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

Existence and uniqueness of strong solutions for the system of interaction between a compressible Navier-Stokes-Fourier fluid and a damped plate equation

Debayan Maity Centre for Applicable Mathematics, TIFR,
Post Bag No. 6503, GKVK Post Office, Bangalore-560065, India.
([email protected])
 and  Takéo Takahashi Université de Lorraine, CNRS, Inria, IECL, F-54000 Nancy, France.
([email protected])
Abstract.

The article is devoted to the mathematical analysis of a fluid-structure interaction system where the fluid is compressible and heat conducting and where the structure is deformable and located on a part of the boundary of the fluid domain. The fluid motion is modeled by the compressible Navier-Stokes-Fourier system and the structure displacement is described by a structurally damped plate equation. Our main results are the existence of strong solutions in an LpLqL^{p}-L^{q} setting for small time or for small data. Through a change of variables and a fixed point argument, the proof of the main results is mainly based on the maximal regularity property of the corresponding linear systems. For small time existence, this property is obtained by decoupling the linear system into several standard linear systems whereas for global existence and for small data, the maximal regularity property is proved by showing that the corresponding linear coupled fluid-structure operator is \mathcal{R}-sectorial.

Debayan Maity was partially supported by INSPIRE faculty fellowship (IFA18-MA128) and by Department of Atomic Energy, Government of India, under project no. 12-R & D-TFR-5.01-0520. Takéo Takahashi was partially supported by the ANR research project IFSMACS (ANR-15-CE40-0010).

Keywords. fluid-structure interaction, compressible Navier–Stokes–Fourier system, maximal regularity, \mathcal{R}-sectorial operators, strong solutions
AMS subject classifications. 35Q30, 76D05, 76N10

1. Introduction

In this work, we study the interaction between a viscous compressible heat conducting fluid and a viscoelastic structure located on a part of the fluid domain boundary. More precisely, we consider a smooth bounded domain 3\mathcal{F}\subset\mathbb{R}^{3} such that its boundary \partial\mathcal{F} contains a flat part ΓS:=𝒮×{0},\Gamma_{S}:=\mathcal{S}\times\{0\}, where 𝒮\mathcal{S} is a smooth bounded domain of 2.\mathbb{R}^{2}. We also set

Γ0=ΓS¯.\Gamma_{0}=\partial\mathcal{F}\setminus\overline{\Gamma_{S}}.

The set Γ0\Gamma_{0} is rigid and remains unchanged whereas on the flat part, we assume that there is a plate that can deform only in the transversal direction, and if we denote by η\eta the corresponding displacement, then ΓS\Gamma_{S} is transformed into

ΓS(η):={[x1,x2,η(x1,x2)];[x1,x2]𝒮}.\Gamma_{S}(\eta):=\left\{[x_{1},x_{2},\eta(x_{1},x_{2})]^{\top}\ ;\ [x_{1},x_{2}]^{\top}\in\mathcal{S}\right\}.

In our study, we consider only displacements η\eta regular enough and satisfying the boundary conditions (the plate is clamped):

η=sηnS=0on 𝒮\eta=\nabla_{s}\eta\cdot n_{S}=0\quad\text{on} \ \partial\mathcal{S} (1.1)

and a condition insuring that the deformed plate does not have any contact with the other part of the boundary of the fluid domain:

Γ0ΓS(η)=.\Gamma_{0}\cap\Gamma_{S}(\eta)=\emptyset. (1.2)

We have denoted by nSn_{S} the unitary exterior normal to 𝒮\partial\mathcal{S} and in the whole article we add the index ss in the gradient and in the Laplace operators if they apply to functions defined on 𝒮2\mathcal{S}\subset\mathbb{R}^{2} (and we keep the usual notation for functions defined on a domain of 3\mathbb{R}^{3}).

With the above notations and hypotheses, Γ0ΓS(η)¯\Gamma_{0}\cup\overline{\Gamma_{S}(\eta)} corresponds to a closed simple and regular surface whose interior is the fluid domain (η)\mathcal{F}(\eta). In what follows, we consider that η\eta is also a function of time and its evolution is governed by a damped plate equation.

(η)\mathcal{F}(\eta)ΓS(η)\Gamma_{S}(\eta)Γ0\Gamma_{0}

In (η(t))\mathcal{F}(\eta(t)), we assume that there is a viscous compressible heat conducting fluid and we denote by ρ~\widetilde{\rho}, v~\widetilde{v}, and ϑ~\widetilde{\vartheta} respectively its density, velocity and temperature. The equations modeling the evolution of these quantities can be written as follows:

{tρ~+div(ρ~v~)=0,t>0,x(η(t)),ρ~(tv~+(v~)v~)div𝕋(v~,π~)=0,t>0,x(η(t)),cvρ~(tϑ~+v~ϑ~)+π~divv~κΔϑ~=α(divv~)2+2μ|𝔻v~|2t>0,x(η(t)),ttη+Δs2ηΔstη=η(v~,π~)t>0,s𝒮,\begin{dcases}\partial_{t}\widetilde{\rho}+\operatorname{div}(\widetilde{\rho}\widetilde{v})=0,&t>0,\ x\in\mathcal{F}(\eta(t)),\\ \widetilde{\rho}\left(\partial_{t}\widetilde{v}+(\widetilde{v}\cdot\nabla)\widetilde{v}\right)-\operatorname{div}\mathbb{T}(\widetilde{v},\widetilde{\pi})=0,&t>0,\ x\in\mathcal{F}(\eta(t)),\\ c_{v}\widetilde{\rho}\left(\partial_{t}\widetilde{\vartheta}+\widetilde{v}\cdot\nabla\widetilde{\vartheta}\right)+\widetilde{\pi}\operatorname{div}\widetilde{v}-\kappa\Delta\widetilde{\vartheta}=\alpha(\operatorname{div}\widetilde{v})^{2}+2\mu\left|\mathbb{D}\widetilde{v}\right|^{2}&t>0,\ x\in\mathcal{F}(\eta(t)),\\ \partial_{tt}\eta+\Delta_{s}^{2}\eta-\Delta_{s}\partial_{t}\eta=\mathbb{H}_{\eta}(\widetilde{v},\widetilde{\pi})&t>0,\ s\in\mathcal{S},\end{dcases} (1.3)

with the boundary conditions

{v~(t,s,η(t,s))=tη(t,s)e3t>0,s𝒮,v~=0t>0,xΓ0,ϑ~n~(t,x)=0t>0,x(η(t)),η=sηnS=0t>0,s𝒮,\begin{dcases}\widetilde{v}(t,s,\eta(t,s))=\partial_{t}\eta(t,s)e_{3}&t>0,\ s\in\mathcal{S},\\ \widetilde{v}=0&t>0,\ x\in\Gamma_{0},\\ \frac{\partial\widetilde{\vartheta}}{\partial\widetilde{n}}(t,x)=0&t>0,\ x\in\partial\mathcal{F}({\eta(t)}),\\ \eta=\nabla_{s}\eta\cdot n_{S}=0&t>0,\ s\in\partial\mathcal{S},\end{dcases} (1.4)

and the initial conditions

{η(0,)=η10,tη(0,)=η20in𝒮,ρ~(0,)=ρ~0,v~(0,)=v~0,ϑ~(0,)=ϑ~0in(η10).\begin{dcases}\eta(0,\cdot)=\eta_{1}^{0},\quad\partial_{t}\eta(0,\cdot)=\eta_{2}^{0}\quad\text{in}\ \mathcal{S},\\ \widetilde{\rho}(0,\cdot)=\widetilde{\rho}^{0},\quad\widetilde{v}(0,\cdot)=\widetilde{v}^{0},\quad\widetilde{\vartheta}(0,\cdot)=\widetilde{\vartheta}^{0}\quad\text{in}\ \mathcal{F}(\eta_{1}^{0}).\end{dcases} (1.5)

In the above system (e1,e2,e3)(e_{1},e_{2},e_{3}) is the canonical basis of 3\mathbb{R}^{3}, the fluid stress tensor is defined by

𝕋(v~,π~)=2μ𝔻(v~)+(αdivv~π~)I3,𝔻(v~)=12(v~+v~),\mathbb{T}(\widetilde{v},\widetilde{\pi})=2\mu\mathbb{D}(\widetilde{v})+(\alpha\operatorname{div}\widetilde{v}-\widetilde{\pi})I_{3},\quad\mathbb{D}(\widetilde{v})=\frac{1}{2}\left(\nabla\widetilde{v}+\nabla\widetilde{v}^{\top}\right),

and the pressure law is given by

π~=R0ρ~ϑ~+π0.\widetilde{\pi}=R_{0}\widetilde{\rho}\widetilde{\vartheta}+\pi_{0}. (1.6)

The above physical constants satisfy

R0>0,μ>0 (viscosity),α+23μ>0,κ>0,cv>0,π0.R_{0}>0,\quad\mu>0\mbox{ (viscosity)},\quad\alpha+\frac{2}{3}\mu>0,\quad\kappa>0,\quad c_{v}>0,\quad\pi_{0}\in\mathbb{R}. (1.7)

For any matrix A,Bd()A,B\in\mathcal{M}_{d}(\mathbb{R}), we use the canonical scalar product and norm:

A:B=i,jaijbij,|A|=A:A.A:B=\sum_{i,j}a_{ij}b_{ij},\quad|A|=\sqrt{A:A}.

We have set

s=[y1,y2],Δs=y12+y22.\nabla_{s}=[\partial_{y_{1}},\partial_{y_{2}}]^{\top},\quad\Delta_{s}=\partial_{y_{1}}^{2}+\partial_{y_{2}}^{2}.

The function \mathbb{H} is defined by

η(v~,π~)=1+|sη|2(𝕋(v~,π~)n~)|ΓS(η(t))e3,\mathbb{H}_{\eta}(\widetilde{v},\widetilde{\pi})=-\sqrt{1+|\nabla_{s}\eta|^{2}}\left(\mathbb{T}(\widetilde{v},\widetilde{\pi})\widetilde{n}\right)|_{\Gamma_{S}(\eta(t))}\cdot e_{3}, (1.8)

where

n~=11+|sη|2[sη,1],\widetilde{n}=\frac{1}{\sqrt{1+|\nabla_{s}\eta|^{2}}}\left[-\nabla_{s}\eta,1\right]^{\top},

is the unit normal to ΓS(η(t))\Gamma_{S}(\eta(t)) outward (η(t)).\mathcal{F}(\eta(t)). Let us mention that the boundary conditions (1.4) are obtained by assuming that the fluid does not slip on the boundaries and that the plate is thermally insulated.

Fluid-structure interaction problems have been an active area of research among the engineers, physicist and mathematicians over the last few decades due to the numerous practical applications and the corresponding scientific challenges. The type of model considered in this article appears in the design of many engineering structures, e.g aircraft and bridges etc., ([4]) as well as in biomechanics ([7]).

Let us mention some related works from the literature. In the last two decades, there has been considerable number of works on similar fluid-structure systems where the fluid is modelled by incompressible flows. We refer to, for instance [23] and references therein for a concise description of recent progress regarding incompressible flows interacting with deformable structure (beam or plate) located on a part of the fluid domain boundary. Moreover, in some recent articles ([22, 5, 6]) existence and uniqueness of strong solutions (either local in time or for small initial data) were proved without the additional damping term (i.e., without the term Δstη-\Delta_{s}\partial_{t}\eta) in the beam/plate equation.

Concerning compressible fluids interacting with plate/beam equations through boundary of the fluid domain, there are only few results available in the literature. Global existence of weak solutions until the structure touches the boundary of the fluid domain were proved in [19, 9]. Local in time existence of strong solutions in the corresponding 2D/1D2D/1D case was recently obtained in [33]. Well-posedness and stability of linear compressible fluid-structure systems were studied in [10, 4].

Let us mention that all the above mentioned works correspond to a “Hilbert” space framework. In this article, we are interested in studying existence and uniqueness of strong solutions, local in time or global in time for small initial data, within an “LpLqL^{p}-L^{q}” framework. More precisely, we look for solutions in the spaces of functions which are LpL^{p} with respect to time and LqL^{q} with respect to space variable, with arbitrary p,q>1.p,q>1. In the context of fluid-solid interaction problems, there are only few articles available in the literature that studies well-posedness in an LpLqL^{p}-L^{q} framework. Let us mention [20, 32] (viscous incompressible fluid and rigid bodies), [25, 31, 24] (viscous compressible fluid and rigid bodies) and [30, 13] (viscous incompressible fluid interacting with viscoelastic structure located at the boundary of the fluid domain). In fact, this article is a compressible counterpart of our previous work [30].

The main novelties that we bring in this article are :

  • The full nonlinear free boundary system coupling viscous compressible Navier-Stokes-Fourier system and a viscoelastic structure located on a part of the fluid domain has not, at the best of our knowledge, been studied in the literature.

  • The existence and uniqueness results are proved in LpLqL^{p}-L^{q} setting.

  • Global in time existence for small initial data seems to be a new result for such coupled systems.

Let us emphasize that using the LpLqL^{p}-L^{q} setting allows us to weaken the regularity on the initial conditions (see for instance [33]). Moreover, this “LpLqL^{p}-{L^{q}}” framework is interesting even for studies in fluid-structure interaction problems done in the “L2L2L^{2}-L^{2}” framework: let us quote for instance the uniqueness of weak solutions ([21, 8]), the asymptotic behavior for large time ([17, 18]), and the asymptotic behavior for small structures ([28]).

1.1. Notation

To state our main results, we need to introduce some notations for the functional spaces. For Ωn\Omega\subset\mathbb{R}^{n} is an open set, q>1q>1 and k,k\in\mathbb{N}, we denote by Lq(Ω)L^{q}(\Omega) and Wk,q(Ω)W^{k,q}(\Omega) the standard Lebesgue and Sobolev spaces respectively. Ws,q(Ω)W^{s,q}(\Omega), with q>1q>1 and s+s\in\mathbb{R}_{+}^{*}, denotes the usual Sobolev-Slobodeckij space. Moreover, W0k,q(Ω)W^{k,q}_{0}(\Omega) is the completion of Cc(Ω)C_{c}^{\infty}(\Omega) with respect to the Wk,q(Ω)W^{k,q}(\Omega) norm. Let k,mk,m\in\mathbb{N}, k<mk<m. For 1p<1\leqslant p<\infty, 1q<1\leqslant q<\infty, we consider the standard definition of the Besov spaces by real interpolation of Sobolev spaces

Bq,ps()=(Wk,q(),Wm,q())θ,p where s=(1θ)k+θm,θ(0,1).B^{s}_{q,p}(\mathcal{F})=\left(W^{k,q}(\mathcal{F}),W^{m,q}(\mathcal{F})\right)_{\theta,p}\mbox{ where }s=(1-\theta)k+\theta m,\quad\theta\in(0,1).

We refer to [1] and [38] for a detailed presentation of the Besov spaces. We denote by CbkC^{k}_{b} is the set of continuous and bounded functions with derivatives continuous and bounded up to the order kk. For s(0,1)s\in(0,1) and a Banach space U,U, Fp,qs(0,T,U)F^{s}_{p,q}(0,T,U) stands for UU valued Lizorkin-Triebel space. For precise definition of such spaces we refer to [38]. If T(0,]T\in(0,\infty], we set

Wp,q1,2((0,T);)=Lp(0,T;W2,q())W1,p(0,T;Lq()),\displaystyle W^{1,2}_{p,q}((0,T);\mathcal{F})=L^{p}(0,T;W^{2,q}(\mathcal{F}))\cap W^{1,p}(0,T;L^{q}(\mathcal{F})),
Wp,q2,4((0,T);𝒮)=Lp(0,T;W4,q(𝒮))W1,p(0,T;W2,q(𝒮))W2,p(0,T;Lq(𝒮)),\displaystyle W^{2,4}_{p,q}((0,T);\mathcal{S})=L^{p}(0,T;W^{4,q}(\mathcal{S}))\cap W^{1,p}(0,T;W^{2,q}(\mathcal{S}))\cap W^{2,p}(0,T;L^{q}(\mathcal{S})),
Wp,q1,2((0,T);𝒮)=Lp(0,T;W2,q(𝒮))W1,p(0,T;Lq(𝒮)).\displaystyle W^{1,2}_{p,q}((0,T);\mathcal{S})=L^{p}(0,T;W^{2,q}(\mathcal{S}))\cap W^{1,p}(0,T;L^{q}(\mathcal{S})).

We have the following embeddings (see, for instance, [3, Theorem 4.10.2, p.180]),

Wp,q1,2((0,T);)Cb0([0,T);Bq,p2(11/p)()),W^{1,2}_{p,q}((0,T);\mathcal{F})\hookrightarrow C_{b}^{0}([0,T);B^{2(1-1/p)}_{q,p}(\mathcal{F})), (1.9)
Wp,q2,4((0,T);𝒮)Cb0([0,T);Bq,p2(21/p)(𝒮))Cb1([0,T);Bq,p2(11/p)(𝒮)).W^{2,4}_{p,q}((0,T);\mathcal{S})\hookrightarrow C_{b}^{0}([0,T);B^{2(2-1/p)}_{q,p}(\mathcal{S}))\cap C_{b}^{1}([0,T);B^{2(1-1/p)}_{q,p}(\mathcal{S})). (1.10)

In particular, in what follows, we use the following norm for Wp,q1,2((0,T);)W^{1,2}_{p,q}((0,T);\mathcal{F}):

fWp,q1,2((0,T);):=fLp(0,T;W2,q())+fW1,p(0,T;Lq())+fCb0([0,T);Bq,p2(11/p)())\|f\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})}:=\|f\|_{L^{p}(0,T;W^{2,q}(\mathcal{F}))}+\|f\|_{W^{1,p}(0,T;L^{q}(\mathcal{F}))}+\|f\|_{C_{b}^{0}([0,T);B^{2(1-1/p)}_{q,p}(\mathcal{F}))}

and we proceed similarly for the two other spaces.

We also introduce functional spaces with time decay. We write for any β\beta\in\mathbb{R}

𝔼β:,teβt.\mathbb{E}_{\beta}:\mathbb{R}\to\mathbb{R},t\mapsto e^{\beta t}.

We denote by Lβp(0,)L^{p}_{\beta}(0,\infty) the space 𝔼βLp(0,)\mathbb{E}_{-\beta}L^{p}(0,\infty), that is the set of functions ff such that teβtf(t)t\mapsto e^{\beta t}f(t) is in Lp(0,)L^{p}(0,\infty). The corresponding norm is

fLβp(0,):=𝔼βfLp(0,).\|f\|_{L^{p}_{\beta}(0,\infty)}:=\|\mathbb{E}_{\beta}f\|_{L^{p}(0,\infty)}.

We proceed similarly for all spaces on (0,)(0,\infty) or on [0,)[0,\infty).

Finally, we also need to introduce functional spaces for the fluid density, velocity and temperature depending on the displacement η\eta of the structure. Assume T(0,]T\in(0,\infty] and that ηWp,q2,4((0,T);𝒮)\eta\in W^{2,4}_{p,q}((0,T);\mathcal{S}) satisfies (1.1) and (1.2). We show in Section 3 that there exists a mapping X=XηX=X_{\eta} such that X(t,)X(t,\cdot) is a C1C^{1}-diffeomorphism from \mathcal{F} onto (η(t))\mathcal{F}(\eta(t)) and for any function f~\widetilde{f} defined for t(0,T)t\in(0,T) and x(η(t))x\in\mathcal{F}(\eta(t)), we then define

f(t,y):=f~(t,X(t,y))(t(0,T),y).f(t,y):=\widetilde{f}(t,X(t,y))\quad(t\in(0,T),\ y\in\mathcal{F}).

Then we define the following sets as follows

f~Wr,p(0,T;Ws,q((η())))\displaystyle\widetilde{f}\in W^{r,p}(0,T;W^{s,q}(\mathcal{F}(\eta(\cdot)))) iffWr,p(0,T;Ws,q()),\displaystyle\quad\text{if}\quad f\in W^{r,p}(0,T;W^{s,q}(\mathcal{F})),
f~Wp,q1,2((0,T);(η()))\displaystyle\widetilde{f}\in W^{1,2}_{p,q}((0,T);\mathcal{F}(\eta(\cdot))) iffWp,q1,2((0,T);),\displaystyle\quad\text{if}\quad f\in W^{1,2}_{p,q}((0,T);\mathcal{F}),
f~C0([0,T];Bq,p2(11/p)((η())))\displaystyle\widetilde{f}\in C^{0}([0,T];B^{2(1-1/p)}_{q,p}(\mathcal{F}(\eta(\cdot)))) iffC0([0,T];Bq,p2(11/p)()\displaystyle\quad\text{if}\quad f\in C^{0}([0,T];B^{2(1-1/p)}_{q,p}(\mathcal{F})

and a similar definition for all the other spaces.

1.2. Statement of the main results

Let us give the conditions we require on (p,q)(p,q) and on the initial data for the system (1.3)–(1.8):

2<p<,3<q<,1p+12q12,2<p<\infty,\quad 3<q<\infty,\quad\frac{1}{p}+\frac{1}{2q}\neq\frac{1}{2}, (1.11)
η10Bq,p2(21/p)(𝒮),η20Bq,p2(11/p)(𝒮),ρ~0W1,q((η10)),min(η10)¯ρ~0>0,\displaystyle\eta_{1}^{0}\in B^{2(2-1/p)}_{q,p}(\mathcal{S}),\quad\eta_{2}^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{S}),\quad\widetilde{\rho}^{0}\in W^{1,q}({\mathcal{F}({\eta_{1}^{0}})}),\quad\min_{\overline{\mathcal{F}(\eta_{1}^{0})}}\widetilde{\rho}^{0}>0, (1.12)
v~0Bq,p2(11/p)((η10))3,ϑ~0Bq,p2(11/p)((η10)),\displaystyle\widetilde{v}^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{F}({\eta_{1}^{0}}))^{3},\qquad\widetilde{\vartheta}^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{F}({\eta_{1}^{0}})), (1.13)

with the compatibility conditions

η10=η10nS=η20=0 on 𝒮,v~0=0 on Γ0,v~0=η20e3 on ΓS(η10),\displaystyle\eta_{1}^{0}=\displaystyle\nabla\eta_{1}^{0}\cdot n_{S}=\eta_{2}^{0}=0\mbox{ on }\partial\mathcal{S},\quad\widetilde{v}^{0}=0\mbox{ on }\Gamma_{0},\quad\widetilde{v}^{0}=\eta_{2}^{0}e_{3}\mbox{ on }\Gamma_{S}(\eta_{1}^{0}), (1.14)
η20nS on 𝒮andϑ~0n=0 on (η10), if 1p+12q<12.\displaystyle\displaystyle\nabla\eta_{2}^{0}\cdot n_{S}\mbox{ on }\partial\mathcal{S}\quad\text{and}\quad\frac{\partial\widetilde{\vartheta}^{0}}{\partial n}=0\mbox{ on }\partial\mathcal{F}(\eta_{1}^{0}),\quad\mbox{ if }\quad\frac{1}{p}+\frac{1}{2q}<\frac{1}{2}. (1.15)

Note that, all the traces in the above relation makes sense for our choice of pp and qq (see for instance, [38, p. 200]).

We also need a geometrical condition on the initial deformation. Using that \mathcal{F} is a smooth domain, there exist two smooth surfaces η:𝒮\eta_{-}:\mathcal{S}\to\mathbb{R}_{-}^{*}, η+:𝒮+\eta_{+}:{\mathcal{S}}\to\mathbb{R}_{+}^{*} such that

{[y1,y2,y3]𝒮×;y3(η(y1,y2),0)},\left\{[y_{1},y_{2},y_{3}]^{\top}\in{\mathcal{S}}\times\mathbb{R}\ ;\ y_{3}\in(\eta_{-}(y_{1},y_{2}),0)\right\}\subset\mathcal{F}, (1.16)
{[y1,y2,y3]𝒮×;y3(0,η+(y1,y2))}3¯.\left\{[y_{1},y_{2},y_{3}]^{\top}\in{\mathcal{S}}\times\mathbb{R}\ ;\ y_{3}\in(0,\eta_{+}(y_{1},y_{2}))\right\}\subset\mathbb{R}^{3}\setminus\overline{\mathcal{F}}. (1.17)

Then our geometrical condition on the initial deformation writes

η<η10<η+in𝒮.\eta_{-}<\eta_{1}^{0}<\eta_{+}\quad\text{in}\ {\mathcal{S}}. (1.18)

This yields in particular that Γ0ΓS(η10)=\Gamma_{0}\cap\Gamma_{S}(\eta_{1}^{0})=\emptyset. According to the geometry, we can in some situation remove the condition η10<η+\eta_{1}^{0}<\eta_{+}. Note that this condition is not a smallness condition, η+\eta_{+} and η\eta_{-} do not need to be small.

Our main results are the following two theorems. The first one is the local in time existence and uniqueness :

Theorem 1.1.

Assume (p,q)(p,q) satisfies (1.11) and that [ρ~0,v~0,ϑ~0,η10,η20][\widetilde{\rho}^{0},\widetilde{v}^{0},\widetilde{\vartheta}^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top} satisfies (1.12)–(1.15) and (1.18). Then there exists T>0,T>0, depending only on initial data, such that the system (1.3)–(1.8) admits a unique strong solution [ρ~,v~,ϑ~,η][\widetilde{\rho},\widetilde{v},\widetilde{\vartheta},\eta]^{\top} satisfying

ρ~W1,p(0,T;W1,q((η()))),v~Wp,q1,2((0,T);(η()))3,\displaystyle\widetilde{\rho}\in W^{1,p}(0,T;W^{1,q}(\mathcal{F}(\eta(\cdot)))),\quad\widetilde{v}\in W^{1,2}_{p,q}((0,T);\mathcal{F}(\eta(\cdot)))^{3},
ϑWp,q1,2((0,T);(η())),ηWp,q2,4((0,T);𝒮),\displaystyle\vartheta\in W^{1,2}_{p,q}((0,T);\mathcal{F}(\eta(\cdot))),\quad\eta\in W^{2,4}_{p,q}((0,T);\mathcal{S}),
Γ0ΓS(η(t))=(t[0,T]),ρ~(t,x)>0(t[0,T],x(η(t))¯).\displaystyle\Gamma_{0}\cap\Gamma_{S}(\eta(t))=\emptyset\quad(t\in[0,T]),\quad\widetilde{\rho}(t,x)>0\quad(t\in[0,T],\ x\in\overline{\mathcal{F}(\eta(t))}).

Our second main result states the global existence and uniqueness under a smallness condition on the initial data. Let ρ¯\overline{\rho} and ϑ¯\overline{\vartheta} be two given positive constants. Let us take in the pressure law (1.6)

π0=R0ρ¯ϑ¯.{\pi}_{0}=-R_{0}\overline{\rho}\overline{\vartheta}. (1.19)

With the above choice of π0\pi_{0}, [ρ~,v~,ϑ~,η]=[ρ¯,0,ϑ¯,0]\left[\widetilde{\rho},\widetilde{v},\widetilde{\vartheta},\eta\right]^{\top}=\left[\overline{\rho},0,\overline{\vartheta},0\right]^{\top} is a steady state solution to the system (1.3)–(1.8).

Then our result states as follows:

Theorem 1.2.

Assume (p,q)(p,q) satisfies (1.11) and assume that ρ¯\overline{\rho} and ϑ¯\overline{\vartheta} are two given positive constants such that (1.19) holds. Then there exist β>0\beta>0 and R>0R>0 such that for any [ρ~0,v~0,ϑ~0,η10,η20][\widetilde{\rho}^{0},\widetilde{v}^{0},\widetilde{\vartheta}^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top} satisfying (1.12)–(1.15), and

ρ~0ρ¯W1,q((η10))+v~0Bq,p2(11/p)((η10))3+ϑ~0ϑ¯Bq,p2(11/p)((η10))+η10Bq,p2(21/p)(𝒮)+η20Bq,p2(11/p)(𝒮)R,\left\|\widetilde{\rho}^{0}-\overline{\rho}\right\|_{W^{1,q}(\mathcal{F}(\eta_{1}^{0}))}+\left\|\widetilde{v}^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F}(\eta_{1}^{0}))^{3}}+\left\|\widetilde{\vartheta}^{0}-\overline{\vartheta}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F}(\eta_{1}^{0}))}\\ +\left\|\eta_{1}^{0}\right\|_{B^{2(2-1/p)}_{q,p}(\mathcal{S})}+\left\|\eta_{2}^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}\leqslant R, (1.20)

the system (1.3)–(1.8) admits a unique strong solution [ρ~,v~,ϑ~,η]\left[\widetilde{\rho},\widetilde{v},\widetilde{\vartheta},\eta\right]^{\top} satisfying

ρ~Cb0([0,);W1,q((η()))),ρ~Wβ1,p(0,;Lq((η())))3,tρ~Lβp(0,;W1,q((η()))),\displaystyle\widetilde{\rho}\in C^{0}_{b}([0,\infty);W^{1,q}(\mathcal{F}(\eta(\cdot)))),\;\nabla\widetilde{\rho}\in{W^{1,p}_{\beta}(0,\infty;L^{q}(\mathcal{F}(\eta(\cdot))))}^{3},\;\partial_{t}\widetilde{\rho}\in{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}(\eta(\cdot))))},
v~Wp,q,β1,2((0,T);(η()))3,\displaystyle\widetilde{v}\in W^{1,2}_{p,q,\beta}((0,T);\mathcal{F}(\eta(\cdot)))^{3},
ϑ~Cb0([0,);Bq,p2(11/p)((η()))),ϑ~Lβp(0,;W1,q((η())))3,tϑ~Lβp(0,;Lq((η()))),\displaystyle\widetilde{\vartheta}\in C^{0}_{b}([0,\infty);B^{2(1-1/p)}_{q,p}(\mathcal{F}(\eta(\cdot)))),\quad\nabla\widetilde{\vartheta}\in L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}(\eta(\cdot))))^{3},\quad\partial_{t}\widetilde{\vartheta}\in{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}(\eta(\cdot))))},
ηCb0([0,);Bq,p2(21/p)(𝒮)),ηLβp(0,;W4,q(𝒮))+L(0,;W4,q(𝒮)),\displaystyle\eta\in C^{0}_{b}([0,\infty);B^{2(2-1/p)}_{q,p}(\mathcal{S})),\quad\eta\in L^{p}_{\beta}(0,\infty;W^{4,q}(\mathcal{S}))+L^{\infty}(0,\infty;W^{4,q}(\mathcal{S})),
tηWp,q,β1,2((0,);𝒮),\displaystyle\partial_{t}\eta\in W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{S}),

and

Γ0ΓS(η(t))=(t[0,)),ρ~(t,x)>0(t[0,),x(η(t))¯).\Gamma_{0}\cap\Gamma_{S}(\eta(t))=\emptyset\quad(t\in[0,\infty)),\quad\widetilde{\rho}(t,x)>0\quad(t\in[0,\infty),\ x\in\overline{\mathcal{F}(\eta(t))}).
Remark 1.3.

Let us make the following remarks on the above results:

  1. (1)

    Note that, in 1.1 we do not need initial displacement of the plate η10\eta_{1}^{0} to be zero. This is a difference with respect to previous works, for instance [33] or our previous work [30] (with an incompressible fluid). Here we manage to handle this case by modifying our change of variables (see Section 3.1).

  2. (2)

    In 1.1 and 1.2, we do not have any “loss of regularity” at initial time. More precisely, we obtain the continuity of the solution with respect to time in the same space where the initial data belong. Due to the coupling between the fluid system and the structure equations, some results in the literature are stated with this loss of regularity: for instance in [33, Theorem 1.7], there is a loss of order 1/21/2 in the space regularity for the fluid velocity at initial time.

  3. (3)

    As explained above since we work in the “LpLqL^{p}-{L^{q}}” framework, we need less regularity on the initial conditions that in the Hilbert case done by [33]. More precisely, in [33] the author assumes that the initial conditions satisfy

    η10=0,η20W3,2(𝒮),ρ~0W2,2((η10)),v~0W3,2((η10))3,\displaystyle\eta_{1}^{0}=0,\quad\eta_{2}^{0}\in W^{3,2}(\mathcal{S}),\quad\widetilde{\rho}^{0}\in W^{2,2}({\mathcal{F}({\eta_{1}^{0}})}),\quad\widetilde{v}^{0}\in W^{3,2}(\mathcal{F}({\eta_{1}^{0}}))^{3},

    with the corresponding compatibility conditions.

  4. (4)

    1.1 and 1.2 can be adapted to the 2D/1D2D/1D case, that is where \mathcal{F} is a regular bounded domain in 2\mathbb{R}^{2} such that \partial{\mathcal{F}} contains a flat part ΓS=𝒮×{0},\Gamma_{S}=\mathcal{S}\times\{0\}, where 𝒮\mathcal{S} is an open bounded interval of .\mathbb{R}. In that case we can take p,q(2,)p,q\in(2,\infty) such that 1p+12q12.\displaystyle\frac{1}{p}+\frac{1}{2q}\neq\frac{1}{2}.

  5. (5)

    Instead of taking heat conducting fluid, we can also consider barotropic fluid model, i.e., the system (1.3) without the temperature equation and with the pressure law π~=ρ~γ,\widetilde{\pi}=\widetilde{\rho}^{\gamma}, for some constant γ>1.\gamma>1. In that case, we can take 1<p<1<p<\infty and n<q<n<q<\infty (n=2n=2 or 3,3, the dimension of the fluid domain) such that 1p+12q1.\displaystyle\frac{1}{p}+\frac{1}{2q}\neq 1.

The proofs of 1.1 and 1.2 follow a standard approach in the literature on well-posedness for fluid-solid interaction systems. One of the main difficulties in studying fluid-structure models is that the fluid system is written in the deformed configuration (in Eulerian variables) whereas the structure equations are written in the reference configuration (in Lagrangian variables). Since the fluid domain (η(t))\mathcal{F}(\eta(t)) depends on the structure displacement, which is one of unknowns, we first reformulate the problem in a fixed domain. This is achieved thanks to a combination of a geometric change of variables (defined through the initial displacement of the structure) and a Lagrangian change of coordinates. With this combined change of variables, we reformulate the problem in the reference domain .\mathcal{F}. In most of the existing literature, a geometric change of variables via the displacement of the fluid-structure interface is used to rewrite the problem in a fixed domain ([29, 22, 5, 30]). However, in the context of compressible fluid-structure systems, it is more convenient to use a Lagrangian (see for instance [24]) or a combination of geometric and Lagrangian change of coordinates ([25]). In fact, such transformations allow us to use basic contraction mapping theorem. More precisely, this transformation eliminates the difficult term v~ρ~\widetilde{v}\cdot\nabla\widetilde{\rho} from the density equation.

Next, we associate the original nonlinear problem to a linear one involving the non-homogeneous terms. In the case of the local in time existence, this linear system can be partially decoupled (see system (3.24)-(3.27)). The LpLqL^{p}-L^{q} regularity of such linear system over finite time interval is obtained by combining various existing maximal LpLqL^{p}-L^{q} results for parabolic systems. One of the difficulties is that due to the non-zero initial displacement of the beam, we are dealing with linear operators involving variable coefficients. For the global existence part, we use a “monolithic” type approach, which means that the linearized system in consideration is still a coupled system of fluid and structure equations (see system (4.20)-(4.22)). A crucial step is to show the maximal LpLqL^{p}-L^{q} property of the associated fluid-structure linear operator in the infinite time horizon. This is achieved by showing that this operator is \mathcal{R}-sectorial and generates an exponentially stable semigroup in a suitable function space. Finally, for both the existence for small time and the existence for small initial conditions, we end the proof by using the Banach fixed point theorem.

The plan of the paper is as follows. In Section 2, we recall some results concerning \mathcal{R}-sectorial operators that are used both for the proofs of 1.1 and 1.2. Then, we prove 1.1 in Section 3. In Section 3.1, we introduce the combination of Lagrangian and geometric change of coordinates to reformulate the original problem in the reference configuration. Local in time existence for the system written in reference configuration is stated in 3.1. In Section 3.2, we prove the maximal LpLqL^{p}-L^{q} regularity of a linearized system, whereas in Section 3.3, we derive estimates for the nonlinear terms in order to prove 3.1 by using the Banach fixed point theorem. Section 4 is devoted to the proof of 1.2. In Section 4.1 we apply the same change of variables than in Section 3.1 with some slight modifications and then linearize the system around a constant steady state. The global in time existence for small initial data for the system written in the reference configuration is stated in 4.1. In Section 4.2, we introduce the so-called fluid-structure operator and we show that it is an \mathcal{R}-sectorial operator and in Section 4.3 that is generates an exponentially stable semigroup in a suitable function space. The maximal LpLqL^{p}-L^{q} regularity of the linearized system is proved in Section 4.4. Finally, in Section 4.5 we show 4.1. by using the Banach fixed point theorem.

2. Some Background on \mathcal{R}-sectorial Operators

We recall here some definitions and properties related to \mathcal{R}-sectorial operators. First, let us give the definition of \mathcal{R}-boundedness (\mathcal{R} for Randomized) for a family of operators (see, for instance, [40, 11, 27]):

Definition 2.1.

Assume 𝒳\mathcal{X} and 𝒴\mathcal{Y} are Banach spaces and (𝒳,𝒴)\mathcal{E}\subset\mathcal{L}(\mathcal{X},\mathcal{Y}). We say that \mathcal{E} is \mathcal{R}-bounded if there exist p[1,)p\in[1,\infty) and a constant C>0C>0, such that for any integer N1N\geqslant 1, any T1,TNT_{1},\ldots T_{N}\in\mathcal{E}, any independent Rademacher random variables r1,,rNr_{1},\ldots,r_{N}, and any x1,,xN𝒳x_{1},\ldots,x_{N}\in\mathcal{X},

(𝔼j=1NrjTjxj𝒴p)1/pC(𝔼j=1Nrjxj𝒳p)1/p.\left(\mathbb{E}\left\|\sum_{j=1}^{N}r_{j}T_{j}x_{j}\right\|_{\mathcal{Y}}^{p}\right)^{1/p}\leqslant C\left(\mathbb{E}\left\|\sum_{j=1}^{N}r_{j}x_{j}\right\|_{\mathcal{X}}^{p}\right)^{1/p}.

The p\mathcal{R}_{p}-bound of \mathcal{E} on (𝒳,𝒴)\mathcal{L}(\mathcal{X},\mathcal{Y}), denoted by p()\mathcal{R}_{p}(\mathcal{E}), is the smallest constant CC in the above inequality.

Let us recall that a Rademacher random variable is a symmetric random variables with value in {1,1}\{-1,1\} and that 𝔼\mathbb{E} denotes the expectation of a random variable. Note that the above definition is independent of p[1,)p\in[1,\infty) (see, for instance, [11, p.26]). The p\mathcal{R}_{p}-bound has the following properties (see, for instance, Proposition 3.4 in [11]):

p(1+2)p(1)+p(2),p(12)p(1)p(2).\mathcal{R}_{p}(\mathcal{E}_{1}+\mathcal{E}_{2})\leqslant\mathcal{R}_{p}(\mathcal{E}_{1})+\mathcal{R}_{p}(\mathcal{E}_{2}),\quad\mathcal{R}_{p}(\mathcal{E}_{1}\mathcal{E}_{2})\leqslant\mathcal{R}_{p}(\mathcal{E}_{1})\mathcal{R}_{p}(\mathcal{E}_{2}). (2.1)

For any β(0,π)\beta\in(0,\pi), we consider the sector \mathcal{R}-sectorial operators:

Σβ={λ{0};|arg(λ)|<β}.\Sigma_{\beta}=\{\lambda\in\mathbb{C}\setminus\{0\}\ ;\ |\arg(\lambda)|<\beta\}. (2.2)

We can introduce the definition of :

Definition 2.2 (sectorial and \mathcal{R}-sectorial operators).

Let A:𝒟(A)𝒳A:\mathcal{D}(A)\to\mathcal{X} be a densely defined closed linear operator on the Banach space 𝒳\mathcal{X}. The operator AA is (\mathcal{R})-sectorial of angle β(0,π)\beta\in(0,\pi) if

Σβρ(A)\Sigma_{\beta}\subset\rho(A)

and if the set

Rβ={λ(λA)1;λΣβ}R_{\beta}=\left\{\lambda(\lambda-A)^{-1}\ ;\ \lambda\in\Sigma_{\beta}\right\}

is (\mathcal{R})-bounded in (𝒳)\mathcal{L}(\mathcal{X}).

We denote by Mβ(A)M_{\beta}(A) (respectively β(A)\mathcal{R}_{\beta}(A)) the bound (respectively the \mathcal{R}-bound) of RβR_{\beta}. One can replace in the above definitions RβR_{\beta} by the set

Rβ~={A(λA)1;λΣβ}.\widetilde{R_{\beta}}=\left\{A(\lambda-A)^{-1}\ ;\ \lambda\in\Sigma_{\beta}\right\}.

In that case, we denote the uniform bound and the \mathcal{R}-bound by M~β(A)\widetilde{M}_{\beta}(A) and β~(A)\widetilde{\mathcal{R}_{\beta}}(A).

The following result, due to [40] (see also [11, p.45]), shows the important relation between the notion of \mathcal{R}-sectoriality and the maximal regularity of type LpL^{p}:

Theorem 2.3.

Assume 𝒳\mathcal{X} is a UMD Banach space and that A:𝒟(A)𝒳A:\mathcal{D}(A)\to\mathcal{X} is a densely defined, closed linear operator on 𝒳\mathcal{X}. Then the following assertions are equivalent:

  1. (1)

    For any T(0,]T\in(0,\infty] and for any fLp(0,T;𝒳)f\in L^{p}(0,T;\mathcal{X}), the Cauchy problem

    u=Au+fin(0,T),u(0)=0u^{\prime}=Au+f\quad\text{in}\quad(0,T),\quad u(0)=0 (2.3)

    admits a unique solution uu with u,AuLp(0,T;𝒳)u^{\prime},Au\in L^{p}(0,T;\mathcal{X}) and there exists a constant C>0C>0 such that

    uLp(0,T;𝒳)+AuLp(0,T;𝒳)CfLp(0,T;𝒳).\|u^{\prime}\|_{L^{p}(0,T;\mathcal{X})}+\|Au\|_{L^{p}(0,T;\mathcal{X})}\leqslant C\|f\|_{L^{p}(0,T;\mathcal{X})}.
  2. (2)

    AA is \mathcal{R}-sectorial of angle >π2>\frac{\pi}{2}.

In the above definition, we recall that 𝒳\mathcal{X} is a UMD Banach space if the Hilbert transform is bounded in Lp(;𝒳)L^{p}(\mathbb{R};\mathcal{X}) for p(1,)p\in(1,\infty). In particular, the closed subspaces of Lq(Ω)L^{q}(\Omega) for q(1,)q\in(1,\infty) are UMD Banach spaces. We refer the reader to [3, pp.141–147] for more information on UMD spaces.

Combining the above theorem with [15, Theorem 2.4] and [37, Theorem 1.8.2], we can consider the following Cauchy problem

u=Au+fin(0,),u(0)=u0.u^{\prime}=Au+f\quad\text{in}\quad(0,\infty),\quad u(0)=u_{0}. (2.4)
Corollary 2.4.

Assume 𝒳\mathcal{X} is a UMD Banach space, 1<p<1<p<\infty and AA is a closed, densely defined operator in 𝒳\mathcal{X} with domain 𝒟(A).\mathcal{D}(A). Let us suppose also that AA is a \mathcal{R}-sectorial operator of angle >π2>\frac{\pi}{2} and that the semigroup generated by AA has negative exponential type. Then for any u0(𝒳,𝒟(A))11/p,pu_{0}\in(\mathcal{X},\mathcal{D}(A))_{1-1/p,p} and for any fLp(0,;𝒳),f\in L^{p}(0,\infty;\mathcal{X}), the system (2.4) admits a unique solution in Lp(0,;𝒟(A))W1,p(0,;𝒳).L^{p}(0,\infty;\mathcal{D}(A))\cap W^{1,p}(0,\infty;\mathcal{X}).

Finally, we will need the following result ([26, Corollary 2]) on the perturbation theory of \mathcal{R}-sectoriality.

Proposition 2.5.

Suppose AA is a \mathcal{R}-sectorial operator of angle β\beta on a Banach space 𝒳\mathcal{X}. Assume that B:𝒟(B)𝒳B:\mathcal{D}(B)\to\mathcal{X} is a linear operator such that 𝒟(A)𝒟(B)\mathcal{D}(A)\subset\mathcal{D}(B) and such that there exist a,b0a,b\geqslant 0 satisfying

Bx𝒳aAx𝒳+bx𝒳(x𝒟(A)).\|Bx\|_{\mathcal{X}}\leqslant a\|Ax\|_{\mathcal{X}}+b\|x\|_{\mathcal{X}}\quad(x\in\mathcal{D}(A)). (2.5)

If

a<1M~β(A)β~(A)andλ>bMβ(A)β~(A)1aM~β(A)β~(A),a<\frac{1}{\widetilde{M}_{\beta}(A)\widetilde{\mathcal{R}_{\beta}}(A)}\quad\text{and}\quad\lambda>\frac{bM_{\beta}(A)\widetilde{\mathcal{R}_{\beta}}(A)}{1-a\widetilde{M}_{\beta}(A)\widetilde{\mathcal{R}_{\beta}}(A)},

then A+BλA+B-\lambda is \mathcal{R}-sectorial of angle β\beta.

3. Local in time existence

The aim of this section is to prove 1.1.

3.1. Change of variables and Linearization

In this subsection, we consider a change of variables to transform the moving domain (η(t))\mathcal{F}(\eta(t)) into the fixed domain \mathcal{F}. For this we use the Lagrangian change of variables to write everything in (η10)\mathcal{F}(\eta_{1}^{0}) and a geometric change of variables to transform (η10)\mathcal{F}(\eta_{1}^{0}) into \mathcal{F}. Let us start with the second one.

First using that \mathcal{F} is smooth, there exist an open bounded neighborhood 𝒮~\widetilde{\mathcal{S}} of 𝒮¯\overline{\mathcal{S}} in 2\mathbb{R}^{2}, ε~>0\widetilde{\varepsilon}>0 and η~:𝒮~\widetilde{\eta}:\widetilde{\mathcal{S}}\to\mathbb{R} smooth such that

(𝒮~×[ε~,ε~])={(s,η~(s)),s𝒮~}.\left(\widetilde{\mathcal{S}}\times[-\widetilde{\varepsilon},\widetilde{\varepsilon}]\right)\cap\partial\mathcal{F}=\left\{(s,\widetilde{\eta}(s)),\ s\in\widetilde{\mathcal{S}}\right\}.

We have in particular that η~0\widetilde{\eta}\equiv 0 in 𝒮\mathcal{S}. From (1.16), (1.17), we can extend η\eta_{-} and η+\eta_{+} with

{[y1,y2,y3]𝒮~×;y3(η(y1,y2),η~(y1,y2))},\left\{[y_{1},y_{2},y_{3}]^{\top}\in\widetilde{\mathcal{S}}\times\mathbb{R}\ ;\ y_{3}\in(\eta_{-}(y_{1},y_{2}),\widetilde{\eta}(y_{1},y_{2}))\right\}\subset\mathcal{F},
{[y1,y2,y3]𝒮~×;y3(η~(y1,y2),η+(y1,y2))}3¯.\left\{[y_{1},y_{2},y_{3}]^{\top}\in\widetilde{\mathcal{S}}\times\mathbb{R}\ ;\ y_{3}\in(\widetilde{\eta}(y_{1},y_{2}),\eta_{+}(y_{1},y_{2}))\right\}\subset\mathbb{R}^{3}\setminus\overline{\mathcal{F}}.

Using (1.13)–(1.14) and that q>3q>3, we can extend η10\eta_{1}^{0} by 0 in 2𝒮¯\mathbb{R}^{2}\setminus\overline{\mathcal{S}} with η10W2,q(2).\eta_{1}^{0}\in W^{2,q}(\mathbb{R}^{2}). Then (1.18) yields the existence of ε(0,1)\varepsilon\in(0,1) such that

η(1ε)<η10<η+(1ε)in𝒮~.\eta_{-}(1-\varepsilon)<\eta_{1}^{0}<\eta_{+}(1-\varepsilon)\quad\text{in}\ \widetilde{\mathcal{S}}.

We consider χCc(3)\chi\in C_{c}^{\infty}(\mathbb{R}^{3}) such that

suppχ{[y1,y2,y3]𝒮~×;y3(η(y1,y2),η+(y1,y2))},\operatorname{supp}\chi\subset\left\{[y_{1},y_{2},y_{3}]^{\top}\in\widetilde{\mathcal{S}}\times\mathbb{R}\ ;\ y_{3}\in(\eta_{-}(y_{1},y_{2}),\eta_{+}(y_{1},y_{2}))\right\},
χ1in{[y1,y2,y3]𝒮×;y3((1ε)η(y1,y2),(1ε)η+(y1,y2))}.\chi\equiv 1\quad\text{in}\ \left\{[y_{1},y_{2},y_{3}]^{\top}\in\mathcal{S}\times\mathbb{R}\ ;\ y_{3}\in((1-\varepsilon)\eta_{-}(y_{1},y_{2}),(1-\varepsilon)\eta_{+}(y_{1},y_{2}))\right\}.

We also define

Λ(y1,y2,y3)=η10(y1,y2)χ(y1,y2,y3)e3[y1,y2,y3]3\Lambda(y_{1},y_{2},y_{3})=\eta_{1}^{0}(y_{1},y_{2})\chi(y_{1},y_{2},y_{3})e_{3}\quad[y_{1},y_{2},y_{3}]^{\top}\in\mathbb{R}^{3}

and we consider

{ζ(t,y)=Λ(ζ(t,y)),ζ(0,y)=y3.\left\{\begin{array}[]{l}\zeta^{\prime}(t,y)=\Lambda(\zeta(t,y)),\\ \zeta(0,y)=y\in\mathbb{R}^{3}.\end{array}\right. (3.1)

Then

X0:=ζ(1,)X^{0}:=\zeta(1,\cdot) (3.2)

is a C1C^{1}-diffeomorphism such that

X0Idin3{[y1,y2,y3]𝒮×;y3(η(y1,y2),η+(y1,y2))}X^{0}\equiv\operatorname{Id}\quad\text{in}\ \mathbb{R}^{3}\setminus\left\{[y_{1},y_{2},y_{3}]^{\top}\in{\mathcal{S}}\times\mathbb{R}\ ;\ y_{3}\in(\eta_{-}(y_{1},y_{2}),\eta_{+}(y_{1},y_{2}))\right\}
X0(ΓS(0))=ΓS(η10),X^{0}(\Gamma_{S}(0))=\Gamma_{S}(\eta_{1}^{0}),
X0({[y1,y2,y3]𝒮×;y3(η(y1,y2),0)})={[y1,y2,y3]𝒮×;y3(η(y1,y2),η10(y1,y2))}.X^{0}\left(\left\{[y_{1},y_{2},y_{3}]^{\top}\in\mathcal{S}\times\mathbb{R}\ ;\ y_{3}\in(\eta_{-}(y_{1},y_{2}),0)\right\}\right)\\ =\left\{[y_{1},y_{2},y_{3}]^{\top}\in\mathcal{S}\times\mathbb{R}\ ;\ y_{3}\in(\eta_{-}(y_{1},y_{2}),\eta_{1}^{0}(y_{1},y_{2}))\right\}.

In particular, X0X^{0} is a C1C^{1}-diffeomorphism such that X0()=(η10)X^{0}(\mathcal{F})=\mathcal{F}(\eta_{1}^{0}) and such that X0=IdX^{0}=\operatorname{Id} on Γ0\Gamma_{0}.

We consider the characteristics XX associated with the fluid velocity v~\widetilde{v}:

{tX(t,y)=v~(t,X(t,y))(t>0),X(0,y)=X0(y),y¯.\begin{dcases}\partial_{t}X(t,y)=\widetilde{v}(t,X(t,y))\qquad(t>0),\\ X(0,y)=X^{0}(y),\quad y\in\overline{\mathcal{F}}.\end{dcases} (3.3)

Assume that XX is a C1C^{1}-diffeomorphism from ¯\overline{\mathcal{F}} onto (η(t))¯\overline{\mathcal{F}(\eta(t))} for all t(0,T).t\in(0,T). For each t(0,T),t\in(0,T), we denote by Y(t,)=[X(t,)]1Y(t,\cdot)=[X(t,\cdot)]^{-1} the inverse of X(t,)X(t,\cdot). We consider the following change of variables

ρ(t,y)=ρ~(t,X(t,y)),v(t,y)=v~(t,X(t,y)),ϑ(t,y)=ϑ~(t,X(t,y)),π=R0ρϑ+π0,\begin{array}[]{c}\rho(t,y)=\widetilde{\rho}(t,X(t,y)),\qquad v(t,y)=\widetilde{v}(t,X(t,y)),\\ \vartheta(t,y)=\widetilde{\vartheta}(t,X(t,y)),\qquad\pi=R_{0}\rho\vartheta+\pi_{0},\end{array} (3.4)

for (t,y)(0,T)×.(t,y)\in(0,T)\times\mathcal{F}. In particular,

ρ~(t,x)=ρ(t,Y(t,x)),v~(t,x)=v(t,Y(t,x)),ϑ~(t,x)=ϑ(t,Y(t,x)),\widetilde{\rho}(t,x)=\rho(t,Y(t,x)),\quad\widetilde{v}(t,x)=v(t,Y(t,x)),\quad\widetilde{\vartheta}(t,x)=\vartheta(t,Y(t,x)),

for (t,x)(0,T)×(η(t))(t,x)\in(0,T)\times\mathcal{F}(\eta(t)). We introduce the notation

𝔹X:=CofX,δX:=detX,𝔸X:=1δX𝔹X𝔹X,\mathbb{B}_{X}:=\operatorname{Cof}\nabla X,\quad\delta_{X}:=\det\nabla X,\quad\mathbb{A}_{X}:=\frac{1}{\delta_{X}}\mathbb{B}_{X}^{\top}\mathbb{B}_{X}, (3.5)
𝔹0:=𝔹X0,δ0:=δX0,𝔸0:=𝔸X0.\mathbb{B}^{0}:=\mathbb{B}_{X^{0}},\quad\delta^{0}:=\delta_{X^{0}},\quad\mathbb{A}^{0}:=\mathbb{A}_{X^{0}}. (3.6)

This change of variables transforms (1.3)–(1.8) into the following system for [ρ,v,ϑ,η]\left[\rho,v,\vartheta,\eta\right]^{\top}:

{tρ+ρ0δ0v:𝔹0=F1 in (0,T)×,ρ(0,)=ρ0 in ,\left\{\begin{array}[]{ll}\displaystyle\partial_{t}\rho+\frac{\rho^{0}}{\delta^{0}}\nabla v:\mathbb{B}^{0}={F}_{1}&\mbox{ in }(0,T)\times\mathcal{F},\\ \rho(0,\cdot)=\rho^{0}&\mbox{ in }\mathcal{F},\\ \end{array}\right. (3.7)
{tvv=F2 in (0,T)×,v=0 on (0,T)×Γ0,v=tηe3 on (0,T)×ΓS,v(0,)=v0 in ,\left\{\begin{array}[]{ll}\partial_{t}v-\mathcal{L}v={F}_{2}&\mbox{ in }(0,T)\times\mathcal{F},\\ v=0&\mbox{ on }(0,T)\times\Gamma_{0},\\ v=\partial_{t}\eta e_{3}&\mbox{ on }(0,T)\times\Gamma_{S},\\ v(0,\cdot)=v^{0}&\mbox{ in }\mathcal{F},\end{array}\right. (3.8)
{tϑκcvρ0δ0div(𝔸0ϑ)=F3 in (0,T)×,𝔸0ϑn=G on (0,T)×,ϑ(0,)=ϑ0 in ,\left\{\begin{array}[]{ll}\partial_{t}\vartheta-\dfrac{\kappa}{c_{v}\rho^{0}\delta^{0}}\operatorname{div}\left(\mathbb{A}^{0}\nabla\vartheta\right)={F}_{3}&\mbox{ in }(0,T)\times\mathcal{F},\\ \mathbb{A}^{0}\nabla\vartheta\cdot n={G}&\mbox{ on }(0,T)\times\partial\mathcal{F},\\ \vartheta(0,\cdot)=\vartheta^{0}&\mbox{ in }\mathcal{F},\end{array}\right. (3.9)
{ttη+Δs2ηΔstη=H in (0,T)×𝒮,η=ηnS=0 on (0,T)×𝒮,η(0,)=η10,tη(0,)=η20 in 𝒮,\left\{\begin{array}[]{ll}\partial_{tt}\eta+\Delta_{s}^{2}\eta-\Delta_{s}\partial_{t}\eta={H}&\mbox{ in }(0,T)\times\mathcal{S},\\ \eta=\nabla\eta\cdot n_{S}=0&\mbox{ on }(0,T)\times\partial\mathcal{S},\\ \eta(0,\cdot)=\eta_{1}^{0},\quad\partial_{t}\eta(0,\cdot)=\eta_{2}^{0}&\mbox{ in }\mathcal{S},\end{array}\right. (3.10)

where we have used the following notation

ρ0:=ρ~0X0,v0:=v~0X0,ϑ0:=ϑ~0X0,\rho^{0}:=\widetilde{\rho}^{0}\circ X^{0},\quad v^{0}:=\widetilde{v}^{0}\circ X^{0},\quad\vartheta^{0}:=\widetilde{\vartheta}^{0}\circ X^{0}, (3.11)
v=1ρ0δ0div𝕋0(v),𝕋0(v):=μv𝔸0+μ+αδ0𝔹0(v)𝔹0\mathcal{L}v=\frac{1}{\rho^{0}\delta^{0}}\operatorname{div}\mathbb{T}^{0}(v),\quad\mathbb{T}^{0}(v):=\mu\nabla v\mathbb{A}^{0}+\frac{\mu+\alpha}{\delta^{0}}\mathbb{B}^{0}(\nabla v)^{\top}\mathbb{B}^{0} (3.12)
F1(ρ,v,ϑ,η):=ρ0δ0v:𝔹0ρδXv:𝔹XF_{1}(\rho,v,\vartheta,\eta):=\frac{\rho^{0}}{\delta^{0}}\nabla v:\mathbb{B}^{0}-\frac{\rho}{\delta_{X}}\nabla v:\mathbb{B}_{X} (3.13)
F2(ρ,v,ϑ,η):=1ρ0δ0[(ρ0δ0ρδX)tv+μdiv(v(𝔸X𝔸0))+(μ+α)div[1δX𝔹X(v)𝔹X1δ0𝔹0(v)𝔹0]+R0𝔹X(ρϑ)]F_{2}(\rho,v,\vartheta,\eta):=\frac{1}{\rho^{0}\delta^{0}}\Bigg{[}\left(\rho^{0}\delta^{0}-\rho\delta_{X}\right)\partial_{t}v+\mu\operatorname{div}\left(\nabla v\left(\mathbb{A}_{X}-\mathbb{A}^{0}\right)\right)\\ +(\mu+\alpha)\operatorname{div}\left[\frac{1}{\delta_{X}}\mathbb{B}_{X}(\nabla v)^{\top}\mathbb{B}_{X}-\frac{1}{\delta^{0}}\mathbb{B}^{0}(\nabla v)^{\top}\mathbb{B}^{0}\right]+R_{0}\mathbb{B}_{X}\nabla(\rho\vartheta)\Bigg{]} (3.14)
F3(ρ,v,ϑ,η):=1cvρ0δ0[cv(ρ0δ0ρδX)tϑ+κdiv((𝔸X𝔸0)ϑ)+αδX(𝔹X:v)2+μ2δX|v𝔹X+𝔹Xv|2(R0ρϑ+π0)v:𝔹X]F_{3}(\rho,v,\vartheta,\eta):=\frac{1}{c_{v}\rho^{0}\delta^{0}}\Bigg{[}c_{v}\left(\rho^{0}\delta^{0}-\rho\delta_{X}\right)\partial_{t}\vartheta+\kappa\operatorname{div}\left(\left(\mathbb{A}_{X}-\mathbb{A}^{0}\right)\nabla\vartheta\right)\\ +\frac{\alpha}{\delta_{X}}\left(\mathbb{B}_{X}:\nabla v\right)^{2}+\frac{\mu}{2\delta_{X}}\left|\nabla v\mathbb{B}_{X}^{\top}+\mathbb{B}_{X}\nabla v^{\top}\right|^{2}-(R_{0}\rho\vartheta+\pi_{0})\nabla v:\mathbb{B}_{X}\Bigg{]} (3.15)
G(ρ,v,ϑ,η)=(𝔸0𝔸X)ϑn{G}(\rho,v,\vartheta,\eta)=\left(\mathbb{A}^{0}-\mathbb{A}_{X}\right)\nabla\vartheta\cdot n (3.16)
H(ρ,v,ϑ,η)=μδX(v𝔹X+𝔹Xv)[sη1]e3αδXv:𝔹X+R0ρϑ+π0.{H}(\rho,v,\vartheta,\eta)=-\frac{\mu}{\delta_{X}}\left(\nabla v\mathbb{B}_{X}^{\top}+\mathbb{B}_{X}\nabla v^{\top}\right)\begin{bmatrix}-\nabla_{s}\eta\\ 1\end{bmatrix}\cdot e_{3}-\frac{\alpha}{\delta_{X}}\nabla v:\mathbb{B}_{X}+R_{0}\rho\vartheta+\pi_{0}. (3.17)

The characteristics XX defined in (3.3) can now be written as

X(t,y)=X0(y)+0tv(r,y)dr,\displaystyle X(t,y)=X^{0}(y)+\int_{0}^{t}v(r,y)\ {\rm d}r,\quad (3.18)

for every yy\in\mathcal{F} and t0t\geqslant 0.

The hypotheses (1.12)–(1.15) on the initial conditions are transformed into the following conditions

ρ0W1,q(),min¯ρ0>0,\displaystyle\rho^{0}\in W^{1,q}(\mathcal{F}),\quad\min_{\overline{\mathcal{F}}}\rho^{0}>0, (3.19)
η10Bq,p2(21/p)(𝒮),η10=sη10nS=0 on 𝒮,\displaystyle\eta_{1}^{0}\in B^{2(2-1/p)}_{q,p}(\mathcal{S}),\quad\eta_{1}^{0}=\displaystyle\nabla_{s}\eta_{1}^{0}\cdot n_{S}=0\mbox{ on }\mathcal{S}, (3.20)
v0Bq,p2(11/p)()3,ϑ0Bq,p2(11/p)(),η20Bq,p2(11/p)(𝒮),\displaystyle v^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3},\quad\vartheta^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{F}),\quad\quad\eta_{2}^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{S}), (3.21)
v0=0 on Γ0,v0=η20e3 on ΓS,η20=0 on 𝒮,\displaystyle v^{0}=0\mbox{ on }\Gamma_{0},\quad v^{0}=\eta_{2}^{0}e_{3}\mbox{ on }\Gamma_{S},\quad\eta_{2}^{0}=0\mbox{ on }\partial\mathcal{S}, (3.22)
η20nS=0 on 𝒮and𝔸0ϑ0n=0onif1p+12q<12.\displaystyle\displaystyle\nabla\eta_{2}^{0}\cdot n_{S}=0\mbox{ on }\partial\mathcal{S}\quad\text{and}\quad\mathbb{A}^{0}\nabla\vartheta^{0}\cdot n=0\quad\mbox{on}\ \partial\mathcal{F}\quad\mbox{if}\quad\displaystyle\frac{1}{p}+\frac{1}{2q}<\frac{1}{2}. (3.23)

Here nn is the unit normal to \partial\mathcal{F} outward to .\mathcal{F}. The regularity properties in (3.19) and (3.21) can be obtained from (1.12), (1.13) by applying [30, Lemma 2.1]. Using the above change of variables, our main result in 1.1 can be rephrased as

Theorem 3.1.

Assume (p,q)(p,q) satisfies (1.11) and that [ρ0,v0,ϑ0,η10,η20][\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top} satisfies (3.19)–(3.23) and (1.18). Then there exists T>0T>0 such that the system (3.7)–(3.18) admits a unique strong solution

[ρ,v,ϑ,η]W1,p(0,T;W1,q())×(Wp,q1,2((0,T);)3×Wp,q1,2((0,T);)×Wp,q2,4((0,T);𝒮)\left[\rho,v,\vartheta,\eta\right]^{\top}\in W^{1,p}(0,T;W^{1,q}(\mathcal{F}))\times\left(W^{1,2}_{p,q}((0,T);\mathcal{F}\right)^{3}\times W^{1,2}_{p,q}((0,T);\mathcal{F})\times W^{2,4}_{p,q}((0,T);\mathcal{S})

Moreover,

min[0,T]ׯρ>0,Γ0ΓS(η(t))=(t[0,T]),\min_{[0,T]\times\overline{\mathcal{F}}}\rho>0,\quad\Gamma_{0}\cap\Gamma_{S}(\eta(t))=\emptyset\quad(t\in[0,T]),

and for all t[0,T]t\in[0,T], X(t,):(η(t))X(t,\cdot):\mathcal{F}\to\mathcal{F}(\eta(t)) is a C1C^{1}-diffeomorphism.

3.2. Maximal LpL^{p}-LqL^{q} regularity of a linear system.

The proof of 3.1 relies on the Banach fixed point theorem and on maximal LpL^{p}-LqL^{q} estimates of a linearized system. By replacing the nonlinear terms F1,F2,F3,G{F}_{1},{F}_{2},{F}_{3},{G} and H{H} in (3.7)–(3.10) by given source terms f1,f2,f3,gf_{1},f_{2},f_{3},g and hh we obtain the following linear system

{tρ+ρ0δ0v:𝔹0=f1 in (0,T)×,ρ(0,)=ρ^0 in ,\left\{\begin{array}[]{ll}\displaystyle\partial_{t}\rho+\frac{\rho^{0}}{\delta^{0}}\nabla v:\mathbb{B}^{0}={f}_{1}&\mbox{ in }(0,T)\times\mathcal{F},\\ \rho(0,\cdot)=\widehat{\rho}^{0}&\mbox{ in }\mathcal{F},\\ \end{array}\right. (3.24)
{tvv=f2 in (0,T)×,v=0 on (0,T)×ΓF,v=tηe3 on (0,T)×ΓS,v(0,)=v0 in ,\left\{\begin{array}[]{ll}\partial_{t}v-\mathcal{L}v={f}_{2}&\mbox{ in }(0,T)\times\mathcal{F},\\ v=0&\mbox{ on }(0,T)\times\Gamma_{F},\\ v=\partial_{t}\eta e_{3}&\mbox{ on }(0,T)\times\Gamma_{S},\\ v(0,\cdot)=v^{0}&\mbox{ in }\mathcal{F},\end{array}\right. (3.25)
{tϑκcvρ0δ0div(𝔸0ϑ)=f3 in (0,T)×,𝔸0ϑn=g on (0,T)×,ϑ(0,)=ϑ0 in ,\left\{\begin{array}[]{ll}\partial_{t}\vartheta-\dfrac{\kappa}{c_{v}\rho^{0}\delta^{0}}\operatorname{div}\left(\mathbb{A}^{0}\nabla\vartheta\right)={f}_{3}&\mbox{ in }(0,T)\times\mathcal{F},\\ \mathbb{A}^{0}\nabla\vartheta\cdot n={g}&\mbox{ on }(0,T)\times\partial\mathcal{F},\\ \vartheta(0,\cdot)=\vartheta^{0}&\mbox{ in }\mathcal{F},\end{array}\right. (3.26)
{ttη+Δs2ηΔstη=h in (0,T)×𝒮,η=ηnS=0 on (0,T)×𝒮,η(0,)=η^10,tη(0,)=η20 in 𝒮,\left\{\begin{array}[]{ll}\partial_{tt}\eta+\Delta_{s}^{2}\eta-\Delta_{s}\partial_{t}\eta={h}&\mbox{ in }(0,T)\times\mathcal{S},\\ \eta=\nabla\eta\cdot n_{S}=0&\mbox{ on }(0,T)\times\partial\mathcal{S},\\ \eta(0,\cdot)=\widehat{\eta}_{1}^{0},\quad\partial_{t}\eta(0,\cdot)=\eta_{2}^{0}&\mbox{ in }\mathcal{S},\end{array}\right. (3.27)

where 𝔸0,𝔹0,δ0\mathbb{A}^{0},\mathbb{B}^{0},\delta^{0} are defined in (3.6) and where \mathcal{L} is defined by (3.12). Note that we also modify the initial conditions in the above system with respect to (3.7)–(3.10) since ρ0\rho^{0} and η10\eta_{1}^{0} already appear in the coefficients of (3.24)–(3.27). In the next section, we will take

ρ^0=ρ0,η^10=η10\widehat{\rho}^{0}=\rho^{0},\quad\widehat{\eta}_{1}^{0}=\eta_{1}^{0}

but here we do not assume the above relation. In particular, we assume that ρ0\rho^{0} satisfies the second condition of (3.19) and that η10\eta_{1}^{0} satisfies (1.18) but we do not impose these hypotheses on ρ^0\widehat{\rho}^{0} and on η^10\widehat{\eta}_{1}^{0}.

We recall that (p,q)(p,q) satisfies (1.11) and to simplify, we assume throughout this section that

T(0,1].T\in(0,1].

This condition is only used to avoid the dependence in time of the constants in the estimates of this section.

We consider the subset of initial conditions

p,q={[ρ^0,v0,ϑ0,η^10,η20]W1,q()×Bq,p2(11/p)()3×Bq,p2(11/p)()×Bq,p2(21/p)(𝒮)×Bq,p2(11/p)(𝒮),v0=0 on Γ0,v0=η20e3 on ΓS,η^10=η^10nS=η20=0 on 𝒮,η20nS=0 on 𝒮and𝔸0ϑ0n=0 on if1p+12q<12},\mathcal{I}_{p,q}=\Bigg{\{}\left[\widehat{\rho}^{0},v^{0},\vartheta^{0},\widehat{\eta}_{1}^{0},\eta_{2}^{0}\right]^{\top}\in W^{1,q}(\mathcal{F})\times B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}\times B^{2(1-1/p)}_{q,p}(\mathcal{F})\times B^{2(2-1/p)}_{q,p}(\mathcal{S})\times B^{2(1-1/p)}_{q,p}(\mathcal{S}),\\ v^{0}=0\mbox{ on }\Gamma_{0},\quad v^{0}=\eta_{2}^{0}e_{3}\mbox{ on }\Gamma_{S},\quad\widehat{\eta}_{1}^{0}=\displaystyle\frac{\partial\widehat{\eta}_{1}^{0}}{\partial n_{S}}=\eta_{2}^{0}=0\mbox{ on }\partial\mathcal{S},\\ \displaystyle\frac{\partial\eta_{2}^{0}}{\partial n_{S}}=0\mbox{ on }\partial\mathcal{S}\quad\text{and}\quad\mathbb{A}^{0}\nabla\vartheta^{0}\cdot n=0\mbox{ on }\partial\mathcal{F}\quad\mbox{if}\quad\displaystyle\frac{1}{p}+\frac{1}{2q}<\frac{1}{2}\Bigg{\}}, (3.28)

endowed with the norm

[ρ^0,v0,ϑ0,η^10,η20]p,q:=ρ^0W1,q()+v0Bq,p2(11/p)()3+ϑ0Bq,p2(11/p)()+η^10Bq,p2(21/p)(𝒮)+η20Bq,p2(11/p)(𝒮).\left\|\left[\widehat{\rho}^{0},v^{0},\vartheta^{0},\widehat{\eta}_{1}^{0},\eta_{2}^{0}\right]^{\top}\right\|_{\mathcal{I}_{p,q}}:=\|\widehat{\rho}^{0}\|_{W^{1,q}(\mathcal{F})}+\|v^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}}+\|\vartheta^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})}\\ +\|\widehat{\eta}_{1}^{0}\|_{B^{2(2-1/p)}_{q,p}(\mathcal{S})}+\|\eta_{2}^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}.

We also consider the space T,p,q{\mathcal{R}}_{T,p,q} of the source terms in (3.24)–(3.27):

T,p,q={[f1,f2,f3,g,h];f1Lp(0,T,W1,q()),f2Lp(0,T;Lq())3,f3Lp(0,T;Lq()),gFp,q(11/q)/2(0,T;Lq())Lp(0,T;W11/q,q()),hLp(0,T;Lq(𝒮)), with g(0,)=0 if 1p+12q<12},\mathcal{R}_{T,p,q}=\Big{\{}\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\ ;\ f_{1}\in L^{p}(0,T,W^{1,q}(\mathcal{F})),f_{2}\in L^{p}(0,T;L^{q}(\mathcal{F}))^{3},\\ f_{3}\in L^{p}(0,T;L^{q}(\mathcal{F})),g\in F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F}))\cap L^{p}(0,T;W^{1-1/q,q}(\partial\mathcal{F})),\\ h\in L^{p}(0,T;L^{q}(\mathcal{S})),\mbox{ with }g(0,\cdot)=0\mbox{ if }\displaystyle\frac{1}{p}+\frac{1}{2q}<\frac{1}{2}\Big{\}}, (3.29)

with

[f1,f2,f3,g,h]T,p,q=f1Lp(0,T;W1,q())+f2Lp(0,T;Lq())3+f3Lp(0,T;Lq())+gFp,q(11/q)/2(0,T;Lq())Lp(0,T;W11/q,q())+hLp(0,T;Lq(𝒮)).\|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\|_{\mathcal{R}_{T,p,q}}=\|f_{1}\|_{L^{p}(0,T;W^{1,q}(\mathcal{F}))}+\|f_{2}\|_{L^{p}(0,T;L^{q}(\mathcal{F}))^{3}}+\|f_{3}\|_{L^{p}(0,T;L^{q}(\mathcal{F}))}\\ +\|g\|_{F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F}))\cap L^{p}(0,T;W^{1-1/q,q}(\partial\mathcal{F}))}+\|h\|_{L^{p}(0,T;L^{q}(\mathcal{S}))}.

Finally, the space 𝒲T,p,q{\mathcal{W}}_{T,p,q} of the solutions [ρ,u,ϑ,η][\rho,u,\vartheta,\eta]^{\top} of (3.24)–(3.27) is the Cartesian product:

𝒲T,p,q=W1,p(0,T;W1,q())×Wp,q1,2((0,T);)3×Wp,q1,2((0,T);)×Wp,q2,4((0,T);𝒮),\mathcal{W}_{T,p,q}=W^{1,p}(0,T;W^{1,q}(\mathcal{F}))\times W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}\times W^{1,2}_{p,q}((0,T);\mathcal{F})\times W^{2,4}_{p,q}((0,T);\mathcal{S}), (3.30)

with the norm

[ρ,u,ϑ,η]𝒲T,p,q:=ρW1,p(0,T;W1,q())+uWp,q1,2((0,T);)3+ϑWp,q1,2((0,T);)+ηWp,q2,4((0,T);𝒮).\|\left[\rho,u,\vartheta,\eta\right]^{\top}\|_{\mathcal{W}_{T,p,q}}:=\|\rho\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))}+\|u\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}}+\|\vartheta\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})}+\|\eta\|_{W^{2,4}_{p,q}((0,T);\mathcal{S})}.

With the above notation, we can state the main result of this section:

Theorem 3.2.

Assume (1.11) (3.19), (3.20) and (1.18). Then for any

[ρ^0,v0,ϑ0,η^10,η20]p,q,[f1,f2,f3,g,h]T,p,q,\left[\widehat{\rho}^{0},v^{0},\vartheta^{0},\widehat{\eta}_{1}^{0},\eta_{2}^{0}\right]^{\top}\in\mathcal{I}_{p,q},\quad\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\in\mathcal{R}_{T,p,q}, (3.31)

the system (3.24)–(3.27) admits a unique solution [ρ,v,ϑ,η]𝒲T,p,q[\rho,v,\vartheta,\eta]^{\top}\in\mathcal{W}_{T,p,q} and there exists a constant C>0C>0 depending on p,qp,q and independent of TT such that

[ρ,v,ϑ,η]𝒲T,p,qC([ρ^0,v0,ϑ0,η^10,η20]p,q+[f1,f2,f3,g,h]T,p,q).\left\|\left[\rho,v,\vartheta,\eta\right]^{\top}\right\|_{\mathcal{W}_{T,p,q}}\leqslant C\Big{(}\left\|\left[\widehat{\rho}^{0},v^{0},\vartheta^{0},\widehat{\eta}_{1}^{0},\eta_{2}^{0}\right]^{\top}\right\|_{\mathcal{I}_{p,q}}+\left\|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\right\|_{\mathcal{R}_{T,p,q}}\Big{)}. (3.32)

In order to prove the above result, we notice that the system (3.24)–(3.27) can be solved in “cascades”. Systems (3.26) and (3.27) can be solved independently. With the solution of system (3.27) we can solve the system (3.25) and then (3.24).

We first need the following result on the coefficients appearing in the system (3.24)–(3.27):

Lemma 3.3.

Assume (1.11) (3.19), (3.20) and (1.18). Then 𝔸0,𝔹0,δ0\mathbb{A}^{0},\mathbb{B}^{0},\delta^{0} defined in (3.6) satisfy

δ0>0,𝔸0=(𝔸0),1δ0W1,q(),𝔹0,𝔸0W1,q()9,\delta^{0}>0,\quad\mathbb{A}^{0}=(\mathbb{A}^{0})^{\top},\quad\frac{1}{\delta^{0}}\in W^{1,q}({\mathcal{F}}),\quad\mathbb{B}^{0},\mathbb{A}^{0}\in W^{1,q}({\mathcal{F}})^{9},

and there exists c0>0c^{0}>0 such that

𝔸0c0𝕀3in¯.\mathbb{A}^{0}\geqslant c^{0}\mathbb{I}_{3}\quad\text{in}\ \overline{\mathcal{F}}.
Proof.

The proof relies on the dependence of the solutions of (3.1) with respect to the initial conditions. Using that η10W2,q(2)\eta_{1}^{0}\in W^{2,q}(\mathbb{R}^{2}) for q>3q>3 and Sobolev embedding, we have that ΛCb1(3)\Lambda\in C^{1}_{b}(\mathbb{R}^{3}). In particular, from standard results (see, for instance, [2, p.116]), we have that ζC1(×3)\zeta\in C^{1}(\mathbb{R}\times\mathbb{R}^{3}) and by using the ordinary differential equation satisfied by the derivatives of ζ\zeta in space, we find that X0W2,q()3X^{0}\in W^{2,q}({\mathcal{F}})^{3} and X0\nabla X^{0} is invertible. This yields the result. ∎

We are now in a position to prove 3.2:

Proof of 3.2.

The proof is divided in several steps devoted to the resolution of each system.

Step 1: we show here that (3.27) admits a unique solution ηWp,q2,4((0,T)×𝒮)\eta\in W^{2,4}_{p,q}((0,T)\times\mathcal{S}) and that there exists a constant CC independent of TT such that

ηWp,q2,4((0,T);𝒮)+tηWp,q1,2((0,T);𝒮)C(η^10Bq,p2(21/p)(𝒮)+η20Bq,p2(11/p)(𝒮)+hLp(0,T;Lq(𝒮))).\|\eta\|_{W^{2,4}_{p,q}((0,T);\mathcal{S})}+\|\partial_{t}\eta\|_{W^{1,2}_{p,q}((0,T);\mathcal{S})}\leqslant C\left(\|\widehat{\eta}_{1}^{0}\|_{B^{2(2-1/p)}_{q,p}(\mathcal{S})}+\|\eta_{2}^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}+\|h\|_{L^{p}(0,T;L^{q}(\mathcal{S}))}\right). (3.33)

To prove this, we combine [14, Theorem 5.1] and [40, Theorem 4.2]. For the sake of clarity, we provide brief details about the proof. We first consider

𝒳S:=W02,q(𝒮)×Lq(𝒮),\mathcal{X}_{S}:=W^{2,q}_{0}(\mathcal{S})\times L^{q}(\mathcal{S}), (3.34)

and the operator ASA_{S} defined by

𝒟(AS)=(W4,q(𝒮)W02,q(𝒮))×W02,q(𝒮),AS=[0IdΔ2Δ].\mathcal{D}(A_{S})=\left(W^{4,q}(\mathcal{S})\cap W^{2,q}_{0}(\mathcal{S})\right)\times W^{2,q}_{0}(\mathcal{S}),\quad A_{S}=\begin{bmatrix}0&\operatorname{Id}\\ -\Delta^{2}&\Delta\end{bmatrix}. (3.35)

With the above notation, the system (3.27) can be written as

ddt[ηtη]=AS[ηtη]+[0h],[ηtη](0)=[η^10η20].\frac{d}{dt}\begin{bmatrix}\eta\\ \partial_{t}\eta\end{bmatrix}=A_{S}\begin{bmatrix}\eta\\ \partial_{t}\eta\end{bmatrix}+\begin{bmatrix}0\\ h\end{bmatrix},\qquad\begin{bmatrix}\eta\\ \partial_{t}\eta\end{bmatrix}(0)=\begin{bmatrix}\widehat{\eta}_{1}^{0}\\ \eta_{2}^{0}\end{bmatrix}.

Applying Theorem 5.1 in [14], we have that ASA_{S} is \mathcal{R}-sectorial in 𝒳S\mathcal{X}_{S} of angle β0>π/2\beta_{0}>\pi/2 (see Section 2). Thus the operator ASA_{S} has maximal regularity LpL^{p}-regularity in 𝒳S\mathcal{X}_{S} ([40, Theorem 4.2] or 2.4). More precisely, for every hLp(0,T;Lq())h\in L^{p}(0,T;L^{q}(\mathcal{F})) and for every (η^10,η20)(𝒳S,𝒟(AS))11/p,p,(\widehat{\eta}_{1}^{0},\eta_{2}^{0})\in(\mathcal{X}_{S},\mathcal{D}(A_{S}))_{1-1/p,p}, the system (3.27) admits a unique strong solution with

ηLp(0,T;W4,q(𝒮))W2,p(0,T;Lq(𝒮)).\eta\in L^{p}(0,T;W^{4,q}(\mathcal{S}))\cap W^{2,p}(0,T;L^{q}(\mathcal{S})).

In order to obtain the estimate (3.33) independent of TT, we proceed as [24, Proposition 2.2].

Step 2: we show now that the system (3.25) admits a unique solution vWp,q1,2((0,T);)3v\in W^{1,2}_{p,q}((0,T);\mathcal{F})^{3} and that there exists a constant C>0C>0 depending only on the geometry such that

vWp,q1,2((0,T);)3C(η^10Bq,p2(21/p)(𝒮)+η20Bq,p2(11/p)(𝒮)+v0Bq,p2(11/p)()3+hLp(0,T;Lq(𝒮))+f2Lp(0,T;Lq())3).\left\|v\right\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}}\leqslant C\Big{(}\|\widehat{\eta}_{1}^{0}\|_{B^{2(2-1/p)}_{q,p}(\mathcal{S})}+\|\eta_{2}^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}+\left\|v^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}}\\ +\|h\|_{L^{p}(0,T;L^{q}(\mathcal{S}))}+\|f_{2}\|_{L^{p}(0,T;L^{q}(\mathcal{F}))^{3}}\Big{)}. (3.36)

To do this, we are going to apply [12, Theorem 2.3] and for this, we first reduce the problem to the case of homogeneous boundary conditions.

Using that \mathcal{F} is a smooth domain, there exists an open bounded neighborhood 𝒮~\widetilde{\mathcal{S}} of 𝒮¯\overline{\mathcal{S}} in 2\mathbb{R}^{2}, ε~>0\widetilde{\varepsilon}>0 and η~:𝒮~\widetilde{\eta}:\widetilde{\mathcal{S}}\to\mathbb{R} smooth such that

(𝒮~×[ε~,ε~])={(s,η~(s)),s𝒮~}.\left(\widetilde{\mathcal{S}}\times[-\widetilde{\varepsilon},\widetilde{\varepsilon}]\right)\cap\partial\mathcal{F}=\left\{(s,\widetilde{\eta}(s)),\ s\in\widetilde{\mathcal{S}}\right\}. (3.37)

We consider χCc(3)\chi\in C_{c}^{\infty}(\mathbb{R}^{3}) such that

suppχ𝒮~×[ε~,ε~],χ1in𝒮×[ε~/2,ε~/2].\operatorname{supp}\chi\subset\widetilde{\mathcal{S}}\times[-\widetilde{\varepsilon},\widetilde{\varepsilon}],\quad\chi\equiv 1\quad\text{in}\ {\mathcal{S}}\times[-\widetilde{\varepsilon}/2,\widetilde{\varepsilon}/2].

Then we define

w(t,y1,y2,y3):=χ(y1,y2,y3)tη(t,y1,y2)e3((t,y1,y2,y3)(0,T)×3)w(t,y_{1},y_{2},y_{3}):=\chi(y_{1},y_{2},y_{3})\partial_{t}\eta(t,y_{1},y_{2})e_{3}\quad\left((t,y_{1},y_{2},y_{3})\in(0,T)\times\mathbb{R}^{3}\right) (3.38)

and we set u=vwu=v-w so that uu is the solution of

{tuu=f^2:=f2tww in (0,T)×,u=0 on (0,T)×,u(0,)=u0:=v0w(0,) in ,\begin{dcases}\partial_{t}u-\mathcal{L}u=\widehat{f}_{2}:=f_{2}-\partial_{t}w-\mathcal{L}w&\mbox{ in }(0,T)\times\mathcal{F},\\ u=0&\mbox{ on }(0,T)\times\partial\mathcal{F},\\ u(0,\cdot)=u^{0}:=v^{0}-w(0,\cdot)&\mbox{ in }\mathcal{F},\end{dcases} (3.39)

From 3.3 and (3.33), there exists a positive constant CC independent of TT such that

f^2Lp(0,T;Lq())3C(η^10Bq,p2(21/p)(𝒮)+η20Bq,p2(11/p)(𝒮)+hLp(0,T;Lq(𝒮))+f2Lp(0,T;Lq())3),\|\widehat{f}_{2}\|_{L^{p}(0,T;L^{q}(\mathcal{F}))^{3}}\leqslant C\Big{(}\|\widehat{\eta}_{1}^{0}\|_{B^{2(2-1/p)}_{q,p}(\mathcal{S})}+\|\eta_{2}^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}+\|h\|_{L^{p}(0,T;L^{q}(\mathcal{S}))}+\|f_{2}\|_{L^{p}(0,T;L^{q}(\mathcal{F}))^{3}}\Big{)},
u0Bq,p2(11/p)()3v0Bq,p2(11/p)()3+η20Bq,p2(11/p)(𝒮).\|u^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}}\leqslant\|v^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}}+\|\eta_{2}^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}.

Moreover, u0=0u^{0}=0 on .\partial\mathcal{F}. To obtain the result it remains to show that for u0Bq,p2(11/p)()3u^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3} with u0=0u^{0}=0 on \partial\mathcal{F} and for f^2Lp(0,T;Lq())3,\widehat{f}_{2}\in L^{p}(0,T;L^{q}(\mathcal{F}))^{3}, system (3.39) admits a unique strong solution in Wp,q1,2((0,T);)3W^{1,2}_{p,q}((0,T);\mathcal{F})^{3} with an estimate independent of TT. In order to do this, we are going to apply [12, Theorem 2.3].

Let us denote by 0(y,ξ)\mathcal{L}_{0}(y,\xi) the principal symbol of the operator \mathcal{L} defined by (3.12). Then we have

0(,ξ)=μρ0δ0(𝔸0ξξ)I3+μ+αρ0(δ0)2(𝔹0ξ)(𝔹0ξ).\mathcal{L}_{0}(\cdot,\xi)=\frac{\mu}{\rho^{0}\delta^{0}}(\mathbb{A}^{0}\xi\cdot\xi)I_{3}+\frac{\mu+\alpha}{\rho^{0}(\delta^{0})^{2}}(\mathbb{B}^{0}\xi)\otimes(\mathbb{B}^{0}\xi).

In particular, 0(,ξ)\mathcal{L}_{0}(\cdot,\xi) is symmetric and using (3.6) and (1.7), there exists c0c^{0} such that

0(y,ξ)aac0|a|2(y,a,ξ3,|ξ|=1).\mathcal{L}_{0}(y,\xi)a\cdot a\geqslant c^{0}|a|^{2}\quad(y\in\mathcal{F},a,\xi\in\mathbb{R}^{3},\ |\xi|=1). (3.40)

This shows condition (𝐄)\bf{(E)} (ellipticity of the interior symbol) of [12].

Since we are in the case of the Dirichlet boundary conditions, (3.40) yields the Lopatinskii–Shapiro condition (𝐋𝐒)\bf{(LS)}, see for instance, [34, Proposition 6.2.13 and Remark (i), p.270].

Finally, applying again 3.3 and using that q>3q>3, we can verify that (𝐒𝐃𝟏)\bf{(SD1)} and (𝐒𝐁𝟏)\bf{(SB1)} hold true. We can thus apply [12, Theorem 2.3] and deduce that the system (3.39) admits a unique solution uWp,q1,2((0,T);)3.u\in W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}. This yields that the system (3.25) admits a unique solution vWp,q1,2((0,T);)3.v\in W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}. In order to show that the estimate (3.36) holds with a constant independent of T,T, we can proceed as [24, Proposition 2.2].

Step 3: next we prove that the system (3.26) admits a unique strong solution ϑWp,q1,2((0,T);)\vartheta\in W^{1,2}_{p,q}((0,T);\mathcal{F}) and that there exists a constant C>0C>0, depending only on the geometry such that

ϑWp,q1,2((0,T);)C(ϑ0Bq,p2(11/p)()+f3Lp(0,T;Lq())+gFp,q(11/q)/2(0,T;Lq())+gLp(0,T;W11/q,q())).\|\vartheta\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})}\leqslant C\Big{(}\|\vartheta^{0}\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})}+\|f_{3}\|_{L^{p}(0,T;L^{q}(\mathcal{F}))}+\|g\|_{F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F}))}\\ +\|g\|_{L^{p}(0,T;W^{1-1/q,q}(\partial\mathcal{F}))}\Big{)}. (3.41)

As for the previous step, we are going to apply [12, Theorem 2.3]. The principal symbol associated with the operator ϑκcvρ0δ0div(𝔸0ϑ)\vartheta\mapsto-\dfrac{\kappa}{c_{v}\rho^{0}\delta^{0}}\operatorname{div}\left(\mathbb{A}^{0}\nabla\vartheta\right) is

a0(,ξ)=κcvρ0δ0𝔸0ξξa_{0}(\cdot,\xi)=\dfrac{\kappa}{c_{v}\rho^{0}\delta^{0}}\mathbb{A}^{0}\xi\cdot\xi

and from 3.3 it satisfies a0(,ξ)c1>0a_{0}(\cdot,\xi)\geqslant c^{1}>0 for ξ\xi such that |ξ|=1|\xi|=1. This shows condition (𝐄)\bf{(E)} (ellipticity of the interior symbol) of [12].

Due to Theorem 10.4 in [41, p.145], the above operator is properly elliptic and following Example 11.6 in [41, pp.160-161]), we see that the Lopatinskii–Shapiro condition (𝐋𝐒)\bf{(LS)} holds true.

Finally, applying again 3.3 and using that q>3q>3, we can verify that (𝐒𝐃𝟏)\bf{(SD1)} and (𝐒𝐁𝟏)\bf{(SB1)} hold true.

Thus all the conditions of [12, Theorem 2.3] are satisfied. Finally, to obtain the estimate (3.41) with constant independent of TT we can proceed as [24, Proposition 2.2].

Step 4: it only remains to prove the estimate for ρ.\rho. It follows from vWp,q1,2((0,T);)3v\in W^{1,2}_{p,q}((0,T);\mathcal{F})^{3} and 3.3 that the system (3.24) admits a unique solution ρW1,p(0,T;W1,q())\rho\in W^{1,p}(0,T;W^{1,q}(\mathcal{F})) and there exists a constant CC independent of TT such that

ρW1,p(0,T;W1,q())C(vWp,q1,2((0,T);)3+ρ^0W1,q()+f1Lp(0,T;W1,q())).\|\rho\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C\left(\|v\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}}+\|\widehat{\rho}^{0}\|_{W^{1,q}(\mathcal{F})}+\|f_{1}\|_{L^{p}(0,T;W^{1,q}(\mathcal{F}))}\right). (3.42)

Combining Step 1 to Step 4, we deduce the result. ∎

3.3. Proof of 3.1

Here, we show the local in time existence of solutions for (3.7)–(3.18). For this, we notice that a solution of (3.7)–(3.17) is a solution of (3.24)–(3.27) such that the source terms satisfy

[f1,f2,f3,g,h]=[F1,F2,F3,G,H],\left[f_{1},f_{2},f_{3},g,h\right]^{\top}=\left[F_{1},F_{2},F_{3},G,H\right]^{\top},

where F1,F2,F3,GF_{1},F_{2},F_{3},G and HH are given by (3.13)-(3.17). This suggests to prove 3.1 by showing that the following mapping admits a fixed point:

ΞT,R:T,RT,R,[f1,f2,f3,g,h][F1,F2,F2,G,H],\Xi_{T,R}:\mathcal{B}_{T,R}\longrightarrow\mathcal{B}_{T,R},\quad\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\longmapsto\left[F_{1},F_{2},F_{2},G,H\right]^{\top}, (3.43)

where

T,R={[f1,f2,f3,g,h]T,p,q;[f1,f2,f3,g,h]T,p,qR}\mathcal{B}_{T,R}=\left\{\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\in\mathcal{R}_{T,p,q}\ ;\ \|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\|_{\mathcal{R}_{T,p,q}}\leqslant R\right\}

(recall that T,p,q\mathcal{R}_{T,p,q} is defined by (3.29)) and where [ρ,v,ϑ,η]\left[\rho,v,\vartheta,\eta\right]^{\top} is the solution of (3.24)—(3.27) associated with [f1,f2,f3,g,h]\left[f_{1},f_{2},f_{3},g,h\right]^{\top} and with initial conditions [ρ0,v0,ϑ0,η10,η20]p,q\left[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}\right]^{\top}\in\mathcal{I}_{p,q}. More precisely, we take RR large enough so that

[ρ0,v0,ϑ0,η10,η20]p,qR,\|\left[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}\right]^{\top}\|_{\mathcal{I}_{p,q}}\leqslant R, (3.44)

and we assume (1.11) (3.19), (3.20) and (1.18) so that we can apply 3.2: the system (3.24)–(3.27) admits a unique solution (ρ,v,ϑ,η)𝒲T,p,q(\rho,v,\vartheta,\eta)\in\mathcal{W}_{T,p,q} and

[ρ,v,ϑ,η]𝒲T,p,qC([ρ0,v0,ϑ0,η10,η20]p,q+[f1,f2,f3,g,h]T,p,q).\left\|\left[\rho,v,\vartheta,\eta\right]^{\top}\right\|_{\mathcal{W}_{T,p,q}}\leqslant C\Big{(}\|\left[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}\right]^{\top}\|_{\mathcal{I}_{p,q}}+\|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\|_{\mathcal{R}_{T,p,q}}\Big{)}.

To prove 3.1, we need to show that, for TT small enough, the mapping ΞT,R\Xi_{T,R} is well-defined, that ΞT,R(T,R)T,R\Xi_{T,R}(\mathcal{B}_{T,R})\subset\mathcal{B}_{T,R} and ΞT,R|T,R{{\Xi}_{T,R}}|_{\mathcal{B}_{T,R}} is a strict contraction.

In this proof, we write CRC_{R} for any positive constant of the form C(1+RN)C(1+R^{N}) for NN\in\mathbb{N}, with CC a constant that only depends on the geometry and on the physical parameters, and in particular independent of TT. In particular the above inequality can be written as

ρW1,p(0,T;W1,q())+vWp,q1,2((0,T);)3+ϑWp,q1,2((0,T);)+ηWp,q2,4((0,T);𝒮)CR.\|\rho\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))}+\|v\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}}+\|\vartheta\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})}+\|\eta\|_{W^{2,4}_{p,q}((0,T);\mathcal{S})}\leqslant C_{R}. (3.45)

We are going to use several times that since q>3q>3, W1,q()W^{1,q}(\mathcal{F}) is an algebra and W1,q()L()W^{1,q}(\mathcal{F})\subset L^{\infty}(\mathcal{F}). We also have that W11q,q()L()W^{1-\frac{1}{q},q}(\partial\mathcal{F})\subset L^{\infty}(\partial\mathcal{F}).

We also recall the following elementary inequalities:

fLp(0,T)T1p1rfLr(0,T)(fLr(0,T))if r>p,\|f\|_{L^{p}(0,T)}\leqslant T^{\frac{1}{p}-\frac{1}{r}}\|f\|_{L^{r}(0,T)}\quad(f\in L^{r}(0,T))\quad\text{if }\ r>p, (3.46)
ff(0)L(0,T)T1pfW1,p(0,T)(fW1,p(0,T))if 1p+1p=1.\|f-f(0)\|_{L^{\infty}(0,T)}\leqslant T^{\frac{1}{p^{\prime}}}\|f\|_{W^{1,p}(0,T)}\quad(f\in W^{1,p}(0,T))\quad\text{if }\ \frac{1}{p}+\frac{1}{p^{\prime}}=1. (3.47)

In particular, we deduce from (3.45) and the above inequality

ρρ0L(0,T;W1,q())CRT1p,ρL(0,T;W1,q())CR.\|\rho-\rho^{0}\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}T^{\frac{1}{p^{\prime}}},\quad\|\rho\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}. (3.48)

The above estimate with (3.46) yields

ρLp(0,T;W1,q())CRT1p.\|\rho\|_{L^{p}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}T^{\frac{1}{p}}. (3.49)

Since 2<p<2<p<\infty, one has Bq,p2(11/p)()W1,q()B^{2(1-1/p)}_{q,p}(\mathcal{F})\hookrightarrow W^{1,q}(\mathcal{F}). Therefore, using (3.45) and (1.9), we obtain

vL(0,T;W1,q())3+ϑL(0,T;W1,q())CR.\|v\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))^{3}}+\|\vartheta\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}. (3.50)

Using (3.47) and (3.18), we deduce successively

XW1,p(0,T;W2,q())3CR,XX0L(0,T;W2,q())3CRT1p,XL(0,T;W2,q())3CR.\|X\|_{W^{1,p}(0,T;W^{2,q}(\mathcal{F}))^{3}}\leqslant C_{R},\quad\|X-X^{0}\|_{L^{\infty}(0,T;W^{2,q}(\mathcal{F}))^{3}}\leqslant C_{R}T^{\frac{1}{p^{\prime}}},\quad\|X\|_{L^{\infty}(0,T;W^{2,q}(\mathcal{F}))^{3}}\leqslant C_{R}. (3.51)

Since X0X^{0} is a C1C^{1}-diffeomorphism, we deduce from the above estimates that XX is a C1C^{1}-diffeomorphism for TT small enough. Moreover, by combining the above estimates with 3.3 and with (3.5), we also deduce

𝔹XW1,p(0,T;W1,q())9CR,𝔹X𝔹0L(0,T;W1,q())9CRT1p,𝔹XL(0,T;W1,q())9CR,\left\|\mathbb{B}_{X}\right\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))^{9}}\leqslant C_{R},\quad\left\|\mathbb{B}_{X}-\mathbb{B}^{0}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))^{9}}\leqslant C_{R}T^{\frac{1}{p^{\prime}}},\quad\left\|\mathbb{B}_{X}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))^{9}}\leqslant C_{R}, (3.52)
δXW1,p(0,T;W1,q())CR,δXδ0L(0,T;W1,q())CRT1p,δXL(0,T;W1,q())CR,\left\|\delta_{X}\right\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R},\quad\left\|\delta_{X}-\delta^{0}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}T^{\frac{1}{p^{\prime}}},\quad\left\|\delta_{X}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}, (3.53)

and in particular, there exists c0c_{0} depending on η10\eta_{1}^{0} such that for TT small enough,

δXc0>0.\delta_{X}\geqslant c_{0}>0. (3.54)

We thus deduce

1δXW1,p(0,T;W1,q())CR,1δX1δ0L(0,T;W1,q())CRT1p,1δXL(0,T;W1,q())CR.\left\|\frac{1}{\delta_{X}}\right\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R},\quad\left\|\frac{1}{\delta_{X}}-\frac{1}{\delta^{0}}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}T^{\frac{1}{p^{\prime}}},\quad\left\|\frac{1}{\delta_{X}}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant C_{R}. (3.55)

Using the above estimates and (3.5), we also obtain

𝔸XW1,p(0,T;W1,q())9CR,𝔸X𝔸0L(0,T;W1,q())9CRT1p,𝔸XL(0,T;W1,q())9CR.\left\|\mathbb{A}_{X}\right\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))^{9}}\leqslant C_{R},\quad\left\|\mathbb{A}_{X}-\mathbb{A}^{0}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))^{9}}\leqslant C_{R}T^{\frac{1}{p^{\prime}}},\quad\left\|\mathbb{A}_{X}\right\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))^{9}}\leqslant C_{R}. (3.56)

We are now in position to estimate the non linear terms in (3.13)-(3.17). From the above estimates, we deduce

F1(ρ,v,ϑ,η)Lp(0,T;W1,q())+F2(ρ,v,ϑ,η)Lp(0,T;Lq())3+F3(ρ,v,ϑ,η)Lp(0,T;Lq())CRT1p.\left\|F_{1}(\rho,v,\vartheta,\eta)\right\|_{L^{p}(0,T;W^{1,q}(\mathcal{F}))}+\left\|F_{2}(\rho,v,\vartheta,\eta)\right\|_{L^{p}(0,T;L^{q}(\mathcal{F}))^{3}}+\left\|F_{3}(\rho,v,\vartheta,\eta)\right\|_{L^{p}(0,T;L^{q}(\mathcal{F}))}\leqslant C_{R}T^{\frac{1}{p}}. (3.57)

By using the trace theorems, we also have

G(ρ,v,ϑ,η)Lp(0,T;W11/q,q())+H(ρ,v,ϑ,η)Lp(0,T;Lq(𝒮))CRT1p.\left\|G(\rho,v,\vartheta,\eta)\right\|_{L^{p}(0,T;W^{1-1/q,q}(\partial\mathcal{F}))}+\left\|H(\rho,v,\vartheta,\eta)\right\|_{L^{p}(0,T;L^{q}(\mathcal{S}))}\leqslant C_{R}T^{\frac{1}{p}}. (3.58)

It only remains to estimate GG given by (3.16) in Fp,q(11/q)/2(0,T;Lq())F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F})). First, using [12, Proposition 6.4], since ϑWp,q1,2((0,T);)\vartheta\in W^{1,2}_{p,q}((0,T);\mathcal{F}), we have that

i,j,ϑyjniFp,q(11/q)/2(0,T;Lq()),ϑyjniFp,q(11/q)/2(0,T;Lq())CR.\forall i,j,\quad\frac{\partial\vartheta}{\partial y_{j}}n_{i}\in F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F})),\quad\left\|\frac{\partial\vartheta}{\partial y_{j}}n_{i}\right\|_{F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F}))}\leqslant C_{R}.

Then we apply the general result [24, Proposition 2.7] with s=(11q)/2s=(1{-}\frac{1}{q})/2, U1=U3=Lq()U_{1}=U_{3}=L^{q}(\partial\mathcal{F}), U2=W11q,q()U_{2}=W^{1-\frac{1}{q},q}(\partial\mathcal{F}). Note that since 2<p<2<p<\infty, we have the condition s+1p<1s+\frac{1}{p}<1. From [24, Proposition 2.7], we deduce that for some positive constant δ\delta,

(𝔸0𝔸X)ϑnFp,q(11/q)/2(0,T;Lq())CTδ𝔸0𝔸XW1,p(0,T;W11q,q())i,jϑyjniFp,q(11/q)/2(0,T;Lq())CRTδ.\left\|\left(\mathbb{A}^{0}-\mathbb{A}_{X}\right)\nabla\vartheta\cdot n\right\|_{F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F}))}\\ \leqslant CT^{\delta}\|\mathbb{A}^{0}-\mathbb{A}_{X}\|_{W^{1,p}(0,T;W^{1-\frac{1}{q},q}(\partial\mathcal{F}))}\sum_{i,j}\left\|\frac{\partial\vartheta}{\partial y_{j}}n_{i}\right\|_{F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F}))}\leqslant C_{R}T^{\delta}. (3.59)

Combining (3.57), (3.58), (3.59), we deduce

ΞT,R(f1,f2,f3,g,h)T,p,qCRTδ\left\|\Xi_{T,R}(f_{1},f_{2},f_{3},g,h)\right\|_{\mathcal{R}_{T,p,q}}\leqslant C_{R}T^{\delta} (3.60)

for some power δ>0\delta>0. Thus for TT small enough, ΞT,R(T,R)T,R\Xi_{T,R}(\mathcal{B}_{T,R})\subset\mathcal{B}_{T,R}.

To show that ΞT,R|T,R{{\Xi}_{T,R}}|_{\mathcal{B}_{T,R}} is a strict contraction, we proceed similarly: we consider

[f1(i),f2(i),f3(i),g(i),h(i)]T,R,i=1,2\left[f_{1}^{(i)},f_{2}^{(i)},f_{3}^{(i)},g^{(i)},h^{(i)}\right]^{\top}\in\mathcal{B}_{T,R},\quad i=1,2

and we denote by [ρ(i),v(i),ϑ(i),η(i)]\left[\rho^{(i)},v^{(i)},\vartheta^{(i)},\eta^{(i)}\right]^{\top} the solution of (3.24)—(3.27) associated with

[f1(i),f2(i),f3(i),g(i),h(i)]T,p,qand[ρ0,v0,ϑ0,η10,η20]p,q.\left[f_{1}^{(i)},f_{2}^{(i)},f_{3}^{(i)},g^{(i)},h^{(i)}\right]^{\top}\in\mathcal{R}_{T,p,q}\quad\text{and}\quad\left[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}\right]^{\top}\in\mathcal{I}_{p,q}.

We also write

[f1,f2,f3,g,h]=[f1(1),f2(1),f3(1),g(1),h(1)][f1(2),f2(2),f3(2),g(2),h(2)],\left[f_{1},f_{2},f_{3},g,h\right]^{\top}=\left[f_{1}^{(1)},f_{2}^{(1)},f_{3}^{(1)},g^{(1)},h^{(1)}\right]^{\top}-\left[f_{1}^{(2)},f_{2}^{(2)},f_{3}^{(2)},g^{(2)},h^{(2)}\right]^{\top},
[ρ,v,ϑ,η]=[ρ(1),v(1),ϑ(1),η(1)][ρ(2),v(2),ϑ(2),η(2)].\left[\rho,v,\vartheta,\eta\right]^{\top}=\left[\rho^{(1)},v^{(1)},\vartheta^{(1)},\eta^{(1)}\right]^{\top}-\left[\rho^{(2)},v^{(2)},\vartheta^{(2)},\eta^{(2)}\right]^{\top}.

We can apply 3.2 and deduce that

ρW1,p(0,T;W1,q())+vWp,q1,2((0,T);)3+ϑWp,q1,2((0,T);)+ηWp,q2,4((0,T);𝒮)C[f1,f2,f3,g,h]T,p,q,\|\rho\|_{W^{1,p}(0,T;W^{1,q}(\mathcal{F}))}+\|v\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})^{3}}+\|\vartheta\|_{W^{1,2}_{p,q}((0,T);\mathcal{F})}+\|\eta\|_{W^{2,4}_{p,q}((0,T);\mathcal{S})}\\ \leqslant C\left\|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\right\|_{\mathcal{R}_{T,p,q}}, (3.61)

and since the initial conditions of [ρ,v,ϑ,η]\left[\rho,v,\vartheta,\eta\right]^{\top} are null, we can apply (3.47):

ρL(0,T;W1,q())CT1p[f1,f2,f3,g,h]T,p,q.\|\rho\|_{L^{\infty}(0,T;W^{1,q}(\mathcal{F}))}\leqslant CT^{\frac{1}{p^{\prime}}}\left\|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\right\|_{\mathcal{R}_{T,p,q}}. (3.62)

We deduce similarly that

X(1)X(2)L(0,T;W2,q())CT1p[f1,f2,f3,g,h]T,p,q,\|X^{(1)}-X^{(2)}\|_{L^{\infty}(0,T;W^{2,q}(\mathcal{F}))}\leqslant CT^{\frac{1}{p^{\prime}}}\|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\|_{\mathcal{R}_{T,p,q}}, (3.63)

and we obtain similar estimates for 𝔹X(1)𝔹X(2)\mathbb{B}_{X^{(1)}}-\mathbb{B}_{X^{(2)}}, 𝔸X(1)𝔸X(2)\mathbb{A}_{X^{(1)}}-\mathbb{A}_{X^{(2)}}, δX(1)δX(2)\delta_{X^{(1)}}-\delta_{X^{(2)}}. Proceeding as above, we deduce that F1,F2,F3,GF_{1},F_{2},F_{3},G and HH given by (3.13)-(3.17) satisfy

F1(ρ(1),v(1),ϑ(1),η(1))F1(ρ(2),v(2),ϑ(2),η(2))Lp(0,T;W1,q())+F2(ρ(1),v(1),ϑ(1),η(1))F2(ρ(2),v(2),ϑ(2),η(2))Lp(0,T;Lq())3+F3(ρ(1),v(1),ϑ(1),η(1))F3(ρ(2),v(2),ϑ(2),η(2))Lp(0,T;Lq())+G(ρ(1),v(1),ϑ(1),η(1))G(ρ(2),v(2),ϑ(2),η(2))Lp(0,T;W11/q,q())Fp,q(11/q)/2(0,T;Lq())+H(ρ(1),v(1),ϑ(1),η(1))H(ρ(2),v(2),ϑ(2),η(2))Lp(0,T;Lq(𝒮))CRTδ[f1,f2,f3,g,h]T,p,q\left\|F_{1}(\rho^{(1)},v^{(1)},\vartheta^{(1)},\eta^{(1)})-F_{1}(\rho^{(2)},v^{(2)},\vartheta^{(2)},\eta^{(2)})\right\|_{L^{p}(0,T;W^{1,q}(\mathcal{F}))}\\ +\left\|F_{2}(\rho^{(1)},v^{(1)},\vartheta^{(1)},\eta^{(1)})-F_{2}(\rho^{(2)},v^{(2)},\vartheta^{(2)},\eta^{(2)})\right\|_{L^{p}(0,T;L^{q}(\mathcal{F}))^{3}}\\ +\left\|F_{3}(\rho^{(1)},v^{(1)},\vartheta^{(1)},\eta^{(1)})-F_{3}(\rho^{(2)},v^{(2)},\vartheta^{(2)},\eta^{(2)})\right\|_{L^{p}(0,T;L^{q}(\mathcal{F}))}\\ +\left\|G(\rho^{(1)},v^{(1)},\vartheta^{(1)},\eta^{(1)})-G(\rho^{(2)},v^{(2)},\vartheta^{(2)},\eta^{(2)})\right\|_{L^{p}(0,T;W^{1-1/q,q}(\partial\mathcal{F}))\cap F^{(1-1/q)/2}_{p,q}(0,T;L^{q}(\partial\mathcal{F}))}\\ +\left\|H(\rho^{(1)},v^{(1)},\vartheta^{(1)},\eta^{(1)})-H(\rho^{(2)},v^{(2)},\vartheta^{(2)},\eta^{(2)})\right\|_{L^{p}(0,T;L^{q}(\mathcal{S}))}\leqslant C_{R}T^{\delta}\|\left[f_{1},f_{2},f_{3},g,h\right]^{\top}\|_{\mathcal{R}_{T,p,q}} (3.64)

for some positive constant δ\delta. Thus taking TT small enough, we deduce that ΞT,R|T,R{{\Xi}_{T,R}}|_{\mathcal{B}_{T,R}} is a strict contraction and this ends the proof of the theorem. ∎

4. Global in time existence

In this section we prove 1.2.

4.1. Change of variables and Linearization

As in the first part of this work, in order to show global existence in time we use a change of variables to write the system (1.3)–(1.8) in the fixed spatial domain \mathcal{F}. We consider the same transformation as in Section 3.1, that is XX is defined by (3.3). Note that (1.20) for RR small enough yields condition (1.18). However, we modify (3.4) since we linearize here the system around the constant steady state [ρ¯,0,ϑ¯,0],\left[\overline{\rho},0,\overline{\vartheta},0\right]^{\top}, with ρ¯,ϑ¯+\overline{\rho},\overline{\vartheta}\in\mathbb{R}_{+}^{*}:

ρ(t,y)=ρ~(t,X(t,y))ρ¯,v(t,y)=v~(t,X(t,y)),ϑ(t,y)=ϑ~(t,X(t,y))ϑ¯,\rho(t,y)=\widetilde{\rho}(t,X(t,y))-\overline{\rho},\qquad v(t,y)=\widetilde{v}(t,X(t,y)),\quad\vartheta(t,y)=\widetilde{\vartheta}(t,X(t,y))-\overline{\vartheta}, (4.1)

for (t,y)(0,T)×.(t,y)\in(0,T)\times\mathcal{F}. In particular,

ρ~(t,x)=ρ¯+ρ(t,Y(t,x)),v~(t,x)=v(t,Y(t,x)),ϑ~(t,x)=ϑ¯+ϑ(t,Y(t,x)),\widetilde{\rho}(t,x)=\overline{\rho}+\rho(t,Y(t,x)),\quad\widetilde{v}(t,x)=v(t,Y(t,x)),\quad\widetilde{\vartheta}(t,x)=\overline{\vartheta}+\vartheta(t,Y(t,x)), (4.2)

for (t,x)(0,T)×(η(t))(t,x)\in(0,T)\times\mathcal{F}(\eta(t)).

This change of variables transforms (1.3)–(1.8) into the following system for [ρ,v,ϑ,η]\left[\rho,v,\vartheta,\eta\right]^{\top}:

{tρ+ρ¯divv=F1(ρ,v,ϑ,η) in (0,)×,tv1ρ¯div𝕋¯(ρ,v,ϑ)=F2(ρ,v,ϑ,η) in (0,)×,tϑκ¯Δϑ=F3(ρ,v,ϑ,η) in (0,)×,ttη+Δs2ηΔstη=𝕋¯(ρ,v,ϑ)e3e3+H(ρ,v,ϑ,η) in (0,)×𝒮,\left\{\begin{array}[]{ll}\partial_{t}\rho+\overline{\rho}\operatorname{div}v=F_{1}(\rho,v,\vartheta,\eta)&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}v-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(\rho,v,\vartheta)=F_{2}(\rho,v,\vartheta,\eta)&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\vartheta-\overline{\kappa}\Delta\vartheta=F_{3}(\rho,v,\vartheta,\eta)&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{tt}\eta+\Delta_{s}^{2}\eta-\Delta_{s}\partial_{t}\eta=-\overline{\mathbb{T}}(\rho,v,\vartheta)e_{3}\cdot e_{3}+H(\rho,v,\vartheta,\eta)&\mbox{ in }(0,\infty)\times\mathcal{S},\end{array}\right. (4.3)
{v=0 on (0,)×ΓF,v=tηe3 on (0,)×ΓS,ϑn=G(ρ,v,ϑ,η) on (0,)×,η=sηnS=0 on (0,)×𝒮,\left\{\begin{array}[]{ll}v=0&\mbox{ on }(0,\infty)\times\Gamma_{F},\\ v=\partial_{t}\eta e_{3}&\mbox{ on }(0,\infty)\times\Gamma_{S},\\ \dfrac{\partial\vartheta}{\partial n}=G(\rho,v,\vartheta,\eta)&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \eta=\nabla_{s}\eta\cdot n_{S}=0&\mbox{ on }(0,\infty)\times\partial\mathcal{S},\end{array}\right. (4.4)
{η(0,)=η10,η(0,)=η20 in 𝒮,ρ(0,)=ρ0,v(0,)=v0,ϑ(0,)=ϑ0 in ,\left\{\begin{array}[]{ll}\eta(0,\cdot)=\eta_{1}^{0},\quad\partial\eta(0,\cdot)=\eta_{2}^{0}&\mbox{ in }\mathcal{S},\\ \rho(0,\cdot)=\rho^{0},\quad v(0,\cdot)=v^{0},\qquad\vartheta(0,\cdot)=\vartheta^{0}&\mbox{ in }\mathcal{F},\end{array}\right. (4.5)

where

𝕋¯(ρ,v,ϑ)=2μ𝔻v+(αdivvR0ϑ¯ρR0ρ¯ϑ)I3,\overline{\mathbb{T}}(\rho,v,\vartheta)=2\mu\mathbb{D}v+\left(\alpha\operatorname{div}v-R_{0}\overline{\vartheta}\rho-R_{0}\overline{\rho}\vartheta\right)I_{3}, (4.6)
κ¯=κcvρ¯\overline{\kappa}=\frac{\kappa}{c_{v}\overline{\rho}} (4.7)
ρ0=ρ~0X0ρ¯,v0=v~0X0,ϑ0=ϑ~0X0ϑ¯.\rho^{0}=\widetilde{\rho}^{0}\circ X^{0}-\overline{\rho},\qquad v^{0}=\widetilde{v}^{0}\circ X^{0},\qquad\vartheta^{0}=\widetilde{\vartheta}^{0}\circ X^{0}-\overline{\vartheta}. (4.8)

The nonlinear terms in (4.3)–(4.5) can be written as

F1(ρ,v,ϑ,η)=ρdivv(ρ+ρ¯)(1δX𝔹XI3):v,{F}_{1}(\rho,v,\vartheta,\eta)=-\rho\operatorname{div}v-(\rho+\overline{\rho})\Big{(}\frac{1}{\delta_{X}}\mathbb{B}_{X}-I_{3}\Big{)}:\nabla v, (4.9)
F2(ρ,v,ϑ,η)=1ρ¯[ρ¯(δX1)tvρδXtv+μdiv(v(𝔸XI3))+(μ+α)div(1δX𝔹X(v)𝔹X(v))+R0𝔹X(ρϑ)+R0(𝔹XI3)(ρ¯ϑ+ϑ¯ρ)]F_{2}(\rho,v,\vartheta,\eta)=\frac{1}{\overline{\rho}}\bigg{[}-\overline{\rho}(\delta_{X}-1)\partial_{t}v-\rho\delta_{X}\partial_{t}v+\mu\operatorname{div}\left(\nabla v(\mathbb{A}_{X}-I_{3})\right)\\ +(\mu+\alpha)\operatorname{div}\left(\frac{1}{\delta_{X}}\mathbb{B}_{X}(\nabla v)^{\top}\mathbb{B}_{X}-(\nabla v)^{\top}\right)+R_{0}\mathbb{B}_{X}\nabla(\rho\vartheta)+R_{0}(\mathbb{B}_{X}-I_{3})(\overline{\rho}\nabla\vartheta+\overline{\vartheta}\nabla\rho)\bigg{]} (4.10)
F3(ρ,v,ϑ,η)=1cvρ¯[cvδXρtϑcvρ¯(δX1)tϑR0(ρϑ+ρ¯ϑ+ϑ¯ρ)(𝔹X:v)+κdiv((𝔸XI3)ϑ)+αδX(𝔹X:v)2+2μδX|v𝔹X+𝔹Xv|2],{F}_{3}(\rho,v,\vartheta,\eta)=\frac{1}{c_{v}\overline{\rho}}\left[-c_{v}\delta_{X}\rho\partial_{t}\vartheta-c_{v}\overline{\rho}(\delta_{X}-1)\partial_{t}\vartheta-R_{0}\Big{(}\rho\vartheta+\overline{\rho}\vartheta+\overline{\vartheta}\rho\Big{)}\Big{(}\mathbb{B}_{X}:\nabla v\Big{)}\right.\\ \left.+\kappa\operatorname{div}\Big{(}(\mathbb{A}_{X}-I_{3})\nabla\vartheta\Big{)}+\frac{\alpha}{\delta_{X}}\left(\mathbb{B}_{X}:\nabla v\right)^{2}+\frac{2\mu}{\delta_{X}}\left|\nabla v\mathbb{B}_{X}^{\top}+\mathbb{B}_{X}\nabla v^{\top}\right|^{2}\right], (4.11)
G(ρ,v,ϑ,η)=(I3𝔸X)ϑn,{G}(\rho,v,\vartheta,\eta)=\left(I_{3}-\mathbb{A}_{X}\right)\nabla\vartheta\cdot n, (4.12)
H(ρ,v,ϑ,η)=μ[1δX(v𝔹X+𝔹Xv)[sη1]2μ𝔻(v)e3]e3α(1δX𝔹XI3):v+R0ρϑ,{H}(\rho,v,\vartheta,\eta)=-\mu\left[\frac{1}{\delta_{X}}\left(\nabla v\mathbb{B}_{X}^{\top}+\mathbb{B}_{X}\nabla v^{\top}\right)\begin{bmatrix}-\nabla_{s}\eta\\ 1\end{bmatrix}-2\mu\mathbb{D}(v)e_{3}\right]\cdot e_{3}\\ -\alpha\Big{(}\frac{1}{\delta_{X}}\mathbb{B}_{X}-I_{3}\Big{)}:\nabla v+R_{0}\rho\vartheta, (4.13)

where 𝔸X,\mathbb{A}_{X}, 𝔹X\mathbb{B}_{X} and δX\delta_{X} are defined in (3.5). The hypotheses (1.12)–(1.15) on the initial conditions are transformed into (3.20)–(3.23) and

ρ0W1,q(),min¯ρ0+ρ¯>0,.\rho^{0}\in W^{1,q}(\mathcal{F}),\quad\min_{\overline{\mathcal{F}}}\rho^{0}+\overline{\rho}>0,. (4.14)

Using the above change of variables, 1.2 can be reformulated as

Theorem 4.1.

Assume (p,q)(p,q) satisfies (1.11) and assume that ρ¯\overline{\rho} and ϑ¯\overline{\vartheta} are two given positive constants such that (1.19) holds. Then there exist β>0\beta>0 and R>0R>0 such that, for any [ρ0,v0,ϑ0,η10,η20]\left[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}\right]^{\top} satisfying (1.18), (4.14), (3.20)–(3.23) and

ρ0W1,q()+v0Bq,p2(11/p)()3+ϑ0Bq,p2(11/p)()+η10Bq,p2(21/p)(𝒮)+η20Bq,p2(11/p)(𝒮)R,\left\|\rho^{0}\right\|_{W^{1,q}(\mathcal{F})}+\left\|v^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}}+\left\|\vartheta^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})}\\ +\left\|\eta_{1}^{0}\right\|_{B^{2(2-1/p)}_{q,p}(\mathcal{S})}+\left\|\eta_{2}^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}\leqslant R,

the system (4.3)–(4.13) admits a unique strong solution [ρ,v,ϑ,η]\left[\rho,v,\vartheta,\eta\right]^{\top} in the class of functions satisfying

ρCb0([0,);W1,q()),ρWβ1,p(0,;Lq()),tρLβp(0,;W1,q()),\displaystyle\rho\in C^{0}_{b}([0,\infty);W^{1,q}(\mathcal{F})),\quad\nabla\rho\in{W^{1,p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))},\quad\partial_{t}\rho\in{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}, (4.15)
ϑCb0([0,);Bq,p2(11/p)()),ϑLβp(0,;W1,q()),tϑLβp(0,;Lq()),\displaystyle\vartheta\in C_{b}^{0}([0,\infty);B^{2(1-1/p)}_{q,p}(\mathcal{F})),\quad\nabla\vartheta\in L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F})),\quad\partial_{t}\vartheta\in{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))}, (4.16)
vWp,q,β1,2((0,);)3,tηWp,q,β1,2((0,);𝒮),\displaystyle v\in{W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{F})^{3}},\quad\partial_{t}\eta\in W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{S}), (4.17)
ηCb0([0,);Bq,p2(21/p)(𝒮)),ηLβp(0,;W4,q(𝒮))+L(0,;W4,q(𝒮)).\displaystyle\eta\in C^{0}_{b}([0,\infty);B^{2(2-1/p)}_{q,p}(\mathcal{S})),\quad\eta\in L^{p}_{\beta}(0,\infty;W^{4,q}(\mathcal{S}))+L^{\infty}(0,\infty;W^{4,q}(\mathcal{S})). (4.18)

Moreover,

min[0,)ׯρ+ρ¯>0,Γ0ΓS(η(t))=(t[0,)),\min_{[0,\infty)\times\overline{\mathcal{F}}}\rho+\overline{\rho}>0,\quad\Gamma_{0}\cap\Gamma_{S}(\eta(t))=\emptyset\quad(t\in[0,\infty)),

and for all t[0,)t\in[0,\infty), X(t,):(η(t))X(t,\cdot):\mathcal{F}\to\mathcal{F}(\eta(t)) is a C1C^{1}-diffeomorphism.

The proof of 4.1 relies on the Banach fixed point theorem and on the maximal LpLqL^{p}-L^{q} regularity of a linearized system over the time interval (0,).(0,\infty). In order to introduce the linearized system associated with (4.3)–(4.13), we introduce the following operator 𝒯:W02,q(𝒮)W2,q()3\mathcal{T}:W^{2,q}_{0}(\mathcal{S})\to W^{2,q}(\partial\mathcal{F})^{3} defined by

(𝒯η)(y)={η(s)e3 if y=(s,0)ΓS,0 if yΓ0.(\mathcal{T}\eta)(y)=\begin{dcases}\eta(s)e_{3}&\mbox{ if }y=(s,0)\in\Gamma_{S},\\ 0&\mbox{ if }y\in\Gamma_{0}.\end{dcases} (4.19)

We also write η1=η\eta_{1}=\eta and η2=tη\eta_{2}=\partial_{t}\eta and we consider the following system where we have replaced in (4.3)–(4.8), the nonlinearities F1,F2,F3,G,HF_{1},F_{2},F_{3},G,H by given source terms f1,f2,f3,g,hf_{1},f_{2},f_{3},g,h:

{tρ+ρ¯divv=f1 in (0,)×,tv1ρ¯div𝕋¯(ρ,v,ϑ)=f2 in (0,)×,tϑκ¯Δϑ=f3 in (0,)×,tη1η2=0 in (0,)×,tη2+Δs2η1Δsη2=𝕋¯(ρ,v,ϑ)e3e3+h in (0,)×𝒮,\left\{\begin{array}[]{ll}\partial_{t}\rho+\overline{\rho}\operatorname{div}v=f_{1}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}v-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(\rho,v,\vartheta)=f_{2}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\vartheta-\overline{\kappa}\Delta\vartheta=f_{3}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\eta_{1}-\eta_{2}=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\eta_{2}+\Delta_{s}^{2}\eta_{1}-\Delta_{s}\eta_{2}=-\overline{\mathbb{T}}(\rho,v,\vartheta)e_{3}\cdot e_{3}+h&\mbox{ in }(0,\infty)\times\mathcal{S},\end{array}\right. (4.20)
{v=𝒯η2 on (0,)×,ϑn=g on (0,)×,η1=sη1nS=0 on (0,)×𝒮,\left\{\begin{array}[]{ll}v=\mathcal{T}\eta_{2}&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \dfrac{\partial\vartheta}{\partial n}=g&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \eta_{1}=\nabla_{s}\eta_{1}\cdot n_{S}=0&\mbox{ on }(0,\infty)\times\partial\mathcal{S},\end{array}\right. (4.21)
{η1(0,)=η10,η2(0,)=η20 in 𝒮,ρ(0,)=ρ0,v(0,)=v0,ϑ(0,)=ϑ0 in ,\left\{\begin{array}[]{ll}\eta_{1}(0,\cdot)=\eta_{1}^{0},\quad\eta_{2}(0,\cdot)=\eta_{2}^{0}&\mbox{ in }\mathcal{S},\\ \rho(0,\cdot)=\rho^{0},\quad v(0,\cdot)=v^{0},\quad\vartheta(0,\cdot)=\vartheta^{0}&\mbox{ in }\mathcal{F},\end{array}\right. (4.22)

Our aim is to show that the linearized operator associated to the above linear system is \mathcal{R}-sectorial in a suitable function space.

4.2. The fluid-structure operator

Here we introduce the operator associated to the linear system (4.20)–(4.22). To this aim, we first define

𝒟(Av)={vW2,q()3;v=0 on },Av=μρ¯Δ+α+μρ¯div,\mathcal{D}(A_{\text{v}})=\Big{\{}v\in W^{2,q}(\mathcal{F})^{3}\ ;\ v=0\mbox{ on }\partial\mathcal{F}\Big{\}},\quad A_{\text{v}}=\frac{\mu}{\overline{\rho}}\Delta+\frac{\alpha+\mu}{\overline{\rho}}\nabla\operatorname{div}, (4.23)

and

𝒟(Aϑ)={ϑW2,q();ϑn=0 on },Aϑ=κ¯Δ.\mathcal{D}(A_{\vartheta})=\left\{\vartheta\in W^{2,q}(\mathcal{F})\ ;\ \frac{\partial\vartheta}{\partial n}=0\mbox{ on }\partial\mathcal{F}\right\},\quad A_{\vartheta}=\overline{\kappa}\Delta. (4.24)

From [35, Theorem 1.4], AvA_{\text{v}} is an isomorphism from 𝒟(Av)\mathcal{D}(A_{\text{v}}) onto Lq()3L^{q}(\mathcal{F})^{3} for any q(1,)q\in(1,\infty). Using trace properties, this allows us to introduce the operator

Dv(W02,q(𝒮);W2,q()3),D_{\text{v}}\in\mathcal{L}(W^{2,q}_{0}(\mathcal{S});W^{2,q}(\mathcal{F})^{3}), (4.25)

where w=Dvgw=D_{\text{v}}g is the solution to the system

{μρ¯Δwα+μρ¯(divw)=0 in ,w=𝒯g on .\begin{dcases}-\frac{\mu}{\overline{\rho}}\Delta w-\frac{\alpha+\mu}{\overline{\rho}}\nabla(\operatorname{div}w)=0&\mbox{ in }\mathcal{F},\\ w=\mathcal{T}g&\mbox{ on }\partial\mathcal{F}.\end{dcases} (4.26)

By a standard transposition method, the operator DvD_{\text{v}} can be extended as a bounded operator from Lq(𝒮)L^{q}(\mathcal{S}) to Lq()3.L^{q}(\mathcal{F})^{3}.

Using the above definitions and recalling the definitions (3.34), (3.35) of ASA_{S} and 𝒳S\mathcal{X}_{S}, we can write (4.20)–(4.22) as follows (in the case g=0g=0):

ddt[ρvϑη1η2]=𝒜FS[ρvϑη1η2]+[f1f2f30h],[ρvϑη1η2](0)=[ρ0v0ϑ0η10η20],\frac{d}{dt}\begin{bmatrix}\rho\\ v\\ \vartheta\\ \eta_{1}\\ \eta_{2}\end{bmatrix}=\mathcal{A}_{FS}\begin{bmatrix}\rho\\ v\\ \vartheta\\ \eta_{1}\\ \eta_{2}\end{bmatrix}+\begin{bmatrix}f_{1}\\ f_{2}\\ f_{3}\\ 0\\ h\end{bmatrix},\qquad\begin{bmatrix}\rho\\ v\\ \vartheta\\ \eta_{1}\\ \eta_{2}\end{bmatrix}(0)=\begin{bmatrix}\rho^{0}\\ v^{0}\\ \vartheta^{0}\\ \eta_{1}^{0}\\ \eta_{2}^{0}\end{bmatrix}, (4.27)

where 𝒜FS:𝒟(𝒜FS)𝒳\mathcal{A}_{FS}:\mathcal{D}(\mathcal{A}_{FS})\to\mathcal{X} is defined by

𝒳=W1,q()×Lq()3×Lq()×W02,q(𝒮)×Lq(𝒮),\mathcal{X}=W^{1,q}(\mathcal{F})\times L^{q}(\mathcal{F})^{3}\times L^{q}(\mathcal{F})\times W^{2,q}_{0}(\mathcal{S})\times L^{q}(\mathcal{S}), (4.28)
𝒟(𝒜FS)={[ρ,v,ϑ,η1,η2]W1,q()×W2,q()3×𝒟(Aϑ)×𝒟(AS);vDvη2𝒟(Av)},\mathcal{D}(\mathcal{A}_{FS})=\Big{\{}[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\in W^{1,q}(\mathcal{F})\times W^{2,q}(\mathcal{F})^{3}\times\mathcal{D}(A_{\vartheta})\times\mathcal{D}(A_{S})\ ;\ v-D_{\textnormal{v}}\eta_{2}\in\mathcal{D}(A_{\textnormal{v}})\Big{\}}, (4.29)

and

𝒜FS=𝒜FS0+FS,\mathcal{A}_{FS}=\mathcal{A}^{0}_{FS}+\mathcal{B}_{FS},

with

𝒜FS0=[ρvϑη1η2]=[ρ¯divvAv(vDvη2)Aϑϑη2Δs2η1+Δsη2]andFS[ρvϑη1η2]=[0R0ϑ¯ρ¯ρR0ϑ00𝕋¯(ρ,v,ϑ)e3e3].\mathcal{A}_{FS}^{0}=\begin{bmatrix}\rho\\ v\\ \vartheta\\ \eta_{1}\\ \eta_{2}\end{bmatrix}=\begin{bmatrix}-\overline{\rho}\operatorname{div}v\\ A_{\textnormal{v}}(v-D_{\textnormal{v}}\eta_{2})\\ A_{\vartheta}\vartheta\\ \eta_{2}\\ -\Delta_{s}^{2}\eta_{1}+\Delta_{s}\eta_{2}\end{bmatrix}\quad\text{and}\quad\mathcal{B}_{FS}\begin{bmatrix}\rho\\ v\\ \vartheta\\ \eta_{1}\\ \eta_{2}\end{bmatrix}=\begin{bmatrix}0\\ -\dfrac{R_{0}\overline{\vartheta}}{\overline{\rho}}\nabla\rho-R_{0}\nabla\vartheta\\ 0\\ 0\\ -\overline{\mathbb{T}}(\rho,v,\vartheta)e_{3}\cdot e_{3}\end{bmatrix}. (4.30)

We recall that the definition of a \mathcal{R}-sectorial operator is given in 2.2. We now prove the following theorem :

Theorem 4.2.

Let 1<q<.1<q<\infty. Then there exists γ>0\gamma>0 such that 𝒜FSγ\mathcal{A}_{FS}-\gamma is an \mathcal{R}-sectorial operator in 𝒳\mathcal{X} of angle β>π/2.\beta>\pi/2.

Proof.

In order to prove the theorem, we first combine [16, Theorem 2.5], [11, Theorem 8.2] and [14, Theorem 5.1]: there exist γ>0\gamma>0 and β>π/2\beta>\pi/2 such that the operators AvγA_{\textnormal{v}}-\gamma, AϑγA_{\vartheta}-\gamma and ASγA_{S}-\gamma are \mathcal{R}-sectorial operators of angle β\beta.

Second, standard calculation shows that for λγ+Σβ\lambda\in\gamma+\Sigma_{\beta} (see (2.2)),

λ(λI𝒜FS0)1=[Idρ¯div(λIAv)10ρ¯divAv(λIAv)1Dv~(λIAS)10λ(λIAv)10Av(λIAv)1Dv~λ(λIAS)100λ(λIAϑ)10000λ(λIAS)1],\lambda(\lambda I-\mathcal{A}^{0}_{FS})^{-1}=\begin{bmatrix}\operatorname{Id}&-\overline{\rho}\operatorname{div}(\lambda I-A_{\textnormal{v}})^{-1}&0&\overline{\rho}\operatorname{div}A_{\textnormal{v}}(\lambda I-A_{\textnormal{v}})^{-1}\widetilde{D_{\textnormal{v}}}(\lambda I-A_{S})^{-1}\\ 0&\lambda(\lambda I-A_{\textnormal{v}})^{-1}&0&-A_{\textnormal{v}}(\lambda I-A_{\textnormal{v}})^{-1}\widetilde{D_{\textnormal{v}}}\lambda(\lambda I-A_{S})^{-1}\\ 0&0&\lambda(\lambda I-A_{\vartheta})^{-1}&0\\ 0&0&0&\lambda(\lambda I-A_{S})^{-1}\end{bmatrix},

where Dv~[η1,η2]=Dvη2.\displaystyle\widetilde{D_{\textnormal{v}}}[\eta_{1},\eta_{2}]^{\top}=D_{\textnormal{v}}\eta_{2}. Using the properties of \mathcal{R}-boundedness recalled in Section 2, we deduce that 𝒜FS0γ\mathcal{A}_{FS}^{0}-\gamma is \mathcal{R}-sectorial operator in 𝒳\mathcal{X} of angle β.\beta. Note that in instance, we can write

divAv(λIAv)1Dv~(λIAS)1=divDv~(λIAS)1+div(λIAv)1Dv~λ(λIAS)1\operatorname{div}A_{\textnormal{v}}(\lambda I-A_{\textnormal{v}})^{-1}\widetilde{D_{\textnormal{v}}}(\lambda I-A_{S})^{-1}=-\operatorname{div}\widetilde{D_{\textnormal{v}}}(\lambda I-A_{S})^{-1}+\operatorname{div}(\lambda I-A_{\textnormal{v}})^{-1}\widetilde{D_{\textnormal{v}}}\lambda(\lambda I-A_{S})^{-1}

and then use that Dv(W02,q(𝒮);W2,q()3)(L0q(𝒮);Lq()3).D_{\text{v}}\in\mathcal{L}(W^{2,q}_{0}(\mathcal{S});W^{2,q}(\mathcal{F})^{3})\cap\mathcal{L}(L^{q}_{0}(\mathcal{S});L^{q}(\mathcal{F})^{3}).

Next, using trace results, for s(1/q,1)s\in(1/q,1) there exists a constant CC such that

FS[ρ,v,ϑ,η1,η2]𝒳C(ρW1,q()+vW1+s,q()3+ϑW1+s,q())[ρ,v,ϑ,η1,η2]𝒟(𝒜FS).\left\|\mathcal{B}_{FS}[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\right\|_{\mathcal{X}}\leqslant C\left(\left\|\rho\right\|_{W^{1,q}(\mathcal{F})}+\left\|v\right\|_{W^{1+s,q}(\mathcal{F})^{3}}+\left\|\vartheta\right\|_{W^{1+s,q}(\mathcal{F})}\right)\quad[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\in\mathcal{D}(\mathcal{A}_{FS}).

Since the embedding W1+s,q()W2,q()W^{1+s,q}(\mathcal{F})\hookrightarrow W^{2,q}(\mathcal{F}) is compact for s(1/q,1),s\in(1/q,1), for any ε>0\varepsilon>0 there exists C(ε)>0C(\varepsilon)>0 such that

FS[ρ,v,ϑ,η1,η2]𝒳ε𝒜FS0[ρ,v,ϑ,η1,η2]𝒳+C(ε)[ρ,v,ϑ,η1,η2]𝒳.\left\|\mathcal{B}_{FS}[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\right\|_{\mathcal{X}}\leqslant\varepsilon\left\|\mathcal{A}^{0}_{FS}[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\right\|_{\mathcal{X}}+C(\varepsilon)\left\|[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\right\|_{\mathcal{X}}. (4.31)

Finally using 2.5 we conclude the proof of the theorem. ∎

4.3. Exponential stability of the fluid-structure semigroup

The aim of this subsection is to show that the operator 𝒜FS\mathcal{A}_{FS} generates an analytic semigroup of negative type in the following subspace of 𝒳:\mathcal{X}:

𝒳m={[f1,f2,f3,h1,h2]𝒳;f1dy+ρ¯𝒮h1ds=0,f3dy=0}.\mathcal{X}_{\textnormal{m}}=\left\{[f_{1},f_{2},f_{3},h_{1},h_{2}]^{\top}\in\mathcal{X}\ ;\ \int_{\mathcal{F}}f_{1}{\rm d}y\;+\overline{\rho}\int_{\mathcal{S}}h_{1}{\rm d}s=0,\;\int_{\mathcal{F}}f_{3}{\rm d}y=0\right\}. (4.32)

We can verify that 𝒳m\mathcal{X}_{\textnormal{m}} is invariant under (et𝒜FS)t0\left(e^{t\mathcal{A}_{FS}}\right)_{t\geqslant 0}. Therefore we can consider the restriction of 𝒜FS\mathcal{A}_{FS} to the domain 𝒟(𝒜FS)𝒳m\mathcal{D}(\mathcal{A}_{FS})\cap\mathcal{X}_{\textnormal{m}} ([39, Definition 2.4.1]). For this operator, we have the following result:

Theorem 4.3.

Let 1<q<.1<q<\infty. The part of 𝒜FS\mathcal{A}_{FS} in 𝒳m\mathcal{X}_{\textnormal{m}} generates an exponentially stable semigroup (et𝒜FS)t0\left(e^{t\mathcal{A}_{FS}}\right)_{t\geqslant 0} on 𝒳m:\mathcal{X}_{\textnormal{m}}: there exists constants C>0C>0 and β0>0\beta_{0}>0 such that

et𝒜FS[ρ0,v0,ϑ0,η10,η20]𝒳Ceβ0t[ρ0,v0,ϑ0,η10,η20]𝒳,(t0),\left\|e^{t\mathcal{A}_{FS}}[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\right\|_{\mathcal{X}}\leqslant Ce^{-\beta_{0}t}\left\|[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\right\|_{\mathcal{X}},\qquad(t\geqslant 0), (4.33)

for all [ρ0,v0,ϑ0,η10,η20]𝒳m.[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\in\mathcal{X}_{\textnormal{m}}.

To show the above theorem, it sufficient to show that +ρ(𝒜FS|𝒟(𝒜FS)𝒳m).\mathbb{C}^{+}\subset\rho({\mathcal{A}_{FS}}_{|\mathcal{D}(\mathcal{A}_{FS})\cap\mathcal{X}_{\textnormal{m}}}). We thus consider the following resolvent problem

{λρ+ρ¯divv=f1 in ,λv1ρ¯div𝕋¯(ρ,v,ϑ)=f2 in ,λϑκ¯Δϑ=f3 in ,v=𝒯η2,ϑn=0 on ,λη1η2=h1 in 𝒮,λη2+Δs2η1Δsη2=𝕋¯(ρ,v,ϑ)e3e3+h2 in 𝒮,η1=sη1nS=η2=0 on 𝒮.\begin{dcases}\lambda\rho+\overline{\rho}\operatorname{div}v=f_{1}&\mbox{ in }\mathcal{F},\\ \lambda v-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(\rho,v,\vartheta)=f_{2}&\mbox{ in }\mathcal{F},\\ \lambda\vartheta-\overline{\kappa}\Delta\vartheta=f_{3}&\mbox{ in }\mathcal{F},\\ v=\mathcal{T}\eta_{2},\quad\frac{\partial\vartheta}{\partial n}=0&\mbox{ on }\partial\mathcal{F},\\ \lambda\eta_{1}-\eta_{2}=h_{1}&\mbox{ in }\mathcal{S},\\ \lambda\eta_{2}+\Delta_{s}^{2}\eta_{1}-\Delta_{s}\eta_{2}=-\overline{\mathbb{T}}(\rho,v,\vartheta)e_{3}\cdot e_{3}+h_{2}&\mbox{ in }\mathcal{S},\\ \eta_{1}=\nabla_{s}\eta_{1}\cdot n_{S}=\eta_{2}=0&\mbox{ on }\partial\mathcal{S}.\end{dcases} (4.34)
Remark 4.4.

If λ=0,\lambda=0, integrating the first and third equation of (4.34) and using the boundary conditions of vv and ϑ\vartheta we obtain

f1dy+ρ¯𝒮h1ds=0 and f3dy=0.\int_{\mathcal{F}}f_{1}{\rm d}y\;+\overline{\rho}\int_{\mathcal{S}}h_{1}{\rm d}s=0\mbox{ and }\;\int_{\mathcal{F}}f_{3}{\rm d}y=0.

Therefore, in order to study exponential stability of the semigroup it is necessary to consider the space 𝒳m\mathcal{X}_{\textnormal{m}} instead of 𝒳.\mathcal{X}.

Proof.

Assume λ+\lambda\in\mathbb{C}^{+} and [f1,f2,f3,h1,h2]𝒳m.[f_{1},f_{2},f_{3},h_{1},h_{2}]^{\top}\in\mathcal{X}_{\textnormal{m}}. We need to show that the system (4.34) admits a unique solution [ρ,v,ϑ,η1,η2]𝒟(𝒜FS)𝒳m[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\in\mathcal{D}(\mathcal{A}_{FS})\cap\mathcal{X}_{\textnormal{m}} together with an estimate

[ρ,v,ϑ,η1,η2]𝒟(𝒜FS)C[f1,f2,f3,h1,h2]𝒳.\left\|[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\right\|_{\mathcal{D}(\mathcal{A}_{FS})}\leqslant C\left\|[f_{1},f_{2},f_{3},h_{1},h_{2}]^{\top}\right\|_{\mathcal{X}}.

The proof is divided into several parts.

Step 1: Uniqueness. Let us assume that [ρ,v,ϑ,η1,η2]𝒟(𝒜FS)𝒳m[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}\in\mathcal{D}(\mathcal{A}_{FS})\cap\mathcal{X}_{\textnormal{m}} solves the system (4.34) with [f1,f2,f3,h1,h2]=0[f_{1},f_{2},f_{3},h_{1},h_{2}]^{\top}=0. We notice that

[ρ,v,ϑ,η1,η2]W1,2()×W2,2()3×W2,2()×W4,2(𝒮)×W2,2(𝒮).\left[\rho,v,\vartheta,\eta_{1},\eta_{2}\right]^{\top}\in W^{1,2}(\mathcal{F})\times W^{2,2}(\mathcal{F})^{3}\times W^{2,2}(\mathcal{F})\times W^{4,2}(\mathcal{S})\times W^{2,2}(\mathcal{S}). (4.35)

If q2q\geqslant 2 then it is a consequence of Hölder’s inequality. Else, 1<q<21<q<2 and we take λ0ρ(𝒜FS)\lambda_{0}\in\rho(\mathcal{A}_{FS}) to rewrite (4.34) as

(λ0𝒜FS)[ρ,v,ϑ,η1,η2]=(λ0λ)[ρ,v,ϑ,η1,η2].\left(\lambda_{0}-\mathcal{A}_{FS}\right)[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}=\left(\lambda_{0}-\lambda\right)[\rho,v,\vartheta,\eta_{1},\eta_{2}]^{\top}.

Since W2,q()L2()W^{2,q}(\mathcal{F})\hookrightarrow L^{2}(\mathcal{F}) and W2,q(𝒮)L2(𝒮),W^{2,q}(\mathcal{S})\hookrightarrow L^{2}(\mathcal{S}), we deduce (4.35) from the the invertibility of the operator (λ0𝒜FS).(\lambda_{0}-\mathcal{A}_{FS}).

Multiplying (4.34)3\eqref{eq:resolvent-0}_{3} by ϑ\vartheta, we obtain after integration by parts

λ|ϑ|2dy+κ¯|ϑ|2dy=0.\lambda\int_{\mathcal{F}}|\vartheta|^{2}\ {\rm d}y+\overline{\kappa}\int_{\mathcal{F}}|\nabla\vartheta|^{2}\ {\rm d}y=0.

Since Reλ0\operatorname{Re}\lambda\geqslant 0 and ϑdy=0,\displaystyle\int_{\mathcal{F}}\vartheta\ {\rm d}y=0, we obtain ϑ=0.\vartheta=0.

Next, multiplying (4.34)2\eqref{eq:resolvent-0}_{2} by v,v, (4.34)6\eqref{eq:resolvent-0}_{6} by η2,\eta_{2}, after integration by parts and taking the real part, we deduce

R0ϑ¯ρ¯(Reλ)|ρ|2dy+ρ¯(Reλ)|v|2dy+2μ|𝔻v|2dy+α(divv)2dy+Reλ𝒮|η2|2ds+Reλ𝒮|Δsη1|2+𝒮|sη2|2=0.\frac{R_{0}\overline{\vartheta}}{\overline{\rho}}(\operatorname{Re}\lambda)\int_{\mathcal{F}}|\rho|^{2}\ {\rm d}y+\overline{\rho}(\operatorname{Re}\lambda)\int_{\mathcal{F}}|v|^{2}\ {\rm d}y+2\mu\int_{\mathcal{F}}|\mathbb{D}v|^{2}\ {\rm d}y+\alpha\int_{\mathcal{F}}(\operatorname{div}v)^{2}\ {\rm d}y\\ +\operatorname{Re}\lambda\int_{\mathcal{S}}|\eta_{2}|^{2}\ {\rm d}s+\operatorname{Re}\lambda\int_{\mathcal{S}}|\Delta_{s}\eta_{1}|^{2}+\int_{\mathcal{S}}|\nabla_{s}\eta_{2}|^{2}=0.

Since Reλ0,\operatorname{Re}\lambda\geqslant 0, using (1.7)\eqref{visco-relation} and using the boundary conditions we obtain v=η2=0v=\eta_{2}=0 and that ρ\rho is a constant. Using that [ρ,v,ϑ,η1,η2]𝒳m\left[\rho,v,\vartheta,\eta_{1},\eta_{2}\right]^{\top}\in\mathcal{X}_{\textnormal{m}} we deduce that η1\eta_{1} solves

{Δs2η1+R0ϑ¯ρ¯||𝒮η1ds=0 in 𝒮,η1=sη1nS=0 on 𝒮.\begin{dcases}\Delta_{s}^{2}\eta_{1}+\frac{R_{0}\overline{\vartheta}\overline{\rho}}{|\mathcal{F}|}\int_{\mathcal{S}}\eta_{1}{\rm d}s=0&\mbox{ in }\mathcal{S},\\ \eta_{1}=\nabla_{s}\eta_{1}\cdot n_{S}=0&\mbox{ on }\partial\mathcal{S}.\end{dcases} (4.36)

Multiplying the first equation of the above system by η1\eta_{1} and integrating by parts, we deduce that η1=0\eta_{1}=0 and that ρ=0\rho=0.

Step 2. Existence for λ=0\lambda=0. We consider the system (4.34) with λ=0.\lambda=0. It can be written as follows

η2=h1in𝒮,\eta_{2}=-h_{1}\ \text{in}\ \mathcal{S},
κ¯Δϑ=f3 in ,ϑn=0 on ,ϑ𝑑y=0,-\overline{\kappa}\Delta\vartheta=f_{3}\mbox{ in }\mathcal{F},\quad\frac{\partial\vartheta}{\partial n}=0\mbox{ on }\partial\mathcal{F},\quad\int_{\mathcal{F}}\vartheta\ dy=0,
{μΔv+R0ϑ¯ρ=ρ¯f2+α+μρ¯f1R0ρ¯ϑ in ,divv=1ρ¯f1 in ,v=𝒯h1 in ,\begin{dcases}-\mu\Delta v+R_{0}\overline{\vartheta}\nabla\rho=\overline{\rho}f_{2}+\frac{\alpha+\mu}{\overline{\rho}}\nabla f_{1}-R_{0}\overline{\rho}\nabla\vartheta&\mbox{ in }\mathcal{F},\\ \operatorname{div}v=\frac{1}{\overline{\rho}}f_{1}&\mbox{ in }\mathcal{F},\\ v=-\mathcal{T}h_{1}&\mbox{ in }\partial\mathcal{F},\end{dcases} (4.37)
{Δs2η1=𝕋¯(ρ,v,ϑ)e3e3Δsh1+h2 in 𝒮,η1=sη1nS=0 on 𝒮.\begin{dcases}\Delta_{s}^{2}\eta_{1}=-\overline{\mathbb{T}}(\rho,v,\vartheta)e_{3}\cdot e_{3}-\Delta_{s}h_{1}+h_{2}&\mbox{ in }\mathcal{S},\\ \eta_{1}=\nabla_{s}\eta_{1}\cdot n_{S}=0&\mbox{ on }\partial\mathcal{S}.\end{dcases} (4.38)
ρdy+ρ¯𝒮η1ds=0,\int_{\mathcal{F}}\rho{\rm d}y+\overline{\rho}\int_{\mathcal{S}}\eta_{1}{\rm d}s=0, (4.39)

We can solve the two first equations and obtain the existence and uniqueness of ϑW2,q()\vartheta\in W^{2,q}(\mathcal{F}) and η2W02,q(𝒮)\eta_{2}\in W^{2,q}_{0}(\mathcal{S}) and we have the following estimate

ϑW2,q()Cf3Lq(),η2W2,q(𝒮)=h1W2,q(𝒮).\left\|\vartheta\right\|_{W^{2,q}(\mathcal{F})}\leqslant C\left\|f_{3}\right\|_{L^{q}(\mathcal{F})},\qquad\left\|\eta_{2}\right\|_{W^{2,q}(\mathcal{S})}=\left\|h_{1}\right\|_{W^{2,q}(\mathcal{S})}.

Using that [f1,f2,f3,h1,h2]𝒳m,[f_{1},f_{2},f_{3},h_{1},h_{2}]^{\top}\in\mathcal{X}_{\textnormal{m}}, we can solve (4.37) (see, for instance [36, Proposition 2.3, p.35]) and we obtain the existence and uniqueness of (ρ,v)(W1,q()/)×W2,q()3(\rho,v)\in\left(W^{1,q}(\mathcal{F})/\mathbb{R}\right)\times W^{2,q}(\mathcal{F})^{3} with the following estimate

vW2,q()3+ρW1,q()/(f1W1,q()+f2Lq()3+f3Lq()+h1W02,q(𝒮)).\left\|v\right\|_{W^{2,q}(\mathcal{F})^{3}}+\left\|\rho\right\|_{W^{1,q}(\mathcal{F})/\mathbb{R}}\leqslant\left(\left\|f_{1}\right\|_{W^{1,q}(\mathcal{F})}+\left\|f_{2}\right\|_{L^{q}(\mathcal{F})^{3}}+\left\|f_{3}\right\|_{L^{q}(\mathcal{F})}+\left\|h_{1}\right\|_{W^{2,q}_{0}(\mathcal{S})}\right).

Then we decompose ρ=ρm+ρavg\rho=\rho_{\textnormal{m}}+\rho_{\textnormal{avg}}, with

ρavg=1||ρ𝑑y=ρ¯||𝒮η1ds\rho_{\textnormal{avg}}=\frac{1}{|\mathcal{F}|}\int_{\mathcal{F}}\rho\ dy=-\frac{\overline{\rho}}{|\mathcal{F}|}\int_{\mathcal{S}}\eta_{1}{\rm d}s

and we can rewrite (4.38) as

{Δs2η1+R0ϑ¯ρ¯||𝒮η1ds=𝕋¯(ρm,v,ϑ)e3e3Δsh1+h2 in 𝒮,η1=sη1nS=0 on 𝒮.\begin{dcases}\Delta_{s}^{2}\eta_{1}+\frac{R_{0}\overline{\vartheta}\overline{\rho}}{|\mathcal{F}|}\int_{\mathcal{S}}\eta_{1}{\rm d}s=-\overline{\mathbb{T}}(\rho_{\textnormal{m}},v,\vartheta)e_{3}\cdot e_{3}-\Delta_{s}h_{1}+h_{2}&\mbox{ in }\mathcal{S},\\ \eta_{1}=\nabla_{s}\eta_{1}\cdot n_{S}=0&\mbox{ on }\partial\mathcal{S}.\end{dcases} (4.40)

Using the Fredholm alternative, the above system admits a unique solution η1W4,q(𝒮)\eta_{1}\in W^{4,q}(\mathcal{S}) and

η1W4,q(𝒮)C[f1,f2,f3,h1,h2]𝒳.\|\eta_{1}\|_{W^{4,q}(\mathcal{S})}\leqslant C\left\|[f_{1},f_{2},f_{3},h_{1},h_{2}]^{\top}\right\|_{\mathcal{X}}.

Step 3. Existence for λ+,λ0.\lambda\in\mathbb{C}^{+},\lambda\neq 0. By setting ρ=1λ(f1ρ¯divv),\displaystyle\rho=\frac{1}{\lambda}(f_{1}-\overline{\rho}\operatorname{div}v), the system (4.34) can be rewritten as

{λv1ρ¯div𝕋^λ(v,ϑ)=f^2 in ,λϑκ¯Δϑ=f3 in ,v=𝒯η2,ϑn=0 on ,λη1η2=h1 in 𝒮,λη2+Δs2η1Δsη2=𝕋^λ(v,ϑ)e3e3+h^2 in 𝒮,η1=sη1nS=η2=0 on 𝒮.\begin{dcases}\lambda v-\dfrac{1}{\overline{\rho}}\operatorname{div}\widehat{\mathbb{T}}_{\lambda}(v,\vartheta)=\widehat{f}_{2}&\mbox{ in }\mathcal{F},\\ \lambda\vartheta-\overline{\kappa}\Delta\vartheta=f_{3}&\mbox{ in }\mathcal{F},\\ v=\mathcal{T}\eta_{2},\quad\frac{\partial\vartheta}{\partial n}=0&\mbox{ on }\partial\mathcal{F},\\ \lambda\eta_{1}-\eta_{2}=h_{1}&\mbox{ in }\mathcal{S},\\ \lambda\eta_{2}+\Delta_{s}^{2}\eta_{1}-\Delta_{s}\eta_{2}=-\widehat{\mathbb{T}}_{\lambda}(v,\vartheta)e_{3}\cdot e_{3}+\widehat{h}_{2}&\mbox{ in }\mathcal{S},\\ \eta_{1}=\nabla_{s}\eta_{1}\cdot n_{S}=\eta_{2}=0&\mbox{ on }\partial\mathcal{S}.\end{dcases} (4.41)

where

𝕋^λ(v,ϑ)=2μ𝔻(v)+((α+R0ϑ¯ρ¯λ)divvR0ρ¯ϑ)I3,\displaystyle\widehat{\mathbb{T}}_{\lambda}(v,\vartheta)=2\mu\mathbb{D}(v)+\left(\left(\alpha+\frac{R_{0}\overline{\vartheta}\overline{\rho}}{\lambda}\right)\operatorname{div}v-R_{0}\overline{\rho}\vartheta\right)I_{3},
f^2=f2R0ϑ¯λρ¯f1,h^2=h2+R0ϑ¯λf1|𝒮.\displaystyle\widehat{f}_{2}=f_{2}-\frac{R_{0}\overline{\vartheta}}{\lambda\overline{\rho}}\nabla f_{1},\qquad\widehat{h}_{2}=h_{2}+\frac{R_{0}\overline{\vartheta}}{\lambda}f_{1}|_{\mathcal{S}}.

Let us set 𝒳^=Lq()3×Lq()×W02,q(𝒮)×Lq(𝒮).\widehat{\mathcal{X}}=L^{q}(\mathcal{F})^{3}\times L^{q}(\mathcal{F})\times W^{2,q}_{0}(\mathcal{S})\times L^{q}(\mathcal{S}). We define (see (4.23))

𝒟(Av,λ)=𝒟(Av),Av,λ=μρ¯Δ+(α+μρ¯+R0ρ¯ϑ¯λ)div.\mathcal{D}(A_{\textnormal{v},\lambda})=\mathcal{D}(A_{\textnormal{v}}),\quad A_{\textnormal{v},\lambda}=\frac{\mu}{\overline{\rho}}\Delta+\left(\frac{\alpha+\mu}{\overline{\rho}}+\frac{R_{0}\overline{\rho}\overline{\vartheta}}{\lambda}\right)\nabla\operatorname{div}.

In view of [35, Theorem 1.4] and of the Fredholm theorem, for each λ\lambda with Reλ0,\operatorname{Re}\lambda\geqslant 0, Av,λA_{\textnormal{v},\lambda} is an isomorphism from 𝒟(Av,λ)\mathcal{D}(A_{\textnormal{v},\lambda}) onto Lq()3L^{q}(\mathcal{F})^{3} for any q(1,).q\in(1,\infty). Let Dv,λ(W02,q(𝒮),W2,q()3)D_{\textnormal{v},\lambda}\in\mathcal{L}(W^{2,q}_{0}(\mathcal{S}),W^{2,q}(\mathcal{F})^{3}) defined by Dv,λg=w,D_{\textnormal{v},\lambda}g=w, where ww is the solution to the problem

{μρ¯Δw(α+μρ¯+R0ρ¯ϑ¯λ)(divw)=0 in ,w=𝒯g on .\begin{dcases}-\frac{\mu}{\overline{\rho}}\Delta w-\left(\frac{\alpha+\mu}{\overline{\rho}}+\frac{R_{0}\overline{\rho}\overline{\vartheta}}{\lambda}\right)\nabla(\operatorname{div}w)=0&\mbox{ in }\mathcal{F},\\ w=\mathcal{T}g&\mbox{ on }\partial\mathcal{F}.\end{dcases}

We introduce the unbounded operator 𝒜λ:𝒟(𝒜λ)𝒳^\mathcal{A}_{\lambda}:\mathcal{D}(\mathcal{A}_{\lambda})\to\widehat{\mathcal{X}} defined by

𝒟(𝒜λ)={[v,ϑ,η1,η2]W2,q()3×𝒟(Aϑ)×𝒟(AS);vDv,λη2𝒟(Av,λ)},\mathcal{D}(\mathcal{A}_{\lambda})=\left\{[v,\vartheta,\eta_{1},\eta_{2}]^{\top}\in W^{2,q}(\mathcal{F})^{3}\times\mathcal{D}(A_{\vartheta})\times\mathcal{D}(A_{S})\ ;\ v-D_{\textnormal{v},\lambda}\eta_{2}\in\mathcal{D}(A_{\textnormal{v},\lambda})\right\},

and

𝒜λ[vϑη1η2]=[Av,λ(vDv,λη2)R0ϑAϑϑη2Δs2η1+Δsη2𝕋^λ(v,ϑ)e3e3].\mathcal{A}_{\lambda}\begin{bmatrix}v\\ \vartheta\\ \eta_{1}\\ \eta_{2}\end{bmatrix}=\begin{bmatrix}A_{\textnormal{v},\lambda}(v-D_{\textnormal{v},\lambda}\eta_{2})-R_{0}\nabla\vartheta\\ A_{\vartheta}\vartheta\\ \eta_{2}\\ -\Delta_{s}^{2}\eta_{1}+\Delta_{s}\eta_{2}-\widehat{\mathbb{T}}_{\lambda}(v,\vartheta)e_{3}\cdot e_{3}\end{bmatrix}.

With the above notations, the system (4.41) can be written as

(λI𝒜λ)[v,ϑ,η1,η2]=[f^2,f3,h1,h^2].(\lambda I-\mathcal{A}_{\lambda})[v,\vartheta,\eta_{1},\eta_{2}]^{\top}=\left[\widehat{f}_{2},f_{3},h_{1},\widehat{h}_{2}\right]^{\top}. (4.42)

Proceeding as in the proof of 4.2, one can show the existence of λ~ρ(𝒜λ)\widetilde{\lambda}\in\rho(\mathcal{A}_{\lambda}). Using that 𝒜λ\mathcal{A}_{\lambda} has compact resolvent and the Fredholm alternative theorem, the existence and uniqueness of a solution to the system (4.41) are equivalent. Let us consider a solution of (4.41) with [f^2,f3,h1,h^2]=0\left[\widehat{f}_{2},f_{3},h_{1},\widehat{h}_{2}\right]^{\top}=0. As in Step 1, we can deduce that

[v,ϑ,η1,η2]W2,2()3×W2,2()×W4,2(𝒮)×W2,2(𝒮).[v,\vartheta,\eta_{1},\eta_{2}]^{\top}\in W^{2,2}(\mathcal{F})^{3}\times W^{2,2}(\mathcal{F})\times W^{4,2}(\mathcal{S})\times W^{2,2}(\mathcal{S}). (4.43)

Then ϑ=0\vartheta=0 and multiplying (4.41) by vv and by η2\eta_{2}, we deduce as in Step 1 that

[v,ϑ,η1,η2]=0.[v,\vartheta,\eta_{1},\eta_{2}]^{\top}=0.

This completes the proof of the proposition. ∎

4.4. Maximal LpL^{p}-LqL^{q} regularity of the linear system

Assume

p,q(1,),1p+12q1,1p+12q12.p,q\in(1,\infty),\quad\frac{1}{p}+\frac{1}{2q}\neq 1,\quad\frac{1}{p}+\frac{1}{2q}\neq\frac{1}{2}. (4.44)

Note that (1.11) implies (4.44). In order to show the maximal LpL^{p}-LqL^{q} regularity of the system (4.20)–(4.22), we first introduce the following decomposition: for any fL1()f\in L^{1}(\mathcal{F}),

f=fm+favg,withfmdy=0,favg=||1f(y)dy.f=f_{\textnormal{m}}+f_{\textnormal{avg}},\quad\text{with}\quad\int_{\mathcal{F}}f_{\textnormal{m}}\ {\rm d}y=0,\quad f_{\textnormal{avg}}=|\mathcal{F}|^{-1}\int_{\mathcal{F}}f(y)\ {\rm d}y. (4.45)

We use the same decomposition and the same notation for L1()L^{1}(\partial\mathcal{F}) and L1(𝒮).L^{1}(\mathcal{S}).

Let us recall some standard results on the heat equation and on the linearized compressible Navier-Stokes system:

Lemma 4.5.

There exists β1>0\beta_{1}>0 such that, for any β(0,β1)\beta\in(0,\beta_{1}) and for any η2,Wp,q,β2,4((0,);𝒮)\eta_{2,\dagger}\in W^{2,4}_{p,q,\beta}((0,\infty);\mathcal{S}) with

η2,(0,)0,\eta_{2,\dagger}(0,\cdot)\equiv 0,

the following linear system

{tρ+ρ¯divv=0 in (0,)×,tv1ρ¯div𝕋¯(ρ,v,0)=0 in (0,)×,v=𝒯η2, on (0,)×,ρ(0,)=0,v(0,)=0 in .\left\{\begin{array}[]{ll}\partial_{t}\rho_{\dagger}+\overline{\rho}\operatorname{div}v_{\dagger}=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}v_{\dagger}-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(\rho_{\dagger},v_{\dagger},0)=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ v_{\dagger}=\mathcal{T}\eta_{2,\dagger}&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \rho_{\dagger}(0,\cdot)=0,\quad v_{\dagger}(0,\cdot)=0&\mbox{ in }\mathcal{F}.\end{array}\right. (4.46)

admits a unique solution

ρ=ρ,m+ρ,avg,ρ,mWβ1,p(0,;W1,q()),tρ,avgLβp(0,),\displaystyle\rho_{\dagger}=\rho_{\dagger,\textnormal{m}}+\rho_{\dagger,\textnormal{avg}},\quad\rho_{\dagger,\textnormal{m}}\in W^{1,p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F})),\quad\partial_{t}\rho_{\dagger,\textnormal{avg}}\in L^{p}_{\beta}(0,\infty), (4.47)
vWp,q,β1,2((0,)×).\displaystyle v_{\dagger}\in W^{1,2}_{p,q,\beta}((0,\infty)\times\mathcal{F}). (4.48)

Moreover, the following estimate holds

ρ,mWβ1,p(0,;W1,q())+ρ,avgL(0,)+tρ,avgLβp(0,)+vWp,q,β1,2((0,)×)Cη2,Wp,q,β2,4((0,)×𝒮).\left\|\rho_{\dagger,\textnormal{m}}\right\|_{W^{1,p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|\rho_{\dagger,\textnormal{avg}}\right\|_{L^{\infty}(0,\infty)}+\left\|\partial_{t}\rho_{\dagger,\textnormal{avg}}\right\|_{L^{p}_{\beta}(0,\infty)}\\ +\left\|v_{\dagger}\right\|_{W^{1,2}_{p,q,\beta}((0,\infty)\times\mathcal{F})}\leqslant C\left\|\eta_{2,\dagger}\right\|_{W^{2,4}_{p,q,\beta}((0,\infty)\times\mathcal{S})}. (4.49)
Proof.

Let χ\chi be the cut-off function defined in (3.38) and we define

w(t,y1,y2,y3):=χ(y1,y2,y3)η2,(t,y1,y2)e3(t,y)(0,)ׯ.w_{\dagger}(t,y_{1},y_{2},y_{3}):=\chi(y_{1},y_{2},y_{3})\eta_{2,\dagger}(t,y_{1},y_{2})e_{3}\quad(t,y)\in(0,\infty)\times\overline{\mathcal{F}}.

Let us set u=vw.u_{\dagger}=v_{\dagger}-w_{\dagger}. Then (ρ,u)(\rho_{\dagger},u_{\dagger}) solves

{tρ+ρ¯divu=f1, in (0,)×,tu1ρ¯div𝕋¯(ρ,u,0)=f2, in (0,)×,u=0 on (0,)×,ρ(0,)=0,v(0,)=0 in ,\begin{dcases}\partial_{t}\rho_{\dagger}+\overline{\rho}\operatorname{div}u_{\dagger}=f_{1,\dagger}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}u_{\dagger}-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(\rho_{\dagger},u_{\dagger},0)=f_{2,\dagger}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ u_{\dagger}=0&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \rho_{\dagger}(0,\cdot)=0,\quad v_{\dagger}(0,\cdot)=0&\mbox{ in }\mathcal{F},\end{dcases} (4.50)

where

f1,=ρ¯divw,f2,=tw1ρ¯div𝕋¯(0,w,0).f_{1,\dagger}=-\overline{\rho}\operatorname{div}w_{\dagger},\quad f_{2,\dagger}=-\partial_{t}w_{\dagger}-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(0,w_{\dagger},0).

It is easy to see that

f1,Lβp(0,;W1,q())+f2Lβp(0,;Lq())Cη2,Wp,q,β2,4((0,)×𝒮),\left\|f_{1,\dagger}\right\|_{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|f_{2}\right\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))}\leqslant C\left\|\eta_{2,\dagger}\right\|_{W^{2,4}_{p,q,\beta}((0,\infty)\times\mathcal{S})},

for any β>0.\beta>0. We look for a solution to the system (4.50) of the form ρ=ρ,m+ρ,avg,\rho_{\dagger}=\rho_{\dagger,\textnormal{m}}+\rho_{\dagger,\textnormal{avg}}, where (ρ,m,u)(\rho_{\dagger,\textnormal{m}},u_{\dagger}) solves the system (4.50) with f1,f_{1,\dagger} replaced by f1,,mf_{1,\dagger,\textnormal{m}} and ρ,avg=0tf1,,avg(s)ds.\rho_{\dagger,\textnormal{avg}}=\displaystyle\int_{0}^{t}f_{1,\dagger,\textnormal{avg}}(s)\ {\rm d}s. By [16, Theorem 2.9], there exists β1>0\beta_{1}>0 such that for any β(0,β1)\beta\in(0,\beta_{1}), (f1,,m,f2,)Lβp(0,;W1,q())×Lβp(0,;Lq()),(f_{1,\dagger,\textnormal{m}},f_{2,\dagger})\in L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))\times L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F})), we have

ρ,mWβ1,p(0,;W1,q())+vWp,q,β1,2((0,)×)Cf1,Lβp(0,;W1,q())+f2Lβp(0,;Lq()).\left\|\rho_{\dagger,\textnormal{m}}\right\|_{W^{1,p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|v_{\dagger}\right\|_{W^{1,2}_{p,q,\beta}((0,\infty)\times\mathcal{F})}\leqslant C\left\|f_{1,\dagger}\right\|_{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|f_{2}\right\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))}.

Combining the above estimates we obtain the conclusion of the lemma. ∎

Combining Step 3 of the proof of 3.2 and [12, Proposition 6.4], we deduce the following result:

Lemma 4.6.

Assume β>0\beta>0. There exists γ1>0\gamma_{1}>0 such that for any

ϑ0Bq,p2(11/p)(),f3Lβp(0,;Lq()),\vartheta^{0}\in B^{2(1-1/p)}_{q,p}(\mathcal{F}),\quad f_{3}\in L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F})),
gFp,q,β(11/q)/2(0,;Lq())Lβp(0,;W11/q,q())g\in F^{(1-1/q)/2}_{p,q,\beta}(0,\infty;L^{q}(\partial\mathcal{F}))\cap L^{p}_{\beta}(0,\infty;W^{1-1/q,q}(\partial\mathcal{F}))

with

ϑ0n=g(0,) on ,\frac{\partial\vartheta^{0}}{\partial n}=g(0,\cdot)\quad\mbox{ on }\partial\mathcal{F},

the following heat equation

{tϑ+γ1ϑκ¯Δϑ=f3 in (0,)×,ϑn=g on (0,)×,ϑ(0,)=ϑ0 in .\begin{dcases}\partial_{t}\vartheta_{\sharp}+\gamma_{1}\vartheta_{\sharp}-\overline{\kappa}\Delta\vartheta_{\sharp}=f_{3}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \frac{\partial\vartheta_{\sharp}}{\partial n}=g&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \vartheta_{\sharp}(0,\cdot)=\vartheta^{0}&\mbox{ in }\mathcal{F}.\end{dcases} (4.51)

admits a unique solution ϑWp,q,β1,2((0,);)\vartheta_{\sharp}\in W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{F}). Moreover, we have the following estimate

ϑWp,q,β1,2((0,);)C(ϑ0Bq,p2(11/p)()+f3Lβp(0,;Lq())+gFp,q,β(11/q)/2(0,;Lq())+gLβp(0,;W11/q,q())).\left\|\vartheta_{\sharp}\right\|_{W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{F})}\leqslant C\Big{(}\left\|\vartheta^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})}+\left\|f_{3}\right\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))}+\|g\|_{F^{(1-1/q)/2}_{p,q,\beta}(0,\infty;L^{q}(\partial\mathcal{F}))}\\ +\|g\|_{L^{p}_{\beta}(0,\infty;W^{1-1/q,q}(\partial\mathcal{F}))}\Big{)}. (4.52)

We consider the subset of initial conditions

𝒥p,q:={[ρ0,v0,ϑ0,η10,η20]W1,q()×Bq,p2(11/p)()3×Bq,p2(11/p)()×Bq,p2(21/p)(𝒮)×Bq,p2(11/p)(𝒮)η10=sη10nS=0 on 𝒮,v0=𝒯η20 on andη20=0 on 𝒮if1p+12q<1,sη20nS=0 on 𝒮if1p+12q<12}\mathcal{J}_{p,q}:=\Bigg{\{}\left[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}\right]^{\top}\in W^{1,q}(\mathcal{F})\times B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}\times B^{2(1-1/p)}_{q,p}(\mathcal{F})\times B^{2(2-1/p)}_{q,p}(\mathcal{S})\times B^{2(1-1/p)}_{q,p}(\mathcal{S})\\ \eta_{1}^{0}=\nabla_{s}\eta_{1}^{0}\cdot n_{S}=0\quad\mbox{ on }\mathcal{S},\\ v^{0}=\mathcal{T}\eta_{2}^{0}\mbox{ on }\partial\mathcal{F}\quad\text{and}\quad\eta_{2}^{0}=0\mbox{ on }\partial\mathcal{S}\quad\mbox{if}\quad\frac{1}{p}+\frac{1}{2q}<1,\\ \nabla_{s}\eta_{2}^{0}\cdot n_{S}=0\mbox{ on }\partial\mathcal{S}\quad\mbox{if}\quad\frac{1}{p}+\frac{1}{2q}<\frac{1}{2}\Bigg{\}} (4.53)

with

[ρ0,v0,ϑ0,η10,η20]𝒥p,q:=ρ0W1,q()+v0Bq,p2(11/p)()3+ϑ0Bq,p2(11/p)()+η10Bq,p2(21/p)(𝒮)+η20Bq,p2(11/p)(𝒮).\left\|\left[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}\right]^{\top}\right\|_{\mathcal{J}_{p,q}}:=\left\|\rho^{0}\right\|_{W^{1,q}(\mathcal{F})}+\left\|v^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})^{3}}+\left\|\vartheta^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{F})}\\ +\left\|\eta_{1}^{0}\right\|_{B^{2(2-1/p)}_{q,p}(\mathcal{S})}+\left\|\eta_{2}^{0}\right\|_{B^{2(1-1/p)}_{q,p}(\mathcal{S})}.

We also consider the following subset for the source terms:

p,q,βcc={[f1,f2,f3,g,h~,h^];f1Lβp(0,,W1,q()),f2Lβp(0,;Lq())3,f3Lp(0,;Lq()),gFp,q,β(11/q)/2(0,;Lq())Lβp(0,;W11/q,q()),h^L(0,),th^Lβp(0,),h~Lβp(0,;Lq(𝒮)),withg(0,)=ϑ0n if 1p+12q<12},\mathcal{R}_{p,q,\beta}^{cc}=\Big{\{}[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\ ;\ f_{1}\in L^{p}_{\beta}(0,\infty,W^{1,q}(\mathcal{F})),f_{2}\in L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))^{3},\\ f_{3}\in L^{p}_{\infty}(0,\infty;L^{q}(\mathcal{F})),g\in F^{(1-1/q)/2}_{p,q,\beta}(0,\infty;L^{q}(\partial\mathcal{F}))\cap L^{p}_{\beta}(0,\infty;W^{1-1/q,q}(\partial\mathcal{F})),\\ \widehat{h}\in L^{\infty}(0,\infty),\quad\partial_{t}\widehat{h}\in L^{p}_{\beta}(0,\infty),\quad\widetilde{h}\in L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{S})),\\ \mbox{with}\ g(0,\cdot)=\frac{\partial\vartheta_{0}}{\partial n}\mbox{ if }\displaystyle\frac{1}{p}+\frac{1}{2q}<\frac{1}{2}\Big{\}}, (4.54)

with

[f1,f2,f3,g,h~,h^]p,q,βcc=f1Lβp(0,;W1,q())+f2Lβp(0,;Lq())3+f3Lβp(0,;Lq())+gFp,q,β(11/q)/2(0,;Lq())Lβp(0,;W11/q,q())+h~Lβp(0,;Lq(𝒮))+h^L(0,)+th^Lβp(0,).\|[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\|_{\mathcal{R}_{p,q,\beta}^{cc}}=\|f_{1}\|_{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\|f_{2}\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))^{3}}+\|f_{3}\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))}\\ +\|g\|_{F^{(1-1/q)/2}_{p,q,\beta}(0,\infty;L^{q}(\partial\mathcal{F}))\cap L^{p}_{\beta}(0,\infty;W^{1-1/q,q}(\partial\mathcal{F}))}\\ +\|\widetilde{h}\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{S}))}+\|\widehat{h}\|_{L^{\infty}(0,\infty)}+\|\partial_{t}\widehat{h}\|_{L^{p}_{\beta}(0,\infty)}.

We take β=min(β0,β1)>0\beta=\min(\beta_{0},\beta_{1})>0 where β0\beta_{0} is the constant in 4.3 and where β1\beta_{1} is the constant in 4.5. We decompose the solution of the system (4.20)–(4.22) as follows

ρ=ρ+ρ+ρ,v=v+v,ϑ=ϑ+ϑ+ϑ,η1=η1,+η1,,η2=η2,+η2,,\rho=\rho_{\flat}+\rho_{\diamond}+\rho_{\dagger},\quad v=v_{\diamond}+v_{\dagger},\quad\vartheta=\vartheta_{\diamond}+\vartheta_{\sharp}+\vartheta_{\flat},\quad\eta_{1}=\eta_{1,\diamond}+\eta_{1,\dagger},\quad\eta_{2}=\eta_{2,\diamond}+\eta_{2,\dagger}, (4.55)

where ϑ\vartheta_{\sharp} is the solution of (4.51) given by 4.6, where

ϑ(t):=0tγ1ϑ,avg(r)dr,ρ(t)=1||(ρ0dy+ρ¯𝒮η10ds)+0tf1,avg(r)dr.\vartheta_{\flat}(t):=\int_{0}^{t}\gamma_{1}\vartheta_{\sharp,\textnormal{avg}}(r)\ {\rm d}r,\quad\rho_{\flat}(t)=\frac{1}{|\mathcal{F}|}\left(\int_{\mathcal{F}}\rho^{0}\ {\rm d}y+\overline{\rho}\int_{\mathcal{S}}\eta_{1}^{0}\ {\rm d}s\right)+\int_{0}^{t}f_{1,\textnormal{avg}}(r)\ {\rm d}r. (4.56)

where [ρ,v,ϑ,η1,,η2,][\rho_{\diamond},v_{\diamond},\vartheta_{\diamond},\eta_{1,\diamond},\eta_{2,\diamond}]^{\top} is solution of the following system

{tρ+ρ¯divv=f1,m in (0,)×,tv1ρ¯div𝕋¯(ρ,v,ϑ)=f2R0ϑ in (0,)×,tϑκ¯Δϑ=γ1ϑ,m in (0,)×,tη1,η2,=0 in (0,)×,tη2,+Δs2η1,Δsη2,=𝕋¯(ρ,v,ϑ)e3e3+h~+R0ρ¯ϑ|𝒮 in (0,)×𝒮,\left\{\begin{array}[]{ll}\partial_{t}\rho_{\diamond}+\overline{\rho}\operatorname{div}v_{\diamond}=f_{1,\textnormal{m}}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}v_{\diamond}-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(\rho_{\diamond},v_{\diamond},\vartheta_{\diamond})=f_{2}-R_{0}\nabla\vartheta_{\sharp}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\vartheta_{\diamond}-\overline{\kappa}\Delta\vartheta_{\diamond}=\gamma_{1}\vartheta_{\sharp,\textnormal{m}}&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\eta_{1,\diamond}-\eta_{2,\diamond}=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\eta_{2,\diamond}+\Delta_{s}^{2}\eta_{1,\diamond}-\Delta_{s}\eta_{2,\diamond}=-\overline{\mathbb{T}}(\rho_{\diamond},v_{\diamond},\vartheta_{\diamond})e_{3}\cdot e_{3}+\widetilde{h}+R_{0}\overline{\rho}{\vartheta_{\sharp}}_{|\mathcal{S}}&\mbox{ in }(0,\infty)\times\mathcal{S},\end{array}\right. (4.57)
{v=𝒯η2, on (0,)×,ϑn=0 on (0,)×,η1,=sη1,nS=0 on (0,)×𝒮,\left\{\begin{array}[]{ll}v_{\diamond}=\mathcal{T}\eta_{2,\diamond}&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \dfrac{\partial\vartheta_{\diamond}}{\partial n}=0&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \eta_{1,\diamond}=\nabla_{s}\eta_{1,\diamond}\cdot n_{S}=0&\mbox{ on }(0,\infty)\times\partial\mathcal{S},\end{array}\right. (4.58)
{η1,(0,)=η10,η2,(0,)=η20 in 𝒮,ρ(0,)=ρ0ρ(0),v(0,)=v0,ϑ(0,)=0 in .\left\{\begin{array}[]{ll}\eta_{1,\diamond}(0,\cdot)=\eta_{1}^{0},\quad\eta_{2,\diamond}(0,\cdot)=\eta_{2}^{0}&\mbox{ in }\mathcal{S},\\ \rho_{\diamond}(0,\cdot)=\displaystyle\rho^{0}-\rho_{\flat}(0),\quad v_{\diamond}(0,\cdot)=v^{0},\quad\vartheta_{\diamond}(0,\cdot)=0&\mbox{ in }\mathcal{F}.\end{array}\right. (4.59)

and where [ρ,v,ϑ,η1,,η2,][\rho_{\dagger},v_{\dagger},\vartheta_{\dagger},\eta_{1,\dagger},\eta_{2,\dagger}]^{\top} is solution of the following system

{tρ+ρ¯divv=0 in (0,)×,tv1ρ¯div𝕋¯(ρ,v,ϑ)=0 in (0,)×,tϑκ¯Δϑ=0 in (0,)×,tη1,η2,=0 in (0,)×,tη2,+Δs2η1,Δsη2,=𝕋¯(ρ,v,ϑ)e3e3+h^+R0ϑ¯ρ+R0ρ¯ϑ in (0,)×𝒮,\left\{\begin{array}[]{ll}\partial_{t}\rho_{\dagger}+\overline{\rho}\operatorname{div}v_{\dagger}=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}v_{\dagger}-\dfrac{1}{\overline{\rho}}\operatorname{div}\overline{\mathbb{T}}(\rho_{\dagger},v_{\dagger},\vartheta_{\dagger})=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\vartheta_{\dagger}-\overline{\kappa}\Delta\vartheta_{\dagger}=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\eta_{1,\dagger}-\eta_{2,\dagger}=0&\mbox{ in }(0,\infty)\times\mathcal{F},\\ \partial_{t}\eta_{2,\dagger}+\Delta_{s}^{2}\eta_{1,\dagger}-\Delta_{s}\eta_{2,\dagger}=-\overline{\mathbb{T}}(\rho_{\dagger},v_{\dagger},\vartheta_{\dagger})e_{3}\cdot e_{3}+\widehat{h}+R_{0}\overline{\vartheta}\rho_{\flat}+R_{0}\overline{\rho}{\vartheta_{\flat}}&\mbox{ in }(0,\infty)\times\mathcal{S},\end{array}\right. (4.60)
{v=𝒯η2, on (0,)×,ϑn=0 on (0,)×,η1,=sη1,nS=0 on (0,)×𝒮,\left\{\begin{array}[]{ll}v_{\dagger}=\mathcal{T}\eta_{2,\dagger}&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \dfrac{\partial\vartheta_{\dagger}}{\partial n}=0&\mbox{ on }(0,\infty)\times\partial\mathcal{F},\\ \eta_{1,\dagger}=\nabla_{s}\eta_{1,\dagger}\cdot n_{S}=0&\mbox{ on }(0,\infty)\times\partial\mathcal{S},\end{array}\right. (4.61)
{η1,(0,)=0,η2,(0,)=0 in 𝒮,ρ(0,)=0,v(0,)=0,ϑ(0,)=0 in .\left\{\begin{array}[]{ll}\eta_{1,\dagger}(0,\cdot)=0,\quad\eta_{2,\dagger}(0,\cdot)=0&\mbox{ in }\mathcal{S},\\ \rho_{\dagger}(0,\cdot)=0,\quad v_{\dagger}(0,\cdot)=0,\quad\vartheta_{\dagger}(0,\cdot)=0&\mbox{ in }\mathcal{F}.\end{array}\right. (4.62)

Let us show that the decomposition (4.55) is valid. First, we can check that

ϑ,ρCb0([0,)),tϑ,tρLβp(0,).\vartheta_{\flat},\rho_{\flat}\in C^{0}_{b}([0,\infty)),\quad\partial_{t}\vartheta_{\flat},\partial_{t}\rho_{\flat}\in L^{p}_{\beta}(0,\infty).

Second, for the system (4.57)–(4.59), we note that from (4.32) and (4.53)

[f1,m,f2R0ϑ,γ1ϑ,m,0,h~+R0ρ¯ϑ|𝒮]Lβp(0,;𝒳m),\Big{[}f_{1,\textnormal{m}},f_{2}-R_{0}\nabla\vartheta_{\sharp},\gamma_{1}\vartheta_{\sharp,\textnormal{m}},0,\widetilde{h}+R_{0}\overline{\rho}{\vartheta_{\sharp}}_{|\mathcal{S}}\Big{]}^{\top}\in L^{p}_{\beta}(0,\infty;\mathcal{X}_{\textnormal{m}}),
[ρ0ρ(0),v0,0,η10,η20](𝒳m,𝒟(𝒜FS))11/p,p.\Big{[}\rho^{0}-\rho_{\flat}(0),v^{0},0,\eta_{1}^{0},\eta_{2}^{0}\Big{]}^{\top}\in\left(\mathcal{X}_{\textnormal{m}},\mathcal{D}(\mathcal{A}_{FS})\right)_{1-1/p,p}.

From 4.2 and 4.3 we know that 𝒜FS+βI\mathcal{A}_{FS}+\beta I is a \mathcal{R}-sectorial operator on 𝒳m\mathcal{X}_{\textnormal{m}} and generates an analytic exponential stable semigroup on 𝒳m.\mathcal{X}_{\textnormal{m}}. Therefore, by 2.4, the system (4.57)–(4.59) admits a unique solution

[ρ,v,ϑ,η1,,η2,]Lβp(0,;𝒟(𝒜FS)𝒳m)Wβ1,p(0,;𝒳m).\Big{[}\rho_{\diamond},v_{\diamond},\vartheta_{\diamond},\eta_{1,\diamond},\eta_{2,\diamond}\Big{]}^{\top}\in L^{p}_{\beta}(0,\infty;\mathcal{D}(\mathcal{A}_{FS})\cap\mathcal{X}_{\textnormal{m}})\cap W^{1,p}_{\beta}(0,\infty;\mathcal{X}_{\textnormal{m}}). (4.63)

Finally, let us consider the system (4.60)–(4.62). Note that ϑ0\vartheta_{\dagger}\equiv 0. Moreover

ddt[tρtvtϑtη1,tη2,]=𝒜FS[tρtvtϑtη1,tη2,]+[0000th^+R0ϑ¯f1,avg+R0ρ¯γ1ϑ,avg],[tρtvtϑtη1,tη2,](0)=[00000].\frac{d}{dt}\begin{bmatrix}\partial_{t}\rho_{\dagger}\\ \partial_{t}v_{\dagger}\\ \partial_{t}\vartheta_{\dagger}\\ \partial_{t}{\eta}_{1,\dagger}\\ \partial_{t}{\eta}_{2,\dagger}\end{bmatrix}=\mathcal{A}_{FS}\begin{bmatrix}\partial_{t}\rho_{\dagger}\\ \partial_{t}v_{\dagger}\\ \partial_{t}\vartheta_{\dagger}\\ \partial_{t}{\eta}_{1,\dagger}\\ \partial_{t}{\eta}_{2,\dagger}\end{bmatrix}+\begin{bmatrix}0\\ 0\\ 0\\ 0\\ \partial_{t}\widehat{h}+R_{0}\overline{\vartheta}f_{1,\textnormal{avg}}+R_{0}\overline{\rho}\gamma_{1}\vartheta_{\sharp,\textnormal{avg}}\end{bmatrix},\quad\begin{bmatrix}\partial_{t}\rho_{\dagger}\\ \partial_{t}v_{\dagger}\\ \partial_{t}\vartheta_{\dagger}\\ \partial_{t}{\eta}_{1,\dagger}\\ \partial_{t}{\eta}_{2,\dagger}\end{bmatrix}(0)=\begin{bmatrix}0\\ 0\\ 0\\ 0\\ 0\end{bmatrix}. (4.64)

Using that th^+R0ϑ¯f1,avg+R0ρ¯γ1ϑ,avgLβp(0,),\partial_{t}\widehat{h}+R_{0}\overline{\vartheta}f_{1,\textnormal{avg}}+R_{0}\overline{\rho}\gamma_{1}\vartheta_{\sharp,\textnormal{avg}}\in L^{p}_{\beta}(0,\infty), and combining as above 4.2, 4.3 and 2.4, we infer that

[tρ,tv,tϑ,tη1,,tη2,]Lβp(0,;𝒟(𝒜FS)𝒳m)Wβ1,p(0,;𝒳m).\Big{[}\partial_{t}\rho_{\dagger},\partial_{t}v_{\dagger},\partial_{t}\vartheta_{\dagger},\partial_{t}{\eta}_{1,\dagger},\partial_{t}{\eta}_{2,\dagger}\Big{]}^{\top}\in L^{p}_{\beta}(0,\infty;\mathcal{D}(\mathcal{A}_{FS})\cap\mathcal{X}_{\textnormal{m}})\cap W^{1,p}_{\beta}(0,\infty;\mathcal{X}_{\textnormal{m}}). (4.65)

In particular,

η2,Wp,q,β2,4((0,);𝒮).{\eta}_{2,\dagger}\in W^{2,4}_{p,q,\beta}((0,\infty);\mathcal{S}). (4.66)

Then, we use 4.5 to deduce (ρ,v)(\rho_{\dagger},v_{\dagger}) satisfies (4.47)–(4.48).

Let us also write

ρ~=ρ+ρ,m,ρ^=ρ+ρ,avg,ϑ~=ϑ+ϑ,ϑ^=ϑ,\widetilde{\rho}=\rho_{\diamond}+\rho_{\dagger,\textnormal{m}},\quad\widehat{\rho}=\rho_{\flat}+\rho_{\dagger,\textnormal{avg}},\quad\widetilde{\vartheta}=\vartheta_{\diamond}+\vartheta_{\sharp},\quad\widehat{\vartheta}=\vartheta_{\flat}, (4.67)

so that

ρ=ρ~+ρ^,ϑ=ϑ~+ϑ^.\rho=\widetilde{\rho}+\widehat{\rho},\quad\vartheta=\widetilde{\vartheta}+\widehat{\vartheta}. (4.68)

Gathering the above properties, we have obtained the following theorem:

Theorem 4.7.

Assume (4.44). There exists β>0\beta>0 such that for any

[ρ0,v0,ϑ0,η10,η20]𝒥p,q,[f1,f2,f3,g,h~,h^]p,q,βcc,h=h~+h^,[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\in\mathcal{J}_{p,q},\quad[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\in\mathcal{R}_{p,q,\beta}^{cc},\quad h=\widetilde{h}+\widehat{h},

the system (4.20)–(4.22) admits a unique solution satisfying (4.15)–(4.18) and

ρL(0,;W1,q())+ρWβ1,p(0,;Lq())3+tρLβp(0,;W1,q())+vWp,q,β1,2((0,);)3+ϑL(0,;Bq,p2(11/p)())+ϑLβp(0,;W1,q())3+tϑLβp(0,;Lq())+η1L(0,;Bq,p2(21/p)(𝒮))+η2Wp,q,β1,2((0,);𝒮)CL([ρ0,v0,ϑ0,η10,η20]𝒥p,q+[f1,f2,f3,g,h~,h^]p,q,βcc).\|\rho\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}+\|\nabla\rho\|_{W^{1,p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))^{3}}+\|\partial_{t}\rho\|_{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\|v\|_{W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{F})^{3}}\\ +\left\|\vartheta\right\|_{L^{\infty}(0,\infty;B^{2(1-1/p)}_{q,p}(\mathcal{F}))}+\|\nabla\vartheta\|_{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))^{3}}+\|\partial_{t}\vartheta\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))}\\ +\|\eta_{1}\|_{L^{\infty}(0,\infty;B^{2(2-1/p)}_{q,p}(\mathcal{S}))}+\|\eta_{2}\|_{W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{S})}\\ \leqslant C_{L}\Big{(}\left\|[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\right\|_{\mathcal{J}_{p,q}}+\left\|[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}\Big{)}. (4.69)

Moreover, we can decompose the solution as (4.67)-(4.68), with

ρ~Wβ1,p(0,;W1,q()),ϑ~Wp,q,β1,2((0,);),[ρ^,ϑ^]L(0,)2,[tρ^,tϑ^]Lβp(0,)2,\displaystyle\widetilde{\rho}\in W^{1,p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F})),\;\widetilde{\vartheta}\in W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{F}),\;\left[\widehat{\rho},\widehat{\vartheta}\right]^{\top}\in L^{\infty}(0,\infty)^{2},\;\left[\partial_{t}\widehat{\rho},\partial_{t}\widehat{\vartheta}\right]^{\top}\in L^{p}_{\beta}(0,\infty)^{2}, (4.70)

and

ρ~Wβ1,p(0,;W1,q())+ϑ~Wp,q,β1,2((0,);)+[ρ^,ϑ^]L(0,)2+[tρ^,tϑ^]Lβp(0,)2CL([ρ0,v0,ϑ0,η10,η20]𝒥p,q+[f1,f2,f3,g,h~,h^]p,q,βcc).\left\|\widetilde{\rho}\right\|_{W^{1,p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|\widetilde{\vartheta}\right\|_{W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{F})}+\left\|\left[\widehat{\rho},\widehat{\vartheta}\right]^{\top}\right\|_{L^{\infty}(0,\infty)^{2}}+\left\|\left[\partial_{t}\widehat{\rho},\partial_{t}\widehat{\vartheta}\right]^{\top}\right\|_{L^{p}_{\beta}(0,\infty)^{2}}\\ \leqslant C_{L}\Big{(}\left\|[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\right\|_{\mathcal{J}_{p,q}}+\left\|[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}\Big{)}. (4.71)

4.5. Proof of 4.1

In this subsection, we prove 4.1 (or equivalently 1.2): we show the existence and uniqueness of global in time solutions for the system (4.3)–(4.13) under a smallness assumption on the initial data.

Let us assume the hypotheses of 4.1, with β\beta given by 4.7. Assume

[ρ0,v0,ϑ0,η10,η20]𝒥p,q,[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\in\mathcal{J}_{p,q},

where 𝒥p,q\mathcal{J}_{p,q} is defined by (4.53). For R>0,R>0, we define R\mathcal{B}_{R} as follows

R={[f1,f2,f3,g,h~,h^]p,q,βcc;[f1,f2,f3,g,h~,h^]p,q,βccR},\mathcal{B}_{R}=\Big{\{}[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\in\mathcal{R}_{p,q,\beta}^{cc}\ ;\ \left\|\left[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}\right]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}\leqslant R\Big{\}}, (4.72)

where p,q,βcc\mathcal{R}_{p,q,\beta}^{cc} is defined by (4.54). By using 4.6 with f3=0f_{3}=0 and g=0g=0, we see that there exists a constant C>0C>0 independent of RR such that if

[ρ0,v0,ϑ0,η10,η20]𝒥p,qCR,\left\|[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}\right\|_{\mathcal{J}_{p,q}}\leqslant CR, (4.73)

then R\mathcal{B}_{R} is a nonempty closed subset of the Banach space

p,q,β={[f1,f2,f3,g,h~,h^];f1Lβp(0,,W1,q()),f2Lβp(0,;Lq())3,f3Lp(0,;Lq()),gFp,q,β(11/q)/2(0,;Lq())Lβp(0,;W11/q,q()),h^L(0,),th^Lβp(0,),h~Lβp(0,;Lq(𝒮))}.\mathcal{R}_{p,q,\beta}=\Big{\{}[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\ ;\ f_{1}\in L^{p}_{\beta}(0,\infty,W^{1,q}(\mathcal{F})),\quad f_{2}\in L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))^{3},\\ f_{3}\in L^{p}_{\infty}(0,\infty;L^{q}(\mathcal{F})),\quad g\in F^{(1-1/q)/2}_{p,q,\beta}(0,\infty;L^{q}(\partial\mathcal{F}))\cap L^{p}_{\beta}(0,\infty;W^{1-1/q,q}(\partial\mathcal{F})),\\ \widehat{h}\in L^{\infty}(0,\infty),\quad\partial_{t}\widehat{h}\in L^{p}_{\beta}(0,\infty),\quad\widetilde{h}\in L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{S}))\Big{\}}. (4.74)

We define the map

ΞR:RR,[f1,f2,f3,g,h~,h^][F1,F2,F3,G,H~,H^],\Xi_{R}:\mathcal{B}_{R}\longrightarrow\mathcal{B}_{R},\quad[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\longmapsto[F_{1},F_{2},F_{3},G,\widetilde{H},\widehat{H}]^{\top}, (4.75)

where [ρ,v,ϑ,η,tη][\rho,v,\vartheta,\eta,\partial_{t}\eta]^{\top} is the solution to the system (4.20)–(4.22) associated with [f1,f2,f3,g,h][f_{1},f_{2},f_{3},g,h]^{\top} and [ρ0,v0,ϑ0,η10,η20],[\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top}, (see 4.7), where F1,F2,F3F_{1},F_{2},F_{3} and GG are given by (4.9)–(4.12) and where

H~=μ[1δX(v𝔹X+𝔹Xv)[sη1]2μ𝔻(v)e3]e3α(1δX𝔹XI3):v+R0ρ~ϑ~+R0ρ~ϑ^+R0ρ^ϑ~,\widetilde{H}=-\mu\left[\frac{1}{\delta_{X}}\left(\nabla v\mathbb{B}_{X}^{\top}+\mathbb{B}_{X}\nabla v^{\top}\right)\begin{bmatrix}-\nabla_{s}\eta\\ 1\end{bmatrix}-2\mu\mathbb{D}(v)e_{3}\right]\cdot e_{3}-\alpha\Big{(}\frac{1}{\delta_{X}}\mathbb{B}_{X}-I_{3}\Big{)}:\nabla v\\ +R_{0}\widetilde{\rho}\widetilde{\vartheta}+R_{0}\widetilde{\rho}\widehat{\vartheta}+R_{0}\widehat{\rho}\widetilde{\vartheta}, (4.76)

and

H^=R0ρ^ϑ^.\widehat{H}=R_{0}\widehat{\rho}\widehat{\vartheta}. (4.77)

In the above definitions, we have used the decomposition of ρ\rho and ϑ\vartheta given by (4.67)-(4.68). We can check that HH defined by (4.13) satisfies

H=H~+H^.H=\widetilde{H}+\widehat{H}.

In order to prove 4.1, it is enough to show that the mapping ΞR\Xi_{R} is well defined, Ξ(R)R\Xi(\mathcal{B}_{R})\subset\mathcal{B}_{R} and Ξ|R\Xi_{|\mathcal{B}_{R}} is a strict contraction, for RR small enough.

Throughout this subsection, CC will be a positive constant depending on p,qp,q and β\beta but independent of R,R, which may change from line to line. To simplify the computations, we assume that R(0,1).R\in(0,1).

Since 2<p<2<p<\infty and 3<q<,3<q<\infty, one has (see, for instance [38, (7), p. 196])

Bq,p2(11/p)()W1,q()L().B^{2(1-1/p)}_{q,p}(\mathcal{F})\hookrightarrow W^{1,q}(\mathcal{F})\hookrightarrow L^{\infty}(\mathcal{F}).

Therefore, from 4.7, we obtain

vLβ(0,;W1,q())3+ϑL(0,;W1,q())+ϑLβp(0,;L())3+ηL(0,;W3,q(𝒮))+ηL(0,;C2(𝒮¯))CR.\left\|v\right\|_{L^{\infty}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))^{3}}+\left\|\vartheta\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|\nabla\vartheta\right\|_{L^{p}_{\beta}(0,\infty;L^{\infty}(\mathcal{F}))^{3}}\\ +\left\|\eta\right\|_{L^{\infty}(0,\infty;W^{3,q}(\mathcal{S}))}+\left\|\eta\right\|_{L^{\infty}(0,\infty;C^{2}(\overline{\mathcal{S}}))}\leqslant CR. (4.78)

From the definition of X0X^{0} from (3.2) and from (4.73) we deduce that

X0I3W2,q()9CR.\left\|\nabla X^{0}-I_{3}\right\|_{W^{2,q}(\mathcal{F})^{9}}\leqslant CR.

Using the above estimate and the definition of XX (see (3.18)) it follows that

XI3L(0,;W1,q())9X0I3W2,q()9+CvLβp(0,;W1,q())9CR.\left\|\nabla X-I_{3}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant\left\|\nabla X^{0}-I_{3}\right\|_{W^{2,q}(\mathcal{F})^{9}}+C\left\|\nabla v\right\|_{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant CR. (4.79)

In particular, by choosing RR sufficiently small, we have

XI3L((0,)×)912.\left\|\nabla X-I_{3}\right\|_{L^{\infty}((0,\infty)\times\mathcal{F})^{9}}\leqslant\frac{1}{2}.

Thus XX is a C1C^{1}-diffeomorphism for RR small enough. Moreover, by combining the above estimates with (3.5) and using that tX=v\partial_{t}X=v, we also deduce

𝔹XI3L(0,;W1,q())9CR,t𝔹XLp(0,;W1,q())9CR,𝔹XL(0,;W1,q())9C,\left\|\mathbb{B}_{X}-I_{3}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant CR,\quad\left\|\partial_{t}\mathbb{B}_{X}\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant CR,\quad\left\|\mathbb{B}_{X}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant C, (4.80)
δX1L(0,;W1,q())CR,tδXLp(0,;W1,q())CR,δXL(0,;W1,q())C.\left\|\delta_{X}-1\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}\leqslant CR,\quad\left\|\partial_{t}\delta_{X}\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))}\leqslant CR,\quad\left\|\delta_{X}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}\leqslant C. (4.81)

Consequently, for RR small enough

δX12 for all (t,y)(0,)×.\delta_{X}\geqslant\frac{1}{2}\quad\mbox{ for all }(t,y)\in(0,\infty)\times\mathcal{F}. (4.82)

We thus deduce

1δX1L(0,;W1,q())CR,t(1δX)Lp(0,;W1,q())CR,1δXL(0,;W1,q())C.\left\|\frac{1}{\delta_{X}}-1\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}\leqslant CR,\quad\left\|\partial_{t}\left(\frac{1}{\delta_{X}}\right)\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))}\leqslant CR,\quad\left\|\frac{1}{\delta_{X}}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}\leqslant C. (4.83)

Using the above estimates and (3.5), we also obtain

𝔸XI3L(0,;W1,q())9CR,𝔸XL(0,;W1,q())9C,\displaystyle\left\|\mathbb{A}_{X}-I_{3}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant CR,\quad\left\|\mathbb{A}_{X}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant C, (4.84)
t𝔸XLp(0,;W1,q())9CR,𝔸XI3C1/p([0,);W1,q())9CR.\displaystyle\left\|\partial_{t}\mathbb{A}_{X}\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant CR,\quad\left\|\mathbb{A}_{X}-I_{3}\right\|_{C^{1/p^{\prime}}([0,\infty);W^{1,q}(\mathcal{F}))^{9}}\leqslant CR. (4.85)

For more details about the proof of the above estimates, we refer to [24, Lemma 3.19].

From (4.12) and (3.23), we notice that

G(ρ,v,ϑ,η)|t=0=(I3𝔸0)ϑ0n=ϑ0non.{G}(\rho,v,\vartheta,\eta)_{|t=0}=\left(I_{3}-\mathbb{A}^{0}\right)\nabla\vartheta^{0}\cdot n=\frac{\partial\vartheta^{0}}{\partial n}\quad\text{on}\ \partial\mathcal{F}. (4.86)

Using the above estimates we deduce that F1,F2,F3,GF_{1},F_{2},F_{3},G and H~\widetilde{H}, H^\widehat{H} defined by (4.9)–(4.12) and (4.76)–(4.77) satisfy the estimate

[F1,F2,F3,G,H~,H^]p,q,βccCR2.\left\|[F_{1},F_{2},F_{3},G,\widetilde{H},\widehat{H}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}\leqslant CR^{2}. (4.87)

To details on the proof of (4.87) can be found in [24, Proposition 3.20]. This shows that Ξ(R)R\Xi(\mathcal{B}_{R})\subset\mathcal{B}_{R} for RR small enough.

To show that ΞR|R{{\Xi}_{R}}|_{\mathcal{B}_{R}} is a strict contraction, we proceed similarly: we consider

[f1(i),f2(i),f3(i),g(i),h~(i),h^(i)]R,i=1,2\left[f_{1}^{(i)},f_{2}^{(i)},f_{3}^{(i)},g^{(i)},\widetilde{h}^{(i)},\widehat{h}^{(i)}\right]^{\top}\in\mathcal{B}_{R},\quad i=1,2

and we denote by [ρ(i),v(i),ϑ(i),η(i),tη(i)][\rho^{(i)},v^{(i)},\vartheta^{(i)},\eta^{(i)},\partial_{t}\eta^{(i)}]^{\top} the solutions to the system (4.20)–(4.20) associated with [f1(i),f2(i),f3(i),g(i),h~(i),h^(i)][f_{1}^{(i)},f_{2}^{(i)},f_{3}^{(i)},g^{(i)},\widetilde{h}^{(i)},\widehat{h}^{(i)}]^{\top} and [ρ0,v0,ϑ0,η10,η20][\rho^{0},v^{0},\vartheta^{0},\eta_{1}^{0},\eta_{2}^{0}]^{\top} (see 4.7). We can thus define

[F1(i),F2(i),F3(i),G(i),H~(i),H^(i)]:=ΞR([f1(i),f2(i),f3(i),g(i),h~(i),h^(i)]).\left[F_{1}^{(i)},F_{2}^{(i)},F_{3}^{(i)},G^{(i)},\widetilde{H}^{(i)},\widehat{H}^{(i)}\right]^{\top}:=\Xi_{R}\left(\left[f_{1}^{(i)},f_{2}^{(i)},f_{3}^{(i)},g^{(i)},\widetilde{h}^{(i)},\widehat{h}^{(i)}\right]^{\top}\right).

We also write

[f1,f2,f3,g,h~,h^]=[f1(1),f2(1),f3(1),g(1),h~(1),h^(1)][f1(2),f2(2),f3(2),g(2),h~(2),h^(2)],[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}=[f_{1}^{(1)},f_{2}^{(1)},f_{3}^{(1)},g^{(1)},\widetilde{h}^{(1)},\widehat{h}^{(1)}]^{\top}-[f_{1}^{(2)},f_{2}^{(2)},f_{3}^{(2)},g^{(2)},\widetilde{h}^{(2)},\widehat{h}^{(2)}]^{\top},
[ρ,v,ϑ,η]=[ρ(1),v(1),ϑ(1),η(1)][ρ(2),v(2),ϑ(2),η(2)],[\rho,v,\vartheta,\eta]^{\top}=[\rho^{(1)},v^{(1)},\vartheta^{(1)},\eta^{(1)}]^{\top}-[\rho^{(2)},v^{(2)},\vartheta^{(2)},\eta^{(2)}]^{\top},
[ρ~,ϑ~]=[ρ~(1),ϑ~(1)][ρ~(2),ϑ~(2)],[ρ^,ϑ^]=[ρ^(1),ϑ^(1)][ρ^(2),ϑ^(2)].[\widetilde{\rho},\widetilde{\vartheta}]^{\top}=[\widetilde{\rho}^{(1)},\widetilde{\vartheta}^{(1)}]^{\top}-[\widetilde{\rho}^{(2)},\widetilde{\vartheta}^{(2)}]^{\top},\quad[\widehat{\rho},\widehat{\vartheta}]^{\top}=[\widehat{\rho}^{(1)},\widehat{\vartheta}^{(1)}]^{\top}-[\widehat{\rho}^{(2)},\widehat{\vartheta}^{(2)}]^{\top}.

Therefore, from 4.7, we obtain

vLβ(0,;W1,q())3+ϑL(0,;W1,q())+ϑLβp(0,;L())3+ηL(0,;W3,q(𝒮))+ηL(0,;C2(𝒮¯))C[f1,f2,f3,g,h~,h^]p,q,βcc\left\|v\right\|_{L^{\infty}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))^{3}}+\left\|\vartheta\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|\nabla\vartheta\right\|_{L^{p}_{\beta}(0,\infty;L^{\infty}(\mathcal{F}))^{3}}\\ +\left\|\eta\right\|_{L^{\infty}(0,\infty;W^{3,q}(\mathcal{S}))}+\left\|\eta\right\|_{L^{\infty}(0,\infty;C^{2}(\overline{\mathcal{S}}))}\leqslant C\left\|[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}} (4.88)

and

ρ~Wβ1,p(0,;W1,q())+ϑ~Wp,q,β1,2((0,);)+[ρ^,ϑ^]L(0,)2+[tρ^,tϑ^]Lβp(0,)2C([f1,f2,f3,g,h~,h^]p,q,βcc).\left\|\widetilde{\rho}\right\|_{W^{1,p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|\widetilde{\vartheta}\right\|_{W^{1,2}_{p,q,\beta}((0,\infty);\mathcal{F})}+\left\|\left[\widehat{\rho},\widehat{\vartheta}\right]^{\top}\right\|_{L^{\infty}(0,\infty)^{2}}+\left\|\left[\partial_{t}\widehat{\rho},\partial_{t}\widehat{\vartheta}\right]^{\top}\right\|_{L^{p}_{\beta}(0,\infty)^{2}}\\ \leqslant C\Big{(}\left\|[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}\Big{)}. (4.89)

In particular, from (3.18),

X(1)X(2)L(0,;W1,q())9C[f1,f2,f3,g,h~,h^]p,q,βcc.\left\|\nabla X^{(1)}-\nabla X^{(2)}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\leqslant C\left\|[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}. (4.90)

By combining the above estimates with (3.5) and with (4.79), we deduce

𝔹X(1)𝔹X(2)L(0,;W1,q())9+t𝔹X(1)t𝔹X(2)Lp(0,;W1,q())9+δX(1)δX(2)L(0,;W1,q())+tδX(1)tδX(2)Lp(0,;W1,q())+1δX(1)1δX(2)L(0,;W1,q())+t(1δX(1))t(1δX(2))Lp(0,;W1,q())+𝔸X(1)𝔸X(2)L(0,;W1,q())9+t𝔸X(1)t𝔸X(2)Lp(0,;W1,q())9+t𝔸X(1)t𝔸X(2)C1/p([0,);W1,q())9CR[f1,f2,f3,g,h~,h^]p,q,βcc.\left\|\mathbb{B}_{X^{(1)}}-\mathbb{B}_{X^{(2)}}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}+\left\|\partial_{t}\mathbb{B}_{X^{(1)}}-\partial_{t}\mathbb{B}_{X^{(2)}}\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\\ +\left\|\delta_{X^{(1)}}-\delta_{X^{(2)}}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|\partial_{t}\delta_{X^{(1)}}-\partial_{t}\delta_{X^{(2)}}\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))}\\ +\left\|\frac{1}{\delta_{X^{(1)}}}-\frac{1}{\delta_{X^{(2)}}}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|\partial_{t}\left(\frac{1}{\delta_{X^{(1)}}}\right)-\partial_{t}\left(\frac{1}{\delta_{X^{(2)}}}\right)\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))}\\ +\left\|\mathbb{A}_{X^{(1)}}-\mathbb{A}_{X^{(2)}}\right\|_{L^{\infty}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}+\left\|\partial_{t}\mathbb{A}_{X^{(1)}}-\partial_{t}\mathbb{A}_{X^{(2)}}\right\|_{L^{p}(0,\infty;W^{1,q}(\mathcal{F}))^{9}}\\ +\left\|\partial_{t}\mathbb{A}_{X^{(1)}}-\partial_{t}\mathbb{A}_{X^{(2)}}\right\|_{C^{1/p^{\prime}}([0,\infty);W^{1,q}(\mathcal{F}))^{9}}\\ \leqslant CR\left\|[f_{1},f_{2},f_{3},g,\widetilde{h},\widehat{h}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}. (4.91)

Using the above estimates we deduce that F1,F2,F3,GF_{1},F_{2},F_{3},G and H~\widetilde{H}, H^\widehat{H} defined by (4.9)–(4.12), (4.76), (4.77) satisfy the estimate

F1(1)F1(2)Lβp(0,;W1,q())+F2(1)F2(2)Lβp(0,;Lq())3+F3(1)F3(2)Lβp(0,;Lq())+G1(1)G1(2)Fp,q,β(11/q)/2(0,;Lq())Lβp(0,;W11/q,q())+H~(1)H~(2)Lβp(0,;Lq(𝒮))+H^(1)H^(2)L(0,)+tH^(1)tH^(2)Lβp(0,)CR[f1,f2,f3,g,h]p,q,βcc.\left\|F_{1}^{(1)}-F_{1}^{(2)}\right\|_{L^{p}_{\beta}(0,\infty;W^{1,q}(\mathcal{F}))}+\left\|F_{2}^{(1)}-F_{2}^{(2)}\right\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))^{3}}+\left\|F_{3}^{(1)}-F_{3}^{(2)}\right\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{F}))}\\ +\left\|G_{1}^{(1)}-G_{1}^{(2)}\right\|_{F^{(1-1/q)/2}_{p,q,\beta}(0,\infty;L^{q}(\partial\mathcal{F}))\cap L^{p}_{\beta}(0,\infty;W^{1-1/q,q}(\partial\mathcal{F}))}+\left\|\widetilde{H}^{(1)}-\widetilde{H}^{(2)}\right\|_{L^{p}_{\beta}(0,\infty;L^{q}(\mathcal{S}))}\\ +\left\|\widehat{H}^{(1)}-\widehat{H}^{(2)}\right\|_{L^{\infty}(0,\infty)}+\left\|\partial_{t}\widehat{H}^{(1)}-\partial_{t}\widehat{H}^{(2)}\right\|_{L^{p}_{\beta}(0,\infty)}\\ \leqslant CR\left\|[f_{1},f_{2},f_{3},g,{h}]^{\top}\right\|_{\mathcal{R}_{p,q,\beta}^{cc}}. (4.92)

This shows that Ξ|R\Xi_{|\mathcal{B}_{R}} is a strict contraction, for RR small enough. This completes the proof of 4.1 and 1.2. ∎

References

  • [1] R. A. Adams and J. J. F. Fournier, Sobolev spaces, vol. 140 of Pure and Applied Mathematics (Amsterdam), Elsevier/Academic Press, Amsterdam, second ed., 2003.
  • [2] H. Amann, Ordinary differential equations, vol. 13 of De Gruyter Studies in Mathematics, Walter de Gruyter & Co., Berlin, 1990. An introduction to nonlinear analysis, Translated from the German by Gerhard Metzen.
  • [3]  , Linear and quasilinear parabolic problems. Vol. I, vol. 89 of Monographs in Mathematics, Birkhäuser Boston, Inc., Boston, MA, 1995. Abstract linear theory.
  • [4] G. Avalos, P. G. Geredeli, and J. T. Webster, Semigroup well-posedness of a linearized, compressible fluid with an elastic boundary, Discrete Contin. Dyn. Syst. Ser. B, 23 (2018), pp. 1267–1295.
  • [5] M. Badra and T. Takahashi, Gevrey Regularity for a System Coupling the Navier–Stokes System with a Beam Equation, SIAM J. Math. Anal., 51 (2019), pp. 4776–4814.
  • [6] M. Badra and T. Takahashi, Gevrey regularity for a system coupling the Navier-Stokes system with a beam: the non-flat case, https://hal.archives-ouvertes.fr/hal-02303258, (2019).
  • [7] T. Bodnár, G. P. Galdi, and v. Nečasová, eds., Fluid-structure interaction and biomedical applications, Advances in Mathematical Fluid Mechanics, Birkhäuser/Springer, Basel, 2014.
  • [8] M. Bravin, Energy equality and uniqueness of weak solutions of a “viscous incompressible fluid + rigid body” system with Navier slip-with-friction conditions in a 2D bounded domain, J. Math. Fluid Mech., 21 (2019), pp. Art. 23, 31.
  • [9] D. Breit and S. Schwarzacher, Compressible fluids interacting with a linear-elastic shell, Arch. Ration. Mech. Anal., 228 (2018), pp. 495–562.
  • [10] I. Chueshov, Dynamics of a nonlinear elastic plate interacting with a linearized compressible viscous fluid, Nonlinear Anal., 95 (2014), pp. 650–665.
  • [11] R. Denk, M. Hieber, and J. Prüss, \mathscr{R}-boundedness, Fourier multipliers and problems of elliptic and parabolic type, Mem. Amer. Math. Soc., 166 (2003), pp. viii+114.
  • [12]  , Optimal LpL^{p}-LqL^{q}-estimates for parabolic boundary value problems with inhomogeneous data, Math. Z., 257 (2007), pp. 193–224.
  • [13] R. Denk and J. Saal, LpL^{p}-theory for a fluid-structure interaction model, arXiv e-prints, (2019), p. arXiv:1909.09344.
  • [14] R. Denk and R. Schnaubelt, A structurally damped plate equation with Dirichlet-Neumann boundary conditions, J. Differential Equations, 259 (2015), pp. 1323–1353.
  • [15] G. Dore, LpL^{p} regularity for abstract differential equations, in Functional analysis and related topics, 1991 (Kyoto), vol. 1540 of Lecture Notes in Math., Springer, Berlin, 1993, pp. 25–38.
  • [16] Y. Enomoto and Y. Shibata, On the \mathscr{R}-sectoriality and the initial boundary value problem for the viscous compressible fluid flow, Funkcial. Ekvac., 56 (2013), pp. 441–505.
  • [17] S. Ervedoza, M. Hillairet, and C. Lacave, Long-time behavior for the two-dimensional motion of a disk in a viscous fluid, Comm. Math. Phys., 329 (2014), pp. 325–382.
  • [18] S. Ervedoza, D. Maity, and M. Tucsnak, Large time behaviour for the motion of a solid in a viscous incompressible fluid, https://hal.archives-ouvertes.fr/hal-02545798, (2020).
  • [19] F. Flori and P. Orenga, Fluid-structure interaction: analysis of a 3-D compressible model, Ann. Inst. H. Poincaré Anal. Non Linéaire, 17 (2000), pp. 753–777.
  • [20] M. Geissert, K. Götze, and M. Hieber, LpL^{p}-theory for strong solutions to fluid-rigid body interaction in Newtonian and generalized Newtonian fluids, Trans. Amer. Math. Soc., 365 (2013), pp. 1393–1439.
  • [21] O. Glass and F. Sueur, Uniqueness results for weak solutions of two-dimensional fluid-solid systems, Arch. Ration. Mech. Anal., 218 (2015), pp. 907–944.
  • [22] C. Grandmont, M. Hillairet, and J. Lequeurre, Existence of local strong solutions to fluid–beam and fluid–rod interaction systems, in Annales de l’Institut Henri Poincaré C, Analyse non linéaire, vol. 36, Elsevier, 2019, pp. 1105–1149.
  • [23] C. Grandmont, M. Lukáčová-Medvid’ová, and v. Nečasová, Mathematical and numerical analysis of some FSI problems, in Fluid-structure interaction and biomedical applications, Adv. Math. Fluid Mech., Birkhäuser/Springer, Basel, 2014, pp. 1–77.
  • [24] B. H. Haak, D. Maity, T. Takahashi, and M. Tucsnak, Mathematical analysis of the motion of a rigid body in a compressible Navier-Stokes-Fourier fluid, Math. Nachr., 292 (2019), pp. 1972–2017.
  • [25] M. Hieber and M. Murata, The LpL^{p}-approach to the fluid-rigid body interaction problem for compressible fluids, Evol. Equ. Control Theory, 4 (2015), pp. 69–87.
  • [26] P. C. Kunstmann and L. Weis, Perturbation theorems for maximal LpL_{p}--regularity, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 30 (2001), pp. 415–435.
  • [27] P. C. Kunstmann and L. Weis, Maximal LpL_{p}-regularity for parabolic equations, Fourier multiplier theorems and HH^{\infty}-functional calculus, in Functional analytic methods for evolution equations, vol. 1855 of Lecture Notes in Math., Springer, Berlin, 2004, pp. 65–311.
  • [28] C. Lacave and T. Takahashi, Small moving rigid body into a viscous incompressible fluid, Arch. Ration. Mech. Anal., 223 (2017), pp. 1307–1335.
  • [29] J. Lequeurre, Existence of strong solutions to a fluid-structure system, SIAM J. Math. Anal., 43 (2011), pp. 389–410.
  • [30] D. Maity and T. Takahashi, Lp theory for the interaction between the incompressible Navier-Stokes system and a damped beam, https://hal.archives-ouvertes.fr/hal-02294097, (2020).
  • [31] D. Maity and M. Tucsnak, A maximal regularity approach to the analysis of some particulate flows, in Particles in flows, Adv. Math. Fluid Mech., Birkhäuser/Springer, Cham, 2017, pp. 1–75.
  • [32] D. Maity and M. Tucsnak, LpL^{p}-LqL^{q} maximal regularity for some operators associated with linearized incompressible fluid-rigid body problems Selected Recent Results, in Mathematical analysis in fluid mechanics selected recent results, vol. 710 of Contemp. Math., Amer. Math. Soc., Providence, RI, 2018, pp. 175–201.
  • [33] S. Mitra, Local existence of Strong solutions for a fluid-structure interaction model, arXiv e-prints, (2018), p. arXiv:1808.06716.
  • [34] J. Prüss and G. Simonett, Moving interfaces and quasilinear parabolic evolution equations, vol. 105 of Monographs in Mathematics, Birkhäuser/Springer, [Cham], 2016.
  • [35] Y. Shibata and K. Tanaka, On a resolvent problem for the linearized system from the dynamical system describing the compressible viscous fluid motion, Math. Methods Appl. Sci., 27 (2004), pp. 1579–1606.
  • [36] R. Temam, Navier-Stokes equations, vol. 2 of Studies in Mathematics and its Applications, North-Holland Publishing Co., Amsterdam-New York, revised ed., 1979. Theory and numerical analysis, With an appendix by F. Thomasset.
  • [37] H. Triebel, Interpolation theory, function spaces, differential operators, Johann Ambrosius Barth, Heidelberg, second ed., 1995.
  • [38]  , Theory of function spaces, Modern Birkhäuser Classics, Birkhäuser/Springer Basel AG, Basel, 2010. Reprint of 1983 edition [MR0730762], Also published in 1983 by Birkhäuser Verlag [MR0781540].
  • [39] M. Tucsnak and G. Weiss, Observation and control for operator semigroups, Birkhäuser Advanced Texts: Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks], Birkhäuser Verlag, Basel, 2009.
  • [40] L. Weis, Operator-valued Fourier multiplier theorems and maximal LpL_{p}-regularity, Math. Ann., 319 (2001), pp. 735–758.
  • [41] J. Wloka, Partial differential equations, Cambridge University Press, Cambridge, 1987. Translated from the German by C. B. Thomas and M. J. Thomas.