta\sigma,\partial_{2}\sigma\partial_{1}\Delta\sigma-\partial_{1}\sigma\partial_{2}\Delta\sigma)^{\top}.\end{cases}

Exploiting the definition of ρ\rho and the product estimates, we have

(𝐌h,𝐍h)0\displaystyle\|({\mathbf{M}}_{\mathrm{h}},{\mathbf{N}}_{\mathrm{h}})\|_{0}\lesssim (1+κ1σ2)(ut1+u1u2)+σ1,2σ3\displaystyle(1+\kappa^{-1}\|\sigma\|_{2})\left(\|u_{t}\|_{1}+\|\nabla u\|_{1}\|u\|_{2}\right)+\|\sigma\|_{1,2}\|\sigma\|_{3} (3.34)

Taking the inner product of the vortex equation (3.33) and Δσ-\nabla^{\bot}\Delta\sigma in L2L^{2}, and then using Hölder inequality, the product estimate, (3.10) and (3.34), we get that

κΔ(2σ,1σ)02\displaystyle\kappa\|\Delta(\partial_{2}\sigma,\partial_{1}\sigma)\|_{0}^{2}
\displaystyle\lesssim (ρ(tωh+uωh)μΔωh+𝐌h+𝐍h)hΔσdx\displaystyle\int(\rho(\partial_{t}\omega_{\mathrm{h}}+u\cdot\nabla\omega_{\mathrm{h}})-\mu{\Delta\omega_{\mathrm{h}}}+{\mathbf{M}}_{\mathrm{h}}+{\mathbf{N}}_{\mathrm{h}})\cdot\nabla_{\mathrm{h}}^{\bot}\Delta\sigma\mathrm{d}x
\displaystyle\lesssim σ1,22σ3+σ1,2(Δu0+(1+κ1σ2)(ut1+u2u1)),\displaystyle\|\sigma\|_{1,2}^{2}\|\sigma\|_{3}+\|\sigma\|_{1,2}\left(\|\nabla\Delta u\|_{0}+(1+\kappa^{-1}\|\sigma\|_{2})(\|u_{t}\|_{1}+\|u\|_{2}\|\nabla u\|_{1})\right),

which, together with (A.22) with s=1s=1 and 22, yields (3.31). This completes the proof. ∎

Finally we shall derive that σt1\|\sigma_{t}\|_{1} and ut0\|u_{t}\|_{0} can be controlled by the norms of spatial derivatives of (σ,u)(\sigma,u).

Lemma 3.5.

It holds that

σt1κu31+σ1u2\displaystyle\|\sigma_{t}\|_{1}\lesssim\kappa\|u_{3}\|_{1}+\|\nabla\sigma\|_{1}\|u\|_{2} (3.35)

and

ut0\displaystyle\|u_{t}\|_{0}\lesssim κ(Δhσ,h3σ)0+(1+(1+κ1σ2)u1)u2+σ1σ2.\displaystyle{\kappa}\|(\Delta_{\mathrm{h}}\sigma,\nabla_{\mathrm{h}}\partial_{3}\sigma)\|_{0}+(1+(1+\kappa^{-1}\|\sigma\|_{2})\|u\|_{1})\|u\|_{2}+\|\nabla\sigma\|_{1}\|\nabla\sigma\|_{2}. (3.36)
Proof.

It is easy see from the product estimate and (2.6)1(\ref{1.7})_{1} that

σt1\displaystyle\|\sigma_{t}\|_{1}\lesssim κu31+uσ1κu31+σ1u2,\displaystyle\kappa\|u_{3}\|_{1}+\|u\cdot\nabla\sigma\|_{1}\lesssim\kappa\|u_{3}\|_{1}+\|\nabla\sigma\|_{1}\|u\|_{2},

which yields (3.35).

It is well-known from the incompressible condition and the mass equation (1.1)1 that

0<dinfxΩ{ρ¯(x)+κ1σ0}ρ(t,x) for any (x,t)Ω×IT,0<d\leqslant\inf\limits_{x\in\Omega}\{\bar{\rho}(x)+\kappa^{-1}\sigma^{0}\}\leqslant{\rho}(t,x)\mbox{ for any }(x,t)\in\Omega\times I_{T}, (3.37)

where ρ=ρ¯+κ1σ\rho=\bar{\rho}+\kappa^{-1}\sigma. Taking the inner product of (3.32) and utu_{t} in L2L^{2}, and then using the integration by parts, and the boundary condition of u3u_{3} in (2.5), we get

ρut02=\displaystyle\|\sqrt{\rho}u_{t}\|^{2}_{0}= ((μΔuσΔσρuu)ut+κ(tuhh3σΔhσtu3))dx.\displaystyle\int((\mu\Delta u-\nabla{\sigma}\Delta{\sigma}-{\rho}u\cdot\nabla u)\cdot u_{t}+\kappa(\partial_{t}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\partial_{3}\sigma-\Delta_{\mathrm{h}}\sigma\partial_{t}u_{3}))\mathrm{d}x.

Exploiting Hölder inequality, the product estimate, (3.10) with i=0i=0 and (3.37), we can get (3.36) from the above identity. This completes the proof. ∎

3.2 A priori stability estimates

Now we are in the position to building the total energy inequality (2.11) for the initial boundary value problem (2.5)–(2.7).

Proposition 3.1 (A priori estimates).

Let (σ,u)(\sigma,u) be the solution of the initial-boundary value problem (2.5)–(2.7) defined on Ω×IT\Omega\times I_{T} with the initial data (σ0,u0)Hρ¯3×s2(\sigma^{0},u^{0})\in H^{3}_{\bar{\rho}}\times{\mathcal{H}^{2}_{\mathrm{s}}} satisfying (2.8). Then there exist constants c1c_{1}, c2c_{2} and χ\chi, where

c1,c21 and χ=max{κ1,c2}c_{1},\ c_{2}\geqslant 1\mbox{ and }\chi=\max\{\kappa^{-1},c_{2}\}

such that, if (σ,u)(\sigma,u) satisfies

sup0t<TE(t)(2c1χ9)2,\displaystyle\mathop{\rm sup}_{0\leqslant t<T}{{E}}(t)\leqslant(2c_{1}\chi^{9})^{-2}, (3.38)

then the solution satisfies the following a priori stability estimate

sup0t<T(t)+0T𝒟(τ)𝑑τ2c1χ9E0.\displaystyle\mathop{\rm sup}_{0\leqslant t<T}\mathcal{E}(t)+\int_{0}^{T}\mathcal{D}(\tau)d\tau\leqslant 2c_{1}\chi^{9}E^{0}. (3.39)
Proof.

Let

sup0t<TE(t)δ21.\displaystyle\mathop{\rm sup}_{0\leqslant t<T}{{E}}(t)\leqslant\delta^{2}\leqslant 1. (3.40)

In view of (3.35) and (3.36), it is easy to see that

(t)(1+κ1σ2)E(t) for any t[0,T),\displaystyle\mathcal{E}(t)\lesssim(1+\kappa^{-1}\|\sigma\|_{2})E(t)\mbox{ for any }t\in[0,T), (3.41)

where δ\delta will be defined by (3.49). In addition, we can derive from (2.6)1(\ref{1.7})_{1} and (3.40) that

κu31σt1+uσ1σt1+σ2u2σ2+σt1.\displaystyle\kappa\|u_{3}\|_{1}\lesssim\|\sigma_{t}\|_{1}+\|u\cdot\nabla\sigma\|_{1}\lesssim\|\sigma_{t}\|_{1}+\|\sigma\|_{2}\|u\|_{2}\lesssim\|\sigma\|_{2}+\|\sigma_{t}\|_{1}. (3.42)

Exploiting Young’s inequality, (3.35), (3.40), (A.20) with i=1i=1 and (A.22) with s=ts=t, we derive from Lemmas 3.13.3 and (3.31) that, for sufficiently large positive constant χ1\chi\geqslant 1 (independent of κ\kappa),

χddt~(t)+c𝒟~(t)(χ3((χ4(1+κ1)+κ2))+κ4)E1/2(t)𝒟(t)\displaystyle\chi\frac{\mathrm{d}}{\mathrm{d}t}\tilde{\mathcal{E}}(t)+c\tilde{\mathcal{D}}(t)\lesssim\left(\chi^{3}\left((\chi^{4}(1+\kappa^{-1})+\kappa^{-2})\right)+\kappa^{-4}\right){E}^{1/2}(t)\mathcal{D}(t)
+κ2(Δu02+ut12),\displaystyle\qquad\qquad\qquad\qquad+\kappa^{-2}(\|\nabla\Delta u\|_{0}^{2}+\|u_{t}\|_{1}^{2}), (3.43)

where we have defined that

~(t):=\displaystyle\tilde{\mathcal{E}}(t):= χ22((σ,Δσ)2+ρ(u,Δu)02)3χ2(κ172κ23σ)3σ(33σ)2dx\displaystyle\frac{\chi^{2}}{2}\left(\|(\nabla\sigma,\nabla\Delta\sigma)\|^{2}+\|\sqrt{\rho}(u,\Delta u)\|^{2}_{0}\right)-3\chi^{2}\int\left(\kappa^{-1}-\frac{7}{2\kappa^{2}}\partial_{3}\sigma\right)\partial_{3}\sigma(\partial_{3}^{3}\sigma)^{2}\mathrm{d}x
+χ3(μ2u02Δσσtdx)+χ4(μ2Δu02ρΔuutdx)+χ6(σt,ρut)02),\displaystyle+\chi^{3}\left(\frac{\mu}{2}\|\nabla u\|^{2}_{0}-\int\Delta\sigma\sigma_{t}\mathrm{d}x\right)+\chi^{4}\left(\frac{\mu}{2}\|\Delta u\|_{0}^{2}-\int\rho\Delta u\cdot u_{t}\mathrm{d}x\right)+\chi^{6}\|(\nabla\sigma_{t},\sqrt{{\rho}}u_{t})\|^{2}_{0}),
𝒟~(t):=\displaystyle\tilde{\mathcal{D}}(t):= σ1,22+χ(χ2u22+χ4σt22+χ3ut12).\displaystyle\|\sigma\|_{1,2}^{2}+\chi(\chi^{2}\|\nabla u\|^{2}_{2}+\chi^{4}\|\sigma_{t}\|_{2}^{2}+\chi^{3}\|u_{t}\|^{2}_{1}).

In addition, making use of Young’s inequality, the definitions of (t)\mathcal{E}(t), 𝒟(t)\mathcal{D}(t), (3.10) with i=0i=0, (3.30), (3.37), (3.40), (3.42), Poincaré-type inequality (A.3), (A.20) with i=0i=0 and (A.21), we have, for sufficiently large positive constant χ\chi,

(t)~(t)+cχ2(χ2κ1+κ2)σ3(t),\displaystyle{\mathcal{E}}(t)\lesssim\tilde{\mathcal{E}}(t)+c\chi^{2}\left(\chi^{2}\kappa^{-1}+\kappa^{-2}\right)\|\sigma\|_{3}{\mathcal{E}}(t), (3.44)
~(t)(χ6(1+κ1)+χ2κ2)(t),\displaystyle\tilde{\mathcal{E}}(t)\lesssim\left(\chi^{6}(1+\kappa^{-1})+\chi^{2}\kappa^{-2}\right){\mathcal{E}}(t), (3.45)
and 𝒟(t)𝒟~(t).\displaystyle\mbox{and }{\mathcal{D}}(t)\lesssim\tilde{\mathcal{D}}(t). (3.46)

It is easy to see from (3.43) and (3.46) that there exists c21c_{2}\geqslant 1 such that the following estimate holds for any χmax{κ1,c2}\chi\geqslant\max\{\kappa^{-1},c_{2}\}:

χddt~(t)+c𝒟(t)χ8E1/2(t)𝒟(t).\displaystyle\chi\frac{\mathrm{d}}{\mathrm{d}t}\tilde{\mathcal{E}}(t)+c{\mathcal{D}}(t)\lesssim\chi^{8}{E}^{1/2}(t)\mathcal{D}(t). (3.47)

Integrating the above inequality over (0,t)(0,t) we obtain

χ~(t)+0t𝒟(τ)dτ\displaystyle\chi\tilde{\mathcal{E}}(t)+\int_{0}^{t}{\mathcal{D}}(\tau)\mathrm{d}\tau\leqslant χ~(t)|t=0+χ8sup0τtE1/2(τ)0t𝒟(τ)dτ,\displaystyle\chi\tilde{\mathcal{E}}(t)|_{t=0}+\chi^{8}\sup_{0\leqslant\tau\leqslant t}{E}^{1/2}(\tau)\int_{0}^{t}\mathcal{D}(\tau)\mathrm{d}\tau,

which, together with (3.41), (3.44) and (3.45), implies

(t)+0t𝒟(τ)dτ\displaystyle{\mathcal{E}}(t)+\int_{0}^{t}{\mathcal{D}}(\tau)\mathrm{d}\tau\leqslant c1χ9(E0+σ3(t)+sup0τtE1/2(τ)0t𝒟(τ)dτ).\displaystyle c_{1}\chi^{9}\left(E^{0}+\|\sigma\|_{3}\mathcal{E}(t)+\sup_{0\leqslant\tau\leqslant t}{E}^{1/2}(\tau)\int_{0}^{t}\mathcal{D}(\tau)\mathrm{d}\tau\right). (3.48)

If

δ:=(2c1χ9)1<1,\displaystyle\delta:=(2c_{1}\chi^{9})^{-1}<1, (3.49)

we can derive from (3.40) and (3.48) that

(t)+0t𝒟(τ)dτ2c1χ9E0,\displaystyle{\mathcal{E}}(t)+\int_{0}^{t}{\mathcal{D}}(\tau)\mathrm{d}\tau\leqslant 2c_{1}\chi^{9}E^{0}, (3.50)

which yields (3.39). This completes the proof. ∎

4 Proof of Theorem 2.1

Now we introduce the local(-in-time) well-posedness of the initial-boundary value problem (2.5)–(2.7) for any fixed κ\kappa.

Proposition 4.1.

Let μ\mu, κ\kappa be given positive constants, ρ¯\bar{\rho} be defined by (2.1), and (σ0,u0)Hρ¯3×s2(\sigma^{0},u^{0})\in H^{3}_{\bar{\rho}}\times{\mathcal{H}^{2}_{\mathrm{s}}} satisfy (2.8) . Then there exists Tmax>0T^{\max}>0 such that the initial-boundary value problem (2.5)–(2.7) admits a unique local(-in-time) strong solution (ϱ,v)(\varrho,v) defined on Ω×[0,Tmax)\Omega\times[0,T^{\max}) with an associated pressure β\beta. Moreover

  • 1.

    (σ,v,β)𝔓T×𝒱T×C0(IT¯,L2)(\sigma,v,\nabla\beta)\in{\mathfrak{P}}_{T}\times{\mathcal{V}_{T}}\times C^{0}(\overline{I_{T}},L^{2}) for any TITmaxT\in I_{T^{\max}}, and

    0<infxΩ{ρ0(x)}ρ(t,x)supxΩ{ρ0(x)} for any (t,x)Ω×ITmax,0<\inf\limits_{x\in\Omega}\big{\{}{\rho}^{0}(x)\big{\}}\leqslant{\rho}(t,x)\leqslant\sup\limits_{x\in\Omega}\big{\{}{\rho}^{0}(x)\big{\}}\mbox{ for any }(t,x)\in\Omega\times I_{T^{\max}},

    where ρ0:=ϱ0+κ1ρ¯\rho^{0}:=\varrho^{0}+\kappa^{-1}\bar{\rho}.

  • 2.

    lim suptTmax(σ,v)(t)2=\limsup_{t\to T^{\max}}\|(\nabla\sigma,v)(t)\|_{2}=\infty if Tmax<T^{\max}<\infty.

Proof 1.

Since Proposition 4.1 can be easily proved by the standard iteration method as in [58, Theorem 1], we omit the trivial proof. \Box

Thanks to the a priori stability estimate (3.39) in Proposition 3.1, we can easily establish the global solvability in Theorem 2.1 based on the local solvability in Proposition 4.1. Next, we briefly describe the proof for the readers’ convenience.

Assume that (σ0,v0)(\sigma^{0},v^{0}) satisfies (2.8) and (2.10), where c1c_{1} and c2c_{2} are provided by Proposition 3.1. In view of Proposition 4.1, there exists a unique local strong solution (ϱ,v,β)(\varrho,v,\beta) to the initial-boundary value problem (2.5)–(2.7) with the maximal existence time TmaxT^{\max}.

By the regularity of (ϱ,v,β)(\varrho,v,\beta), we can verify that (ϱ,v)(\varrho,v) satisfies stability estimate (3.39) in Proposition 3.1, i.e.

sup0t<T(t)+0T𝒟(t)dt2c1χ9E0,\displaystyle\sup_{0\leqslant t<T}{\mathcal{E}(t)}+\int_{0}^{T}\mathcal{D}(t)\mathrm{d}t\leqslant 2c_{1}\chi^{9}E^{0}, (4.1)

if

sup0t<TE(t)(2c1χ9)2 for TITmax.\sup_{0\leqslant t<T}E(t)\leqslant(2c_{1}\chi^{9})^{-2}\mbox{ for }T\in I_{T^{\max}}.

Let

T=\displaystyle T^{*}= sup{τITmax|E(t)(3c1χ9)2 for any tτ}.\displaystyle\sup\left\{\tau\in I_{T^{\max}}~{}\left|~{}E(t)\leqslant(3c_{1}\chi^{9})^{-2}\mbox{ for any }\ t\leqslant\tau\right.\right\}.

Then, we easily see that the definition of TT^{*} makes sense by the fact

E0(3c1χ9)3<(2c1χ9)2.\displaystyle E^{0}\leqslant(3c_{1}\chi^{9})^{-3}<(2c_{1}\chi^{9})^{-2}. (4.2)

Thus, to show the existence of a global strong solution, it suffices to verify T=T^{*}=\infty. We shall prove this by contradiction below.

Assume T<T^{*}<\infty, then by and the definition of TT^{*} and Proposition 4.1, we have

TITmax.\displaystyle T^{*}\in I_{T^{\max}}. (4.3)

Noting that

sup0t<TE(t)(2c1χ9)2,\sup_{0\leqslant t<{T^{*}}}E(t)\leqslant(2c_{1}\chi^{9})^{-2},

then, by (4.1) with T=TT=T^{*} and (4.2), we have

sup0t<T(t)+0T𝒟(t)dt2c1χ9E0227(c1χ9)2.\displaystyle\sup_{0\leqslant t<T^{*}}{\mathcal{E}(t)}+\int_{0}^{T^{*}}\mathcal{D}(t)\mathrm{d}t\leqslant 2c_{1}\chi^{9}E^{0}\leqslant\frac{2}{27(c_{1}\chi^{9})^{2}}.

In particular,

sup0t<TE(t)827(2c1χ9)2<(2c1χ9)2.\displaystyle\sup_{0\leqslant t<{T^{*}}}E(t)\leqslant\frac{8}{27(2c_{1}\chi^{9})^{2}}<(2c_{1}\chi^{9})^{-2}. (4.4)

Making use of (4.3), (4.4) and the strong continuity (ϱ,v)C0([0,Tmax),H3×H2)(\varrho,v)\in C^{0}([0,T^{\max}),H^{3}\times H^{2}), we deduce that there is a constant T~(T,Tmax)\tilde{T}\in(T^{*},T^{\max}), such that

sup0tT~E(t)(2c1χ9)2,\displaystyle\sup_{0\leqslant t\leqslant{\tilde{T}}}E(t)\leqslant(2c_{1}\chi^{9})^{-2},

which contradicts with the definition of TT^{*}. Hence, T=T^{*}=\infty, which implies Tmax=T^{\max}=\infty. This completes the proof of the existence of a global solution. The uniqueness of the global solution is obvious due to the uniqueness result of local solutions in Proposition 4.1. In addition, exploiting the product estimates and (2.11), we can easily derive (2.12) from (3.32).

To complete the proof of Theorem 2.1, we shall verify (2.13). To this purpose, applying curl\mathrm{curl} to the momentum equation (3.32)1 yields

κcurl(13σ,23σ,Δhσ)=curl(ρ(ut+uu)μΔu+σΔσ).\displaystyle\kappa\mathrm{curl}(\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)^{\top}=\mathrm{curl}(\rho(u_{t}+u\cdot\nabla u)-\mu\Delta u+\nabla\sigma\Delta\sigma). (4.5)

Applying L2(IT,L2)\|\cdot\|_{L^{2}(I_{T},L^{2})} to the above identity, and then using the product estimate and (2.11), we can estimate that

κcurl(13σ,23σ,Δhσ)L2(IT,L2)\displaystyle\|\kappa\mathrm{curl}(\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)^{\top}\|_{L^{2}(I_{T},L^{2})}
curl(ρ(ut+uu)μΔu+σΔσ)L2(IT,L2)\displaystyle\lesssim\|\mathrm{curl}(\rho(u_{t}+u\cdot\nabla u)-\mu\Delta u+\nabla\sigma\Delta\sigma)\|_{L^{2}(I_{T},L^{2})}
c1χ9E0(1+χ+T).\displaystyle\lesssim\sqrt{c_{1}\chi^{9}E^{0}}(1+\chi+\sqrt{T}). (4.6)

Noting that div(13σ,23σ,Δhσ)=0\mathrm{div}(\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)^{\top}=0, we can derive from (2.11), (3.30), (4.6) and Lemma A.5 that

κ(13σ,23σ,Δhσ)L2(IT,H1)c1χ9E0(1+χ+T).\displaystyle\|\kappa(\partial_{1}\partial_{3}\sigma,\partial_{2}\partial_{3}\sigma,-\Delta_{\mathrm{h}}\sigma)\|_{L^{2}(I_{T},H^{1})}\lesssim\sqrt{c_{1}\chi^{9}E^{0}}(1+\chi+\sqrt{T}). (4.7)

Thanks to (2.11) and (4.7), we easily deduce from (3.32) that

𝒫L2(IT,H1)c1χ9E0(1+χ+T),\displaystyle\|\nabla\mathcal{P}\|_{L^{2}(I_{T},H^{1})}\lesssim\sqrt{c_{1}\chi^{9}E^{0}}(1+\chi+\sqrt{T}),

which, together with (4.7), yields (2.13). This completes the proof of Theorem 2.1.

5 Proof of Theorem 2.2

This section is devoted to the proof of the asymptotic behavior of solutions stated in Theorem 2.2. For the simplicity, we denote (ϱκ,σκ,uκ)(\varrho^{\kappa},\sigma^{\kappa},u^{\kappa}) by (ϱ,σ,u)(\varrho,\sigma,u), where ϱκ=κ1σκ\varrho^{\kappa}=\kappa^{-1}\sigma^{\kappa}.

Let T>0T>0 and p1p\geqslant 1 be arbitrary given. Making use of (2.11)–(2.13), Aubin–Lions theorem (see Theorem A.3), Arzelá–Ascoli theorem (see Theorem A.4) and Banach–Alaoglu theorem (see Theorem A.5), there exits a sequence (not relabeled) such that, for κ\kappa\to\infty,

uu~ weakly* in L(IT,s2) and weakly in L2(IT,H3) with u~3=0,\displaystyle u\rightarrow\tilde{u}\mbox{ weakly* in }L^{\infty}(I_{T},{\mathcal{H}^{2}_{\mathrm{s}}})\mbox{ and weakly in }L^{2}(I_{T},H^{3})\mbox{ with }\tilde{u}_{3}=0,
uu~ strongly in Lp(IT,Hloc2)C0(IT¯,Hloc1) with u~h|t=0=wh0,\displaystyle u\rightarrow{\tilde{u}}\mbox{ strongly in }L^{p}(I_{T},H^{2}_{\mathrm{loc}})\cap C^{0}(\overline{I_{T}},H^{1}_{\mathrm{loc}})\mbox{ with }\tilde{u}_{\mathrm{h}}|_{t=0}=w^{0}_{\mathrm{h}}, (5.1)
σϖ~ weakly* in L(IT,H03) and  strongly in C0(IT¯,Hloc2),\displaystyle\sigma\rightarrow\tilde{\varpi}\mbox{ weakly* in }L^{\infty}(I_{T},H^{3}_{0})\mbox{ and }\mbox{ strongly in }C^{0}(\overline{I_{T}},H^{2}_{\mathrm{loc}}), (5.2)
utu~t weakly* in L(IT,L2) and weakly in L2(IT,1),\displaystyle u_{t}\rightarrow\tilde{u}_{t}\mbox{ weakly* in }L^{\infty}(I_{T},L^{2})\mbox{ and weakly in }L^{2}(I_{T},\mathcal{H}^{1}), (5.3)
(ϱ,ϱt)=κ1(σ,σt)(0,0) strongly in L(IT,H3)×L2(IT,H2),\displaystyle(\varrho,\varrho_{t})=\kappa^{-1}(\sigma,\sigma_{t})\rightarrow(0,0)\mbox{ strongly in }L^{\infty}(I_{T},H^{3})\times L^{2}(I_{T},{{H}^{2}}),
κ(h3σ,Δhσ)𝒩 weakly* in L(IT,L2) and weakly in L2(IT,H01),\displaystyle\kappa(\nabla_{\mathrm{h}}\partial_{3}\sigma^{\top},-\Delta_{\rm h}\sigma)^{\top}\rightarrow\mathcal{N}\mbox{ weakly* in }L^{\infty}(I_{T},L^{2})\mbox{ and weakly in }L^{2}(I_{T},H^{1}_{0}), (5.4)
(h3σ,Δhσ)0 strongly in L(IT,L2),\displaystyle(\nabla_{\mathrm{h}}\partial_{3}\sigma,-\Delta_{\rm h}\sigma)^{\top}\rightarrow 0\mbox{ strongly in }L^{\infty}(I_{T},L^{2}), (5.5)
𝒫N~ weakly* in L(IT,L2) and weakly in L2(IT,H1);\displaystyle\nabla\mathcal{P}\rightarrow\tilde{N}\mbox{ weakly* in }L^{\infty}(I_{T},L^{2})\mbox{ and weakly in }L^{2}(I_{T},H^{1}); (5.6)

moreover, for any χC0(IT)\chi\in C^{\infty}_{0}(I_{T}), for any ηC0(0,h)\eta\in C^{\infty}_{0}(0,h), for any ϕC0(Ω)\phi\in C^{\infty}_{0}(\Omega), for any (ψ1,ψ2)C0(Ω)(\psi_{1},\psi_{2})^{\top}\in C^{\infty}_{0}(\Omega) satisfying 1ψ1+2ψ2=0\partial_{1}\psi_{1}+\partial_{2}\psi_{2}=0 and for any φC0(Ω)\varphi\in C^{\infty}_{0}(\Omega) satisfying divφ=0\mathrm{div}\varphi=0,

0T(𝒩h,𝒩3)ϕdxχdτ=0,\displaystyle\int_{0}^{T}\int(\mathcal{N}_{\rm h}^{\top},\mathcal{N}_{3})^{\top}\cdot\nabla\phi\mathrm{d}x\chi\mathrm{d}\tau=0, (5.7)
0T0h2(𝒩1ψ1+𝒩2ψ2)dxhηdx3χdτ=0,\displaystyle\int_{0}^{T}\int_{0}^{h}\int_{\mathbb{R}^{2}}(\mathcal{N}_{1}\psi_{1}+\mathcal{N}_{2}\psi_{2})\mathrm{d}x_{\mathrm{h}}\eta\mathrm{d}x_{3}\chi\mathrm{d}\tau=0, (5.8)
0TN~φdxχdτ=0.\displaystyle\int_{0}^{T}\int\tilde{N}\cdot\varphi\mathrm{d}x\chi\mathrm{d}\tau=0. (5.9)

We can further derive from (5.7)–(5.9) that, for a.e. tITt\in I_{T},

div(𝒩h,𝒩3)=0,\displaystyle\mathrm{div}(\mathcal{N}_{\rm h}^{\top},\mathcal{N}_{3})^{\top}=0, (5.10)
2𝒩hψdxh=0 for a.e. x3(0,h),\displaystyle\int_{\mathbb{R}^{2}}\mathcal{N}_{\mathrm{h}}\cdot\psi\mathrm{d}x_{\mathrm{h}}=0\mbox{ for a.e. }x_{3}\in(0,h), (5.11)
N~φdx=0.\displaystyle\int\tilde{N}\cdot\varphi\mathrm{d}x=0. (5.12)

In view of (5.11), (5.12) and Lemma A.6, we have

N~=𝒬 for some 𝒬Lloc2,\displaystyle\tilde{N}=\nabla\mathcal{Q}\mbox{ for some }\mathcal{Q}\in L^{2}_{\mathrm{loc}}, (5.13)
𝒩h=h for some Lloc2(2) for a.e. x3(0,h),\displaystyle\mathcal{N}_{\mathrm{h}}=\nabla_{\mathrm{h}}\mathcal{M}\mbox{ for some }\mathcal{M}\in L^{2}_{\mathrm{loc}}(\mathbb{R}^{2})\mbox{ for a.e. }x_{3}\in(0,h), (5.14)

which, together with (5.10), yields

Δh=3𝒩3.\displaystyle-\Delta_{\mathrm{h}}\mathcal{M}=\partial_{3}\mathcal{N}_{3}. (5.15)

In addition, it is easy to see from (5.2) and (5.5) that

Δhϖ~=0,\displaystyle\Delta_{\mathrm{h}}\tilde{\varpi}=0, (5.16)

which, together with the regularity of ϖ~L(IT,H3)\tilde{\varpi}\in L^{\infty}(I_{T},H^{3}), imply

ϖ~=0.\displaystyle\tilde{\varpi}=0. (5.17)

By virtue of the product estimate and (2.11), it is easily to see that

σutL(IT,L2)+σuuL(IT,H1)1.\|\sigma u_{t}\|_{L^{\infty}(I_{T},L^{2})}+\|\sigma u\cdot\nabla u\|_{L^{\infty}(I_{T},H^{1})}\lesssim 1.

Thus it holds that

κ1σut0 strongly in L(IT,L2),\displaystyle\kappa^{-1}\sigma u_{t}\to 0\mbox{ strongly in }L^{\infty}(I_{T},L^{2}), (5.18)
κ1σuu0 stronly in L(IT,H1).\displaystyle\kappa^{-1}\sigma u\cdot\nabla u\to 0\mbox{ stronly in }L^{\infty}(I_{T},H^{1}). (5.19)

Making use of (5.1), (5.2), (5.3), (5.17)–(5.19) and (A.2), one has

ρutρ¯u~t weakly in L(IT,L2),\displaystyle{\rho}u_{t}\to\bar{\rho}\tilde{u}_{t}\mbox{ weakly in }L^{\infty}(I_{T},L^{2}), (5.20)
ρuuρ¯u~u~ strongly in Lp(IT,Lloc3),\displaystyle{\rho}{u}\cdot\nabla{u}\to\bar{\rho}\tilde{u}\cdot\nabla\tilde{u}\mbox{ strongly in }L^{p}(I_{T},L^{3}_{\mathrm{loc}}), (5.21)
σΔσ0 strongly in C0(IT¯,Lloc3/2).\displaystyle\nabla\sigma\Delta\sigma\to 0\mbox{ strongly in }C^{0}(\overline{I_{T}},L^{3/2}_{\mathrm{loc}}). (5.22)

Let β~=𝒬\tilde{\beta}=\mathcal{Q}-\mathcal{M}. Making use of (5.1), (5.4), (5.6), (5.13), (5.14) and (5.20) –(5.22), we easily derive from (2.6) with (3σ(h3σ,Δhσ))(\nabla\partial_{3}\sigma-(\nabla_{\mathrm{h}}\partial_{3}\sigma^{\top},-\Delta_{\mathrm{h}}\sigma))^{\top} in place of Δσ𝐞3\Delta\sigma\mathbf{e}^{3} that

{ρ¯(tu~h+u~hu~h)+hβ~μΔu~h=0,divhu~h=0\displaystyle\begin{cases}\bar{\rho}(\partial_{t}\tilde{u}_{\mathrm{h}}+\tilde{u}_{\mathrm{h}}\cdot\nabla\tilde{u}_{\mathrm{h}})+\nabla_{\mathrm{h}}\tilde{\beta}-\mu\Delta\tilde{u}_{\mathrm{h}}=0,\\ \mathrm{div}_{\mathrm{h}}\tilde{u}_{\mathrm{h}}=0\end{cases} (5.23)

and

𝒩3=3𝒬,\displaystyle\mathcal{N}_{3}=\partial_{3}\mathcal{Q}, (5.24)

which, together with (5.15), yields

Δh=32𝒬.\displaystyle-\Delta_{\mathrm{h}}\mathcal{M}=\partial_{3}^{2}\mathcal{Q}. (5.25)

In addition, applying divh\mathrm{div}_{\mathrm{h}} to (5.23)1, and then using (5.23)2, we arrive at

Δ𝒬=ρ¯u~h:huh.\displaystyle-\Delta\mathcal{Q}=\bar{\rho}\nabla{\tilde{u}}_{\mathrm{h}}:\nabla_{\mathrm{h}}u_{\mathrm{h}}. (5.26)

Now let u¯h\bar{u}_{\mathrm{h}}, 𝒬¯\bar{\mathcal{Q}}, ¯\bar{\mathcal{M}} and 𝒩¯3\bar{\mathcal{N}}_{3} enjoy the same regularity as well as u~h\tilde{u}_{\mathrm{h}}, 𝒬{\mathcal{Q}}, {\mathcal{M}} and 𝒩3{\mathcal{N}}_{3}, and satisfy the following initial-boundary value problem

{ρ¯(tu¯h+u¯hhu¯h)+h(𝒬¯¯)μΔu¯h=0,divhu¯h=0,u¯h|t=0=wh0,Δ𝒬¯=ρ¯u¯h:hu¯h,3𝒬¯|Ω=0,Δh¯=32𝒬¯,𝒩¯3=3𝒬¯,\displaystyle\begin{cases}\bar{\rho}(\partial_{t}{\bar{u}}_{\mathrm{h}}+{\bar{u}}_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}{\bar{u}}_{\mathrm{h}})+\nabla_{\rm h}(\bar{\mathcal{Q}}-\bar{\mathcal{M}})-\mu\Delta{\bar{u}}_{\mathrm{h}}=0,\\ \mathrm{div}_{\mathrm{h}}{\bar{u}}_{\mathrm{h}}=0,\\ \bar{u}_{\mathrm{h}}|_{t=0}=w_{\mathrm{h}}^{0},\\ -\Delta\bar{\mathcal{Q}}=\bar{\rho}\nabla{\bar{u}}_{\mathrm{h}}:\nabla_{\mathrm{h}}\bar{u}_{\mathrm{h}},\ \partial_{3}\bar{\mathcal{Q}}|_{\partial\Omega}=0,\\ -\Delta_{\mathrm{h}}\bar{\mathcal{M}}=\partial_{3}^{2}\bar{\mathcal{Q}},\ \bar{\mathcal{N}}_{3}=\partial_{3}\bar{\mathcal{Q}},\end{cases} (5.27)

then it is easy to check that

(u¯h,𝒬¯,h¯,𝒩¯3)=(u~h,𝒬,h,𝒩3).(\bar{u}_{\mathrm{h}},\nabla\bar{\mathcal{Q}},\nabla_{\mathrm{h}}\bar{\mathcal{M}},\bar{\mathcal{N}}_{3})=(\tilde{u}_{\mathrm{h}},\nabla{\mathcal{Q}},\nabla_{\mathrm{h}}{\mathcal{M}},{\mathcal{N}}_{3}).

The uniqueness mentioned above means that any sequence of

{(u,𝒫,κ13σ,κ23σ,κΔhσ)}κ>0\{(u,\nabla\mathcal{P},\kappa\partial_{1}\partial_{3}\sigma,\kappa\partial_{2}\partial_{3}\sigma,-\kappa\Delta_{\rm h}\sigma)\}_{\kappa>0}

converges to the limit function (u~,𝒬,1,2,𝒩3)({\tilde{u}},\nabla\mathcal{Q},\partial_{1}\mathcal{M},\partial_{2}\mathcal{M},\mathcal{N}_{3}) is independent of choosing the sequences of solutions. This completes the proof of Theorem 2.2.

Appendix A Analysis tools

This appendix is devoted to providing some mathematical results, which have been used in previous sections. In addition, aba\lesssim b still denotes acba\leqslant cb where the positive constant cc depends on the parameters and the domain in the lemmas in which cc appears.

Lemma A.1.

Embedding inequality [1, Theorem 4.12]: Let D3D\subset\mathbb{R}^{3} be a domain satisfying the cone condition. It holds that

fC0(D¯)fH2(D)\displaystyle\|f\|_{C^{0}(\overline{D})}\lesssim\|f\|_{H^{2}(D)} (A.1)

for any fH2(D)f\in H^{2}(D) (after possibly being redefined on a set of measure zero), and

ϕL6(D)ϕH1(D) for any ϕH1(D).\displaystyle\|\phi\|_{L^{6}(D)}\lesssim\|\phi\|_{H^{1}(D)}\mbox{ for any }\phi\in H^{1}(D). (A.2)
Lemma A.2.

Poincaré-type inequality [23, Lemma A.10]: For any fH01f\in H_{0}^{1}, it holds that

f03f0.\displaystyle\|f\|_{0}\lesssim\|\partial_{3}f\|_{0}. (A.3)
Lemma A.3.
  1. 1.

    Interpolation inequality in HjH^{j}: Let D3D\subset\mathbb{R}^{3} be a domain satisfying the cone condition and 0j<i0\leqslant j<i, then it holds that

    fjf01jifijiεjijf0+εfi\displaystyle\|f\|_{j}\lesssim\|f\|_{0}^{1-\frac{j}{i}}\|f\|_{i}^{\frac{j}{i}}\lesssim\varepsilon^{-\frac{j}{i-j}}\|f\|_{0}+\varepsilon\|f\|_{i} (A.4)

    for any ε>0\varepsilon>0 and for any fHj(D)f\in H^{j}(D).

  2. 2.

    Interpolation inequalities in LpL^{p}: it holds that

    fL42hf0f1 for any fH1,\displaystyle\|f\|_{L^{4}}^{2}\lesssim\|\nabla_{\mathrm{h}}f\|_{0}\|f\|_{1}\mbox{ for any }f\in H^{1}, (A.5)
    fL2hf1f2 for any fH2.\displaystyle\|f\|_{L^{\infty}}^{2}\lesssim\|\nabla_{\mathrm{h}}f\|_{1}\|f\|_{2}\mbox{ for any }f\in H^{2}. (A.6)
Proof 2.

The result (A.4) can be founded in [1, Theorem 5.2]. Next we derive (A.5) and (A.6).

It is well-known that [52]

f(xh,x3)L4(2)2f(xh,x3)L2(2)hf(xh,x3)L2(2) for a.e. x3Ih,\displaystyle\|f(x_{\mathrm{h}},x_{3})\|_{L^{4}(\mathbb{R}^{2})}^{2}\lesssim\|f(x_{\mathrm{h}},x_{3})\|_{L^{2}(\mathbb{R}^{2})}\|\nabla_{\mathrm{h}}f(x_{\mathrm{h}},x_{3})\|_{L^{2}(\mathbb{R}^{2})}\mbox{ for a.e. }x_{3}\in I_{h}, (A.7)
f(xh,x3)L(Ih)2f(xh,x3)L2(Ih)f(xh,x3)W1,2(Ih) for a.e. xh2\displaystyle\|f(x_{\mathrm{h}},x_{3})\|_{L^{\infty}(I_{h})}^{2}\lesssim\|f(x_{\mathrm{h}},x_{3})\|_{L^{2}(I_{h})}\|f(x_{\mathrm{h}},x_{3})\|_{W^{1,2}(I_{h})}\mbox{ for a.e. }x_{\mathrm{h}}\in\mathbb{R}^{2} (A.8)

and

fL4fL4fL43.\displaystyle\|f\|_{L^{\infty}}^{4}\lesssim\|f\|_{L^{4}}\|\nabla f\|_{L^{4}}^{3}. (A.9)

Exploiting (A.7) and (A.8), we have

fL44\displaystyle\|f\|_{L^{4}}^{4}\lesssim f(xh,x3)L2(2)2hf(xh,x3)L2(2)2L1(Ih)\displaystyle\|\|f(x_{\mathrm{h}},x_{3})\|_{L^{2}(\mathbb{R}^{2})}^{2}\|\nabla_{\mathrm{h}}f(x_{\mathrm{h}},x_{3})\|_{L^{2}(\mathbb{R}^{2})}^{2}\|_{L^{1}(I_{h})}
\displaystyle\lesssim f(xh,x3)L(Ih)L2(2)2hf02f12hf02.\displaystyle\|\|f(x_{\mathrm{h}},x_{3})\|_{L^{\infty}(I_{h})}\|_{L^{2}(\mathbb{R}^{2})}^{2}\|\nabla_{\mathrm{h}}f\|_{0}^{2}\lesssim\|f\|_{1}^{2}\|\nabla_{\mathrm{h}}f\|_{0}^{2}. (A.10)

Hence (A.5) holds. Thus we further deduce from (A.5) and (A.9) that

fL2\displaystyle\|f\|_{L^{\infty}}^{2}\lesssim fW1,42hf1f2,\displaystyle\|f\|_{W^{1,4}}^{2}\lesssim\|\nabla_{\mathrm{h}}f\|_{1}\|f\|_{2}, (A.11)

which yields (A.6). \Box

Lemma A.4.

Product estimates (see [31, Lemma A.3]): Let D3D\subset\mathbb{R}^{3} be a domain satisfying the cone condition, and ff, gg be functions defined in DD. Then

fgi{f1g1for i=0;fig2for 0i2.\displaystyle\|fg\|_{i}\lesssim\begin{cases}\|f\|_{1}\|g\|_{1}&\mbox{for }i=0;\\ \|f\|_{i}\|g\|_{2}&\mbox{for }0\leqslant i\leqslant 2.\end{cases} (A.12)

if the norms on the right hand side of the above inequalities are finite.

Lemma A.5.

A Hodge-type elliptic estimate [32, Lemma A.4]: Let i1i\geqslant 1, then

wi1(divw,curlw)i1 for any wHi with w3|Ω=0.\displaystyle\|\nabla w\|_{i-1}\lesssim\|(\mathrm{div}w,\mathrm{curl}w)\|_{i-1}\mbox{ for any }w\in H^{i}\mbox{ with }w_{3}|_{\partial\Omega}=0. (A.13)
Lemma A.6.

Helmholtz decomposition in L2L^{2}-spaces (see Lemma 2.5.1 in Chapter II in [56]): Let DnD\subseteq\mathbb{R}^{n}, n2n\geqslant 2, be any domain. We define that

2(D):=𝒞0(D)¯L2(D),𝒞0(D):={wC0(D)|divw=0}\mathcal{L}^{2}(D):=\overline{\mathcal{C}_{0}^{\infty}(D)}^{\|\cdot\|_{L^{2}(D)}},\ \mathcal{C}_{0}^{\infty}(D):=\{w\in C_{0}^{\infty}(D)~{}|~{}\mathrm{div}w=0\}

and

G(D):={fL2(D)|f=g for some scalar function gLloc2(D)}.G(D):=\{f\in L^{2}(D)~{}|~{}f=\nabla g\mbox{ for some scalar function }g\in L^{2}_{\mathrm{loc}}(D)\}.

Then

G(D)={fL2(D)|Dfwdx=0 for all w2(D)},G(D)=\left\{f\in L^{2}(D)~{}\bigg{|}~{}\int_{D}f\cdot w\mathrm{d}x=0\mbox{ for all }w\in\mathcal{L}^{2}(D)\right\},

and each fL2(D)f\in L^{2}(D) has a unique decomposition

f=f~+gf=\tilde{f}+\nabla g

with f~2(D)\tilde{f}\in\mathcal{L}^{2}(D), gG(D)\nabla g\in G(D), Df~gdx=0\int_{D}\tilde{f}\cdot\nabla g\mathrm{d}x=0 and

f02=f~02+g02.\|f\|_{0}^{2}=\|\tilde{f}\|_{0}^{2}+\|\nabla g\|_{0}^{2}.
Lemma A.7.

An elliptic estimate for the Dirichlet boundary value condition: Let i0i\geqslant 0, f1Hi{f}^{1}\in H^{i} and f2Hi+2{f}^{2}\in H^{i+2} be given, then there exists a unique solution wHi+2w\in H^{i+2} solving the boundary value problem:

{Δw=f1in Ω,w=f2on Ω.\begin{cases}\Delta w={f}^{1}&\mbox{in }\Omega,\\ w={f}^{2}&\mbox{on }\partial\Omega.\end{cases}

Moreover,

wi+2f1i+f2i+2.\|w\|_{i+2}\lesssim\|{f}^{1}\|_{i}+\|{f}^{2}\|_{i+2}. (A.14)
Proof 3.

Please refer to [32, Lemma A.7] for the proof. \Box

Lemma A.8.

An elliptic estimate for the Neumann boundary value condition: Let aa be a positive constant and i0i\geqslant 0, then there exists a unique solution wHi+2w\in H^{i+2} solving the boundary value problem:

{aΔw=divfin Ω,𝐧w=0on Ω,\begin{cases}-a\Delta w=\mathrm{div}{f}&\mbox{in }\Omega,\\ \partial_{{\mathbf{n}}}w=0&\mbox{on }\partial\Omega,\end{cases}

where 𝐧{\mathbf{n}} denotes the outward unit normal vector to Ω\partial\Omega. Moreover,

wi+1f0+divfi.\|\nabla w\|_{i+1}\lesssim\|{f}\|_{0}+\|\mathrm{div}{f}\|_{i}. (A.15)
Proof 4.

Please refer to [32, Lemma A.7] and [56, 1.3.1 Theorem in Chapter III]. \Box

Lemma A.9.

Let (σ,u)𝔓T×𝒱T(\sigma,u)\in{\mathfrak{P}}_{T}\times{\mathcal{V}}_{T} be the solution of (2.6) with the initial condition σ0Hρ¯3\sigma^{0}\in{{H}}^{3}_{\bar{\rho}}, then σ\sigma satisfies the following boundary condition:

(σ,3ρ,32σ)|Ω=0,\displaystyle(\sigma,\partial_{3}\rho,\partial_{3}^{2}\sigma)|_{\partial\Omega}=0, (A.16)

where ρ=ρ¯+κ1σ\rho=\bar{\rho}+\kappa^{-1}\sigma.

Proof.

The above result (A.16) can be found in [35, Lemma 2.1]. However we provide the proofs for reader’s convenience. In view of the mass equation (2.6)1 and the boundary condition of u3u_{3} in (2.5), it holds that

(σt+uhhσ)|Ω=0.\displaystyle(\sigma_{t}+u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\sigma)|_{\partial\Omega}=0.

Taking the inner product of the above identity and σ\sigma in L2(Ω)L^{2}{(\partial\Omega)}, and then using the integration by parts and the embedding inequality of H2C0(Ω¯)H^{2}\hookrightarrow C^{0}(\overline{\Omega}) in (A.1), we derive that

ddtΩ|σ|2dxh=Ωuhh|σ|2dxhdivhuh2Ω|σ|2dxh.\frac{\mathrm{d}}{\mathrm{d}t}\int_{\partial\Omega}|\sigma|^{2}\mathrm{d}x_{\mathrm{h}}=-\int_{\partial\Omega}u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}|\sigma|^{2}\mathrm{d}x_{\mathrm{h}}\lesssim\|\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}}\|_{2}\int_{\partial\Omega}|\sigma|^{2}\mathrm{d}x_{\mathrm{h}}.

Noting that 0Tdivhuh2dτ<\int_{0}^{T}\|\mathrm{div}_{\mathrm{h}}u_{\mathrm{h}}\|_{2}\mathrm{d}\tau<\infty and σ0|Ω=0\sigma^{0}|_{\partial\Omega}=0, thus applying Gronwall’s inequality to the above inequality yields

σL2(Ω)2=0,\|\sigma\|^{2}_{L^{2}(\partial\Omega)}=0,

which implies

σ|Ω=0.\displaystyle\sigma|_{\partial\Omega}=0. (A.17)

Applying 3\partial_{3} to (2.3)1, then using the boundary condition (2.5), we can compute out that

(3ρt+uhh3ρ+3u33ρ)|Ω=0,(\partial_{3}\rho_{t}+u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\partial_{3}\rho+\partial_{3}u_{3}\partial_{3}\rho)|_{\partial\Omega}=0,

where ρ:=ρ¯+κ1σ\rho:=\bar{\rho}+\kappa^{-1}\sigma. Following the similar argument of (A.17) by further using the incompressible condition (2.6)3 and the boundary condition 3ρ0|Ω=0\partial_{3}\rho^{0}|_{\partial\Omega}=0, we easily derive from the above identity that

3ρ|Ω=0.\partial_{3}\rho|_{\partial\Omega}=0. (A.18)

Similarly, applying 32\partial^{2}_{3} to (2.6)1, and then using 32σ0|Ω=0\partial_{3}^{2}\sigma^{0}|_{\partial\Omega}=0, the incompressible condition and the boundary conditions in (2.5), (A.17), we have

(32σt+uhh32σ+23u332σ)|Ω=0,(\partial^{2}_{3}\sigma_{t}+u_{\mathrm{h}}\cdot\nabla_{\mathrm{h}}\partial^{2}_{3}\sigma+2\partial_{3}u_{3}\partial^{2}_{3}\sigma)|_{\partial\Omega}=0,

which obviously implies that

32σ|Ω=0.\partial_{3}^{2}\sigma|_{\partial\Omega}=0. (A.19)

Thus we arrive at (A.16), which presents that the solution σ\sigma satisfies the same boundary conditions as well as the initial data σ0\sigma^{0}. ∎

Lemma A.10.

Under the assumption of Lemma A.9, we have the following estimates:

iu2iu0+Δiu0 for i=0, 1,\displaystyle\|\nabla^{i}u\|_{2}\lesssim\|\nabla^{i}u\|_{0}+\|\Delta\nabla^{i}u\|_{0}\mbox{ for }i=0,\ 1, (A.20)
σ3Δσ0,\displaystyle\|\sigma\|_{3}\lesssim\|\nabla\Delta\sigma\|_{0}, (A.21)
sσ2sΔσ0 for s=1, 2,t.\displaystyle\|\partial_{s}\sigma\|_{2}\lesssim\|\partial_{s}\Delta\sigma\|_{0}\mbox{ for }s=1,\ 2,\ t. (A.22)
Proof.

Similar results can be found in [35, Lemma 2.2], however we also provide the proofs of (A.20)–(A.22) for reader’s convenience.

Recalling the boundary condition of uu in (2.5), we use both the elliptic estimates in Lemmas A.7 and A.8 to deduce that

u1u0+Δu0\|\nabla u\|_{1}\lesssim\|\nabla u\|_{0}+\|\Delta u\|_{0}

and

ju1ju0+Δju0 for 1j3,\|\nabla\partial_{j}u\|_{1}\lesssim\|\nabla\partial_{j}u\|_{0}+\|\Delta\partial_{j}u\|_{0}\mbox{ for }1\leqslant j\leqslant 3,

which, together with the interpolation inequality (A.4), imply (A.20).

Similarly, thanks to the boundary conditions of (σ,32σ)(\sigma,\partial_{3}^{2}\sigma) in (A.16), we also exploit the elliptic estimate in Lemma A.7 to deduce (A.21)–(A.22). This completes the proof of Lemma A.10. ∎

Theorem A.3.

Aubin–Lions theorem [53, Theorem 1.71]: Let T>0T>0, XBYX\hookrightarrow\hookrightarrow B\hookrightarrow Y be Banach spaces, {fn}n=1\{f_{n}\}_{n=1}^{\infty} a sequence bounded in Lq(IT,B)L1(IT,X)L^{q}(I_{T},B)\cap L^{1}(I_{T},X) and {dfn/dt}n=1\{\mathrm{d}f_{n}/\mathrm{d}t\}_{n=1}^{\infty} bounded in L1(IT,Y)L^{1}(I_{T},Y), where 1<q1<q\leqslant\infty. Then {fn}n=1\{f_{n}\}_{n=1}^{\infty} is relatively compact in Lp(IT,B)L^{p}(I_{T},B) for any 1p<q1\leqslant p<q.

Theorem A.4.

Arzelá–Ascoli theorem [53, Theorem 1.70]: Let T>0T>0 and BB, XX be Banach spaces such that BXB\hookrightarrow\hookrightarrow X is compact. Let fnf_{n} be a sequence of functions IT¯B\overline{I_{T}}\rightarrow B uniformly bounded in BB and uniformly continuous in XX. Then there exists fC0(IT¯,B)f\in C^{0}(\overline{I_{T}},B) such that fnff_{n}\rightarrow f strongly in C0(IT¯,X)C^{0}(\overline{I_{T}},X) at least for a chosen subsequence.

Theorem A.5.

Banach–Alaoglu theorem (see Sections 1.4.5.25 and 1.4.5.26 in [53]):

  1. 1.

    Let XX be a reflexive Banach space and let {un}X\{u_{n}\}\subset X be a bounded sequence. Then there exists a subsequence {unk}k=1\{u_{n_{k}}\}_{k=1}^{\infty} weakly convergent in XX.

  2. 2.

    Another version of the Banach–Alaoglu theorem: Let XX be a separable Banach space and let {fn}X\{f_{n}\}\subset X^{*} be a bounded sequence. Then there exists a subsequence {fnk}k=1\{f_{n_{k}}\}_{k=1}^{\infty} weakly-* convergent in XX^{*}.

Acknowledgements. The research of Fei Jiang was supported by NSFC (Grant Nos. 12371233 and 12231016), and the Natural Science Foundation of Fujian Province of China (Grant Nos. 2024J011011 and 2022J01105) and the Central Guidance on Local Science and Technology Development Fund of Fujian Province (Grant No. 2023L3003).

References

  • Adams and Fourier [2003] R.A. Adams, J.J.F. Fourier, Sobolev Spaces, Academic press, New York, 2003.
  • Antonelli and Spirito [2022] P. Antonelli, S. Spirito, Global existence of weak solutions to the Navier–Stokes–Korteweg equations, Ann. Inst. H. Poincaré C Anal. Non Linéaire 39 (2022) 171–200.
  • Audiard and Haspot [2017] C. Audiard, B. Haspot, Global well-posedness of the Euler–Korteweg system for small irrotational data, Comm. Math. Phys. 351 (2017) 201–247.
  • Benzoni-Gavage et al. [2007] S. Benzoni-Gavage, R. Danchin, S. Descombes, On the well-posedness for the Euler–Korteweg model in several space dimensions, Indiana Univ. Math. J. 56 (2007) 1499–1579.
  • Bian et al. [2014] D. Bian, L. Yao, C. Zhu, Vanishing capillarity limit of the compressible fluid models of Korteweg type to the Navier–Stokes equations, SIAM J. Math. Anal. 46 (2014) 1633–1650.
  • Bresch et al. [2008] D. Bresch, B. Desjardins, M. Gisclon, R. Sart, Instability results related to compressible Korteweg system, Ann. Univ. Ferrara Sez. 54 (2008) 11–36.
  • Bresch et al. [2003] D. Bresch, B. Desjardins, C.K. Lin, On some compressible fluid models: Korteweg, lubrication, and shallow water systems, Comm. Partial Differential Equations 28 (2003) 843–868.
  • Bresch et al. [2019] D. Bresch, M. Gisclon, I. Lacroix-Violet, On Navier–Stokes–Korteweg and Euler–Korteweg systems: application to quantum fluids models, Arch. Ration. Mech. Anal. 233 (2019) 975–1025.
  • Bresch et al. [2022] D. Bresch, M. Gisclon, I. Lacroix-Violet, A. Vasseur, On the exponential decay for compressible Navier–Stokes–Korteweg equations with a drag term, J. Math. Fluid Mech. 24 (2022) Paper No. 11, 16.
  • Browning and Kreiss [1982] G. Browning, H.O. Kreiss, Problems with different time scales for nonlinear partial differential equations, SIAM J. Appl. Math. 42 (1982) 704–718.
  • Burtea and Charve [2017] C. Burtea, F. Charve, Lagrangian methods for a general inhomogeneous incompressible Navier–Stokes–Korteweg system with variable capillarity and viscosity coefficients, SIAM J. Math. Anal. 49 (2017) 3476–3495.
  • Charve [2014] F. Charve, Local in time results for local and non-local capillary Navier–Stokes systems with large data, J. Differential Equations 256 (2014) 2152–2193.
  • Charve et al. [2021] F. Charve, R. Danchin, J. Xu, Gevrey analyticity and decay for the compressible Navier–Stokes system with capillarity, Indiana Univ. Math. J. 70 (2021) 1903–1944.
  • Charve and Haspot [2013] F. Charve, B. Haspot, Existence of a global strong solution and vanishing capillarity-viscosity limit in one dimension for the Korteweg system, SIAM J. Math. Anal. 45 (2013) 469–494.
  • Chen and Li [2021] Z. Chen, Y. Li, Asymptotic behavior of solutions to an impermeable wall problem of the compressible fluid models of Korteweg type with density-dependent viscosity and capillarity, SIAM J. Math. Anal. 53 (2021) 1434–1473.
  • Chen and Zhao [2014] Z. Chen, H. Zhao, Existence and nonlinear stability of stationary solutions to the full compressible Navier–Stokes–Korteweg system, J. Math. Pures Appl. 101 (2014) 330–371.
  • Danchin and Desjardins [2001] R. Danchin, B. Desjardins, Existence of solutions for compressible fluid models of Korteweg type, Ann. Inst. H. Poincaré C Anal. Non Linéaire 18 (2001) 97–133.
  • Ding et al. [2018] S. Ding, Q. Li, Z. Xin, Stability analysis for the incompressible Navier–Stokes equations with Navier boundary conditions, J. Math. Fluid Mech. 20 (2018) 603–629.
  • Dunn and Serrin [1985] J.E. Dunn, J. Serrin, On the thermomechanics of interstitial working, Arch. Ration. Mech. Anal. 88 (1985) 95–133.
  • Fanelli [2016] F. Fanelli, Highly rotating viscous compressible fluids in presence of capillarity effects, Math. Ann. 366 (2016) 981–1033.
  • Germain and LeFloch [2016] P. Germain, P. LeFloch, Finite energy method for compressible fluids: the Navier–Stokes–Korteweg model, Comm. Pure Appl. Math. 69 (2016) 3–61.
  • Goto [1990] S. Goto, Singular limit of the incompressible ideal magneto-fluid motion with respect to the Alfvén number, Hokkaido Math. J. 19 (1990) 175–187.
  • Guo and Tice [2013] Y. Guo, I. Tice, Decay of viscous surface waves without surface tension in horizontally infinitedomains, Anal. PDE 6 (2013) 1429–1533.
  • Haspot [2011] B. Haspot, Existence of global weak solution for compressible fluid models of Korteweg type, J. Math. Fluid Mech. 13 (2011) 223–249.
  • Haspot [2020] B. Haspot, Strong solution for Korteweg system in BMO1(N)\rm{BMO}^{-1}(\mathbb{R}^{N}) with initial density in LL^{\infty}, Proc. Lond. Math. Soc. (3) 121 (2020) 1766–1797.
  • Hattori and Li [1994] H. Hattori, D. Li, Solutions for two-dimensional system for materials of Korteweg type, SIAM J. Math. Anal. 25 (1994) 85–98.
  • Hong [2020] H. Hong, Strong solutions for the compressible barotropic fluid model of Korteweg type in the bounded domain, Z. Angew. Math. Phys. 71 (2020) Paper No. 85, 25.
  • Hong [2022] H. Hong, Stability of stationary solutions and viscous shock wave in the inflow problem for isentropic Navier–Stokes–Korteweg system, J. Differential Equations 314 (2022) 518–573.
  • Hou et al. [2018] X. Hou, H. Peng, C. Zhu, Global well-posedness of the 3D non-isothermal compressible fluid model of Korteweg type, Nonlinear Anal. Real World Applications 43 (2018) 18–53.
  • Hou et al. [2017] X. Hou, L. Yao, C. Zhu, Vanishing capillarity limit of the compressible non-isentropic Navier–Stokes–Korteweg system to Navier–Stokes system, J. Math. Anal. Appl. 448 (2017) 421–446.
  • Jiang et al. [2023a] F. Jiang, H. Jiang, S. Jiang, Rayleigh–Taylor instability in stratified compressible fluids with/without the interfacial surface tension, ariXiv:2309.13370 (2023a).
  • Jiang et al. [2022] F. Jiang, S. Jiang, Y. Zhao, On inhibition of the Rayleigh–Taylor instability by a horizontal magnetic field in ideal MHD fluids with velocity damping, J. Differential Equations 314 (2022) 574–652.
  • Jiang et al. [2023b] F. Jiang, S. Jiang, Y. Zhao, On inhibition of the Rayleigh–Taylor instability by a horizontal magnetic field in 2D non-resistive MHD fluids: the viscous case, CSIAM Trans. Appl. Math. 4 (2023b) 451–514.
  • Jiang et al. [2023c] F. Jiang, F. Li, Z. Zhang, On stability and instability of gravity driven Navier–Stokes–Korteweg model in two dimensions, arXiv preprint arXiv:2302.01013 (2023c).
  • Jiang et al. [2024] F. Jiang, Y. Zhang, Z. Zhang, On the inhibition of Rayleigh–Taylor instability by capillarity in the Navier–Stokes–Korteweg model, arXiv preprint arXiv:2402.07201 (2024).
  • Jiang et al. [2019] S. Jiang, Q. Ju, X. Xu, Small Alfvén number limit for incompressible magneto-hydrodynamics in a domain with boundaries, Sci. China Math. 62 (2019) 2229–2248.
  • Ju and Xu [2022] Q. Ju, J. Xu, Zero-Mach limit of the compressible Navier–Stokes–Korteweg equations, J. Math. Phys. 63 (2022) 111503.
  • Jüngel et al. [2014] A. Jüngel, C.K. Lin, K.C. Wu, An asymptotic limit of a Navier–Stokes system with capillary effects, Comm. Math. Phys. 329 (2014) 725–744.
  • Kawashima et al. [2021] S. Kawashima, Y. Shibata, J. Xu, The LpL^{p} energy methods and decay for the compressible Navier–Stokes equations with capillarity, J. Math. Pures Appl. (9) 154 (2021) 146–184.
  • Kobayashi et al. [2022] T. Kobayashi, M. Murata, H. Saito, Resolvent estimates for a compressible fluid model of Korteweg type and their application, J. Math. Fluid Mech. 24 (2022) Paper No. 12, 42.
  • Korteweg [1901] D.J. Korteweg, Sur la forme que prennent les équations du mouvements des fluides si l’on tient compte des forces capillaires causées par des variations de densité considérables mais connues et sur la théorie de la capillarité dans l’hypothèse d’une variation continue de la densité, Archives Néerlandaises des Sciences exactes et naturelles 6 (1901) 1–24.
  • Kotschote [2008] M. Kotschote, Strong solutions for a compressible fluid model of Korteweg type, Ann. Inst. H. Poincaré C Anal. Non Linéaire 25 (2008) 679–696.
  • Kotschote [2010] M. Kotschote, Strong well-posedness for a Korteweg-type model for the dynamics of a compressible non-isothermal fluid, J. Math. Fluid Mech. 12 (2010) 473–484.
  • Kotschote [2012] M. Kotschote, Dynamics of compressible non-isothermal fluids of non-Newtonian Korteweg type, SIAM J. Math. Anal. 44 (2012) 74–101.
  • Kotschote [2014] M. Kotschote, Existence and time-asymptotics of global strong solutions to dynamic Korteweg models, Indiana Univ. Math. J. 63 (2014) 21–51.
  • Li and Zhang [2023] F. Li, Z. Zhang, Stabilizing effect of capillarity in the Rayleigh–Taylor problem to the viscous incompressible capillary fluids, SIAM J. Math. Anal. 55 (2023) 3287–3315.
  • Li and Ding [2021] Q. Li, S. Ding, Global well-posedness of the Navier–Stokes equations with Navier-slip boundary conditions in a strip domain, Commun. Pure Appl. Anal. 20 (2021) 3561–3581.
  • Li and Yong [2016] Y. Li, W. Yong, Zero Mach number limit of the compressible Navier–Stokes–Korteweg equations, Commun. Math. Sci. 14 (2016) 233–247.
  • Lin [2012] F. Lin, Some analytical issues for elastic complex fluids, Comm. Pure Appl. Math. 65 (2012) 893–919.
  • Murata and Shibata [2020] M. Murata, Y. Shibata, The global well-posedness for the compressible fluid model of Korteweg type, SIAM J. Math. Anal. 52 (2020) 6313–6337.
  • Nguyen [2023] T.T. Nguyen, Influence of capillary number on nonlinear Rayleigh–Taylor instability to the Navier–Stokes–Korteweg equations, arXiv preprint arXiv:2312.05536 (2023).
  • Nirenberg [1959] L. Nirenberg, On elliptic partial differential equations, Estratto Ann. Sc. Norm. Super. Pisa (3) XIII(II) (1959) 1–48.
  • Novotný and Straškraba [2004] A. Novotný, I. Straškraba, Introduction to the Mathematical Theory of Compressible Flow, Oxford University Press, Oxford, 2004.
  • Saito [2020] H. Saito, On the maximal LpL_{p}LqL_{q} regularity for a compressible fluid model of Korteweg type on general domains, J. Differential Equations 268 (2020) 2802–2851.
  • Sha and Li [2019] K. Sha, Y. Li, Low Mach number limit of the three-dimensional full compressible Navier–Stokes–Korteweg equations, Z. Angew. Math. Phys. 70 (2019) 169–175.
  • Sohr [2001] H. Sohr, The Navier–Stokes equations: An elementary functional analytic approach, Birkhäuser Verlag, Basel, 2001.
  • Sy et al. [2006] M. Sy, D. Bresch, F. Guillén-González, J. Lemoine, M.A. Rodríguez-Bellido, Local strong solution for the incompressible Korteweg model, C. R. Math. Acad. Sci. Paris 342 (2006) 169–174.
  • Tan and Wang [2010] Z. Tan, Y. Wang, Strong solutions for the incompressible fluid models of Korteweg type, Acta Math. Sci. Ser. B (Engl. Ed.) 30 (2010) 799–809.
  • Wang and Zhang [2024] P. Wang, Z. Zhang, Vanishing capillarity-viscosity limit of the incompressible Navier–Stokes–Korteweg equations with slip boundary condition, Nonlinear Anal. 243 (2024) 113526.
  • Yang et al. [2015] J. Yang, L. Yao, C. Zhu, Vanishing capillarity-viscosity limit for the incompressible inhomogeneous fluid models of Korteweg type, Z. Angew. Math. Phys. 66 (2015) 2285–2303.
  • Zhang et al. [2023] X. Zhang, F. Tian, W. Wang, On Rayleigh–Taylor instability in Navier–Stokes–Korteweg equations, J. Inequal. Appl. (2023) Paper No. 119, 30.
  • Zhang [2022] Z. Zhang, Rayleigh–Taylor instability for viscous incompressible capillary fluids, J. Math. Fluid Mech. 24 (2022) Paper No. 70, 23.