1 Introduction

In this work, we study the interaction between a viscous incompressible fluid and a deformable structure located on a part of the fluid domain boundary. More precisely, we denote by \({\mathcal {F}}\) the reference domain for the fluid. We assume that it is a smooth bounded domain of \({\mathbb {R}}^3\) such that its boundary \(\partial {\mathcal {F}}\) contains a flat part \(\Gamma _{S}\) corresponding to the reference domain of the plate. We assume \(\Gamma _{S} = {\mathcal {S}} \times \{0\}\), where \({\mathcal {S}}\) is a smooth domain of \({\mathbb {R}}^2\) and we set \(\Gamma _{0} : = \partial {\mathcal {F}} {\setminus } \overline{\Gamma _{S}}.\) The set \(\Gamma _0\) is rigid and remains unchanged whereas the plate domain \(\Gamma _{S}\) can deform through exterior forces and in particular the force coming from the fluid and if we denote by \(\eta \) its displacement, then the plate domain changes from \(\Gamma _{S}\) to

$$\begin{aligned} \Gamma _{S}(\eta ) := \left\{ (s, \eta (s)) \ ; \ s \in {\mathcal {S}} \right\} . \end{aligned}$$

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

$$\begin{aligned} \eta = \nabla _s \eta \cdot n_{{\mathcal {S}}} = 0 \quad \text {on}\ \partial {\mathcal {S}} \end{aligned}$$
(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:

$$\begin{aligned} \Gamma _{0} \cap \Gamma _{S}(\eta ) = \emptyset . \end{aligned}$$
(1.2)

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

With the above notations and hypotheses, \(\Gamma _{0} \cup \overline{\Gamma _{S}(\eta )}\) corresponds to a closed simple and regular surface which interior is the fluid domain \({\mathcal {F}}(\eta )\). In what follows, we consider that \(\eta \) is also a function of time and its evolution is governed by a plate equation. If \(\eta (t,\cdot )\) satisfies the above conditions, we can define the fluid domain \({\mathcal {F}}(\eta (t))\) and we then denote by \((\widetilde{v}, \widetilde{\pi })\) the Eulerian velocity and the pressure of the fluid and we assume that they satisfy the incompressible Navier-Stokes system in \({\mathcal {F}}(\eta (t))\). Then the corresponding system we analyze reads as follows:

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _{t} \widetilde{v} + (\widetilde{v} \cdot \nabla ) \widetilde{v} - \mathrm{div}\,{\mathbb {T}}(\widetilde{v}, \widetilde{\pi }) = 0, &{} t> 0, x \in {\mathcal {F}}(\eta (t)), \\ \mathrm{div}\,\widetilde{v} = 0 &{} t> 0, x \in {\mathcal {F}}(\eta (t)), \\ \widetilde{v}(t, s, \eta (t,s)) = \partial _{t} \eta (t,s) e_{3} &{} t> 0, s \in {\mathcal {S}}, \\ \widetilde{v} = 0 &{} t> 0, x \in \Gamma _{0}, \\ \partial _{tt} \eta + \alpha \Delta _{s}^{2} \eta -\beta \Delta _{s} \eta - \gamma \Delta _{s} \partial _{t}\eta = {\mathbb {H}}(\widetilde{v},\widetilde{\pi }, \eta ) &{} t> 0, s \in {\mathcal {S}}, \\ \eta = \nabla _s \eta \cdot n_{{\mathcal {S}}} = 0 &{} t > 0, s \in \partial {\mathcal {S}}, \end{array}\right. } \end{aligned}$$
(1.3)

where \((e_1,e_3,e_{3})\) is the canonical basis of \({\mathbb {R}}^3\). The fluid stress tensor \({\mathbb {T}}(\widetilde{v},\widetilde{\pi })\) is given by

$$\begin{aligned} {\mathbb {T}}(\widetilde{v},\widetilde{\pi }) = 2\nu D(\widetilde{v}) - \widetilde{\pi } I_3, \quad D(\widetilde{v}) = \frac{1}{2} \left( \nabla \widetilde{v} + \nabla \widetilde{v}^{\top }\right) . \end{aligned}$$
(1.4)

The function \({\mathbb {H}}\) corresponds to the force of the fluid acting on the plate and can be expressed as follows:

$$\begin{aligned} {\mathbb {H}}(\widetilde{v},\widetilde{\pi }, \eta ) =- \sqrt{1 + \left| \nabla _s \eta \right| ^{2}} \left( {\mathbb {T}}(\widetilde{v},\widetilde{\pi }) \widetilde{n} \right) |_{\Gamma _{S}(\eta (t))} \cdot e_{3}, \end{aligned}$$
(1.5)

where

$$\begin{aligned} \widetilde{n} = \frac{1}{\sqrt{1 + \left| \nabla _s \eta \right| ^{2}}} \left[ -\nabla _s \eta , 1 \right] ^{\top }, \end{aligned}$$

is the unit normal to \(\Gamma _{S}(\eta (t))\) outward \({\mathcal {F}}(\eta (t)).\) The above system is completed by the following initial data

$$\begin{aligned} \eta (0,\cdot ) = \eta _{1}^{0} \text{ in } {\mathcal {S}}, \quad \partial _{t}\eta (0,\cdot ) = \eta _{2}^{0} \text{ in } {\mathcal {S}}, \quad \widetilde{v}(0,\cdot ) = \widetilde{v}^{0} \text{ in } {\mathcal {F}}(\eta _{1}^{0}). \end{aligned}$$
(1.6)

System (1.3) is a simplified model for blood flow in arteries (see, for instance the survey article [38]) and \(\alpha ,\beta ,\gamma \) are non negative constants that corresponds to the physical properties of the wall tissue. Our analysis will be done in the case \(\alpha >0\), \(\beta \geqslant 0\) and \(\gamma >0\) and to simplify, we consider in what follows the case

$$\begin{aligned} \alpha =1, \quad \beta =0,\quad \gamma =1, \end{aligned}$$

and the other cases can be done in the same way. Let us remark that the term \(- \gamma \Delta _{s} \partial _{t}\eta \) corresponds to the damping in the plate equation. The other positive constant, appearing in (1.4) is the viscosity \(\nu \).

An important remark in the study of (1.3)–(1.6) is that a solution \((\widetilde{v}, \widetilde{\pi }, \eta )\) satisfies

$$\begin{aligned} 0 = \int _{{\mathcal {F}}(\eta (t))} \mathrm{div}\,\widetilde{v} \ dx =\int _{\Gamma _{S}(\eta (t))} \widetilde{v} \cdot \widetilde{n} \ d\Gamma = \frac{d}{dt} \int _{{\mathcal {S}}} \eta \ ds. \end{aligned}$$

Assuming that \(\eta _1^0\) has a zero mean, we deduce that this property is preserved for \(\eta \) all along. This leads us to consider the space

$$\begin{aligned} L^{q}_{m}({\mathcal {S}}) = \left\{ f \in L^{q}({\mathcal {S}}) \ ; \ \int _{{\mathcal {S}}} f \ ds = 0 \right\} , \end{aligned}$$
(1.7)

and the orthogonal projection \(P_{m} : L^{q}({\mathcal {S}}) \rightarrow L^{q}_{m}({\mathcal {S}})\), that is

$$\begin{aligned} P_{m} f = f - \frac{1}{|{\mathcal {S}}|}\int _{{\mathcal {S}}} f \ ds \qquad (f \in L^{q}({\mathcal {S}})). \end{aligned}$$
(1.8)

Taking the projection of the plate equation in (1.3) onto \(L^{q}_{m}({\mathcal {S}})\) and onto \(L^{q}_{m}({\mathcal {S}})^\perp \) yields the following two equations:

$$\begin{aligned} \partial _{tt} \eta + P_{m}\Delta _{s}^{2} \eta - \Delta _{s} \partial _{t}\eta = P_{m}\left( {\mathbb {H}}(\widetilde{v}, \widetilde{\pi }, \eta ) \right) \qquad t > 0, s \in {\mathcal {S}}, \end{aligned}$$
(1.9)

and

$$\begin{aligned} \int _{{\mathcal {S}}} \widetilde{\pi }(t,s,\eta (t,s)) \; ds = \int _{{\mathcal {S}}} \Delta _{s}^{2} \eta (t,s)\; ds + \int _{{\mathcal {S}}} \sqrt{1 + \left| \nabla _s \eta \right| ^{2}} \left[ (2\nu D\widetilde{v}) \widetilde{n} \right] (t,s,\eta (t,s)) \cdot e_{3} \; ds. \end{aligned}$$
(1.10)

This means that, in contrast to the Navier–Stokes system without structure, the pressure is not determined up to a constant. In what follows, we only keep (1.9) and solve the corresponding system up to constant for the pressure, and Eq. (1.10) is used at the end to fix the constant for the pressure. We thus consider the following system

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _{t}\widetilde{v} + (\widetilde{v} \cdot \nabla ) \widetilde{v} - \mathrm{div}\,{\mathbb {T}}(\widetilde{v}, \widetilde{\pi }) =0&{} t> 0, x \in {\mathcal {F}}(\eta (t)), \\ \mathrm{div}\,\widetilde{v} = 0 &{} t> 0, x \in {\mathcal {F}}(\eta (t)), \\ \widetilde{v}(t, s, \eta (t,s)) = \partial _{t} \eta (t,s) e_{3} &{} t> 0, s \in {\mathcal {S}}, \\ \widetilde{v} = 0 &{} t> 0, x \in \Gamma _{0}, \\ \partial _{tt} \eta + P_{m}\Delta _{s}^{2} \eta - \Delta _{s} \partial _{t}\eta = P_{m}{\mathbb {H}}(\widetilde{v},\widetilde{\pi }, \eta ) \qquad &{} t> 0, s \in {\mathcal {S}},\\ \eta = \nabla _s \eta \cdot n_{{\mathcal {S}}} = 0 &{} t > 0, s \in \partial {\mathcal {S}},\\ \eta (0,\cdot ) = \eta _{1}^{0} \text{ in } {\mathcal {S}}, \quad \partial _{t}\eta (0,\cdot ) = \eta _{2}^{0} \text{ in } {\mathcal {S}}, \quad \widetilde{v}(0,\cdot ) = \widetilde{v}^{0} \text{ in } {\mathcal {F}}(\eta _{1}^{0}). \end{array}\right. } \end{aligned}$$
(1.11)

To state our main result, we introduce some notations for our functional spaces. Firstly \(W^{s,q}(\Omega )\), with \(s\geqslant 0\) and \(q\geqslant 1\), denotes the usual Sobolev space. Let \(k,k' \in {\mathbb {N}}\), \(k<k'\). For \(1\leqslant p < \infty \), \(1\leqslant q < \infty \), we consider the standard definition of the Besov spaces by real interpolation of Sobolev spaces

$$\begin{aligned} B^{s}_{q,p}({\mathcal {F}}) = \left( W^{k,q}({\mathcal {F}}), W^{k',q}({\mathcal {F}})\right) _{\theta , p} \text{ where } s = (1 - \theta ) k + \theta k', \quad \theta \in (0,1). \end{aligned}$$

We refer to [1] and [44] for a detailed presentation of the Besov spaces. We also introduce functional spaces for the fluid velocity and pressure for a spatial domain depending on the displacement \(\eta \) of the structure. Let \(1< p, q < \infty \) and \(\eta \in L^{p}(0,\infty ;W^{4,q}({\mathcal {S}})) \cap W^{2,p}(0,\infty ;L^{q}({\mathcal {S}}))\) satisfying (1.1) and (1.2). We show in Sect. 2 that there exists a mapping \(X=X_\eta \) such that \(X(t,\cdot )\) is a \(C^{1}\)-diffeomorphism from \({\mathcal {F}}\) onto \({\mathcal {F}}(\eta (t))\) and such that \(X\in L^{p}(0,\infty ;W^{2,q}({\mathcal {F}})) \cap W^{2,p}(0,\infty ;L^{q}({\mathcal {F}}))\). Then for \(T \in (0, \infty ]\), we define

$$\begin{aligned}&L^{p}(0,T;L^{q}({\mathcal {F}}(\eta (\cdot ) ))) := \left\{ v \circ X^{-1} \ ; \ v \in L^{p}(0,T;L^{q}({\mathcal {F}}))\right\} , \\&L^{p}(0,T;W^{2,q}({\mathcal {F}}(\eta (\cdot )))) := \left\{ v \circ X^{-1} \ ; \ v \in L^{p}(0,T;W^{2,q}({\mathcal {F}}))\right\} , \\&W^{1,p}(0,T;L^{q}({\mathcal {F}}(\eta (\cdot )))) := \left\{ v \circ X^{-1} \ ; \ v \in W^{1,p}(0,T;L^{q}({\mathcal {F}}))\right\} , \\&C^0([0,T];W^{1,q}({\mathcal {F}}(\eta (\cdot )))) := \left\{ v \circ X^{-1} \ ; \ v \in C^0([0,T];W^{1,q}({\mathcal {F}}))\right\} , \\&C^0([0,T];B^{2(1-1/p)}_{q,p}({\mathcal {F}}(\eta (\cdot )))) := \left\{ v \circ X^{-1} \ ; \ v \in C^0([0,T];B^{2(1-1/p)}_{q,p}({\mathcal {F}}))\right\} , \end{aligned}$$

where we have set \((v\circ X^{-1})(t,x):=v(t,(X(t,\cdot ))^{-1}(x))\) for simplicity.

Finally, let us give the conditions we need on the initial conditions for the system (1.11): we assume

$$\begin{aligned} \eta _{1}^{0} \in B^{2(2-1/p)}_{q,p}({\mathcal {S}}), \quad \eta _{2}^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {S}}), \quad \widetilde{v}^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {F}}(\eta _{1}^{0})) \end{aligned}$$
(1.12)

with the compatibility conditions

$$\begin{aligned} \eta _{1}^{0} = \nabla _s \eta _{1}^{0}\cdot n_{{\mathcal {S}}} = 0 \quad \text{ on } \ \partial {\mathcal {S}}, \quad \Gamma _{0} \cap \Gamma _{S}(\eta _{1}^{0}) = \emptyset , \quad \int _{{\mathcal {S}}} \eta _{1}^{0} \ ds = 0, \quad \int _{{\mathcal {S}}} \eta _{2}^{0} \ ds = 0, \quad \mathrm{div}\,\widetilde{v}^{0} = 0 \quad \text{ in } \ {\mathcal {F}}(\eta _{1}^{0}), \end{aligned}$$
(1.13)

and

$$\begin{aligned} {\left\{ \begin{array}{ll} \widetilde{v}^0(s, \eta _1^0(s))\cdot \widetilde{n}^0 = \eta _2^0(s) e_{3}\cdot \widetilde{n}^0 \quad s \in {\mathcal {S}}, \quad \widetilde{v}^{0} \cdot \widetilde{n}^0 = 0 \quad \text{ on } \ \Gamma _{0} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q} > 1, \\ \widetilde{v}^0(s, \eta _1^0(s)) = \eta _2^0(s) e_{3} \quad s \in {\mathcal {S}}, \quad \widetilde{v}^{0} = 0 \quad \text{ on } \ \Gamma _{0}, \quad \eta _{2}^{0} = 0 \quad \text{ on } \ \partial {\mathcal {S}} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q}< 1, \\ \nabla _{s} \eta _{2}^{0} \cdot n_{S} = 0 \text{ on } \partial {\mathcal {S}} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q} < \frac{1}{2}. \end{array}\right. } \end{aligned}$$
(1.14)

Here \(\widetilde{n}^0\) is the unit exterior normal to \(\Gamma _{S}(\eta _1^0)\) outward \({\mathcal {F}}(\eta _1^0).\)

We are now in a position to state our main results. The first one is the local in time existence and uniqueness of strong solutions for (1.11).

Theorem 1.1

Let \(p,q\in (1,\infty )\) such that

$$\begin{aligned} \frac{1}{p} + \frac{1}{2q} \ne 1, \quad \frac{1}{p} + \frac{1}{2q} \ne \frac{1}{2} \quad \text {and} \quad \frac{1}{p}+\frac{3}{2q}<\frac{3}{2}. \end{aligned}$$
(1.15)

Let us assume that \(\eta _{1}^{0} = 0\) and \((\eta _2^0,\widetilde{v}^0)\) satisfies (1.12), (1.13), (1.14). Then there exists \(T > 0,\) depending only on \((\eta _2^0,\widetilde{v}^0),\) such that the system (1.11) admits a unique strong solution \((\widetilde{v}, \widetilde{\pi }, \eta )\) in the class of functions satisfying

$$\begin{aligned}&\widetilde{v} \in L^{p}(0,T;W^{2,q}({\mathcal {F}}(\eta (\cdot )))) \cap L^{\infty }(0,T;B^{2(1-1/p)}_{q,p}({\mathcal {F}}(\eta (\cdot )))) \cap W^{1,p}(0,T;L^{q}({\mathcal {F}}(\eta (\cdot )))), \\&\widetilde{\pi } \in L^{p}(0,T;W^{1,q}_{m}({\mathcal {F}}(\eta (\cdot )))), \\&\eta \in L^{p}(0,T;W^{4,q}({\mathcal {S}})) \cap L^{\infty }(0,T;B^{2(2-1/p)}_{q,p}({\mathcal {S}})) \cap W^{1,p}(0,T;W^{2,q}({\mathcal {S}})),\\&\partial _t \eta \in L^{p}(0,T;W^{2,q}({\mathcal {S}})) \cap L^{\infty }(0,T;B^{2(1-1/p)}_{q,p}({\mathcal {S}})) \cap W^{1,p}(0,T;L^{q}({\mathcal {S}})). \end{aligned}$$

Moreover, \(\Gamma _{0} \cap \Gamma _{S}(\eta (t)) = \emptyset \) for all \(t \in [0,T].\)

Our second main result asserts the global existence and uniqueness of strong solution for (1.11) under a smallness condition on the initial data.

Theorem 1.2

Let \(p,q\in (1,\infty )\) satisfying the conditions (1.15). Then there exists \(\beta _{0} > 0\) such that, for all \(\beta \in [0,\beta _{0}]\) there exists \(\varepsilon _{0}\) such that for any \((\eta _1^0,\eta _2^0,\widetilde{v}^0)\) satisfying (1.12), (1.13), (1.14) and

$$\begin{aligned} \Vert \widetilde{v}^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {F}} (\eta _{1}^{0}))}+ \Vert \eta _{1}^{0}\Vert _{B^{2(2-1/p)}_{q,p}({\mathcal {S}})} +\Vert \eta _{2}^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {S}})} <\varepsilon _0, \end{aligned}$$
(1.16)

the system (1.11) admits a unique strong solution \((\widetilde{v}, \widetilde{\pi }, \eta )\) in the class of functions satisfying

$$\begin{aligned}&\widetilde{v} \in L^{p}_\beta (0,\infty ;W^{2,q}({\mathcal {F}}(\eta (\cdot )))) \cap L^{\infty }_\beta (0,\infty ;B^{2(1-1/p)}_{q,p}({\mathcal {F}}(\eta (\cdot )))) \cap W^{1,p}_\beta (0,\infty ;L^{q}({\mathcal {F}}(\eta (\cdot )))), \\&\widetilde{\pi } \in L^{p}_\beta (0,\infty ;W^{1,q}_{m} ({\mathcal {F}}(\eta (\cdot )))), \\&\eta \in L^{p}_\beta (0,\infty ;W^{4,q}({\mathcal {S}})) \cap L^{\infty }_\beta (0,\infty ;B^{2(2-1/p)}_{q,p}({\mathcal {S}})) \cap W^{1,p}_\beta (0,\infty ;W^{2,q}({\mathcal {S}})),\\&\partial _t \eta \in L^{p}_\beta (0,\infty ;W^{2,q}({\mathcal {S}})) \cap L^{\infty }_\beta (0,\infty ;B^{2(1-1/p)}_{q,p}({\mathcal {S}})) \cap W^{1,p}_\beta (0,\infty ;L^{q}({\mathcal {S}})). \end{aligned}$$

Moreover, \(\Gamma _{0} \cap \Gamma _{S}(\eta (t)) = \emptyset \) for all \(t \in [0,\infty ).\)

In the above statement, we have used a similar notation as in (1.7):

$$\begin{aligned} L^{q}_{m}({\mathcal {F}}) := \left\{ f \in L^{q}({\mathcal {F}}) \ ; \ \int _{{\mathcal {F}}} f = 0 \ dx \right\} , \quad W^{s,q}_{m}({\mathcal {F}}) := W^{s,q}({\mathcal {F}}) \cap L^{q}_{m}({\mathcal {F}}). \end{aligned}$$

We also set

$$\begin{aligned} W^{s,q}_{m}({\mathcal {S}}) = W^{s,q}({\mathcal {S}}) \cap L^{q}_{m}({\mathcal {S}}). \end{aligned}$$

We denote by \(W^{s,q}_0({\mathcal {S}})\) the closure of \(C^{\infty }_c({\mathcal {S}})\) in \(W^{s,q}({\mathcal {S}})\) and we set

$$\begin{aligned} W^{s,q}_{0,m}({\mathcal {S}})=W^{s,q}_0(\Omega )\cap L^q_m({\mathcal {S}}). \end{aligned}$$

We define similarly \(W^{s,q}_0({\mathcal {F}})\), \(W^{s,q}_{0,m}({\mathcal {F}})\).

Finally, we also need the following notation in what follows: for \(T\in (0,\infty ]\),

$$\begin{aligned}&W^{1,2}_{p,q}((0,T) ; {\mathcal {F}}) =L^{p}(0,T;W^{2,q}({\mathcal {F}})) \cap W^{1,p}(0,T;L^{q}({\mathcal {F}})), \\&W^{2,4}_{p,q}((0,T) ; {\mathcal {S}}) =L^{p}(0,T;W^{4,q}({\mathcal {S}})) \cap W^{1,p}(0,T;W^{2,q}({\mathcal {S}})) \cap W^{2,p}(0,T;L^{q}({\mathcal {S}})), \\&W^{1,2}_{p,q}((0,T) ; {\mathcal {S}}) =L^{p}(0,T;W^{2,q}({\mathcal {S}})) \cap W^{1,p}(0,T;L^{q}({\mathcal {S}})). \end{aligned}$$

We have the following embeddings (see, for instance, [2, Theorem 4.10.2, p.180]),

$$\begin{aligned}&W^{1,2}_{p,q}((0,T) ; {\mathcal {F}}) \hookrightarrow C_{b}^0([0,T);B^{2(1-1/p)}_{q,p}({\mathcal {F}})), \end{aligned}$$
(1.17)
$$\begin{aligned}&W^{2,4}_{p,q}((0,T) ; {\mathcal {S}}) \hookrightarrow C_{b}^0([0,T);B^{2(2-1/p)}_{q,p}({\mathcal {S}})) \cap C_{b}^1([0,T);B^{2(1-1/p)}_{q,p}({\mathcal {S}})) \end{aligned}$$
(1.18)

where \(C^k_b\) is the set of continuous and bounded functions with derivatives continuous and bounded up to the order k. In particular, in what follows, we use the following norm for \(W^{1,2}_{p,q}((0,T); {\mathcal {F}})\):

$$\begin{aligned} \Vert f\Vert _{W^{1,2}_{p,q}((0,T) ; {\mathcal {F}})} :=\Vert f\Vert _{L^{p}(0,T;W^{2,q}({\mathcal {F}}))} +\Vert f\Vert _{W^{1,p}(0,T;L^{q}({\mathcal {F}}))} +\Vert f\Vert _{C_{b}^0([0,T);B^{2(1-1/p)}_{q,p}({\mathcal {F}}))} \end{aligned}$$

and we proceed similarly for the two other spaces.

For \(\beta \geqslant 0\), \(p\in [1,\infty ]\) and for \({\mathbb {X}}\) a Banach space, we also introduce the notation

$$\begin{aligned} L^p_\beta (0,\infty ;{\mathbb {X}}) :=\left\{ f \ ; \ t\mapsto e^{\beta t} f(t) \in L^p (0,\infty ;{\mathbb {X}}) \right\} , \end{aligned}$$

and a similar notation for \(W^{1,2}_{p,q,\beta }((0,\infty ) ; {\mathcal {F}})\), \(W^{2,4}_{p,q,\beta }((0,\infty ) ; {\mathcal {S}})\), etc.

Let us give some remarks on Theorems 1.1 and 1.2. First let us point out that the system (1.11) has already been studied by several authors: existence of weak solutions ([9, 26, 37]), uniqueness of weak solutions ([25]), existence of strong solutions ([7, 32, 34]), feedback stabilization ([5, 40]), global existence of strong solutions and study of the contacts ([22]). Some works consider also the case of a beam/plate without damping (that is without the term \( - \Delta _{s} \partial _{t}\eta \)): [6, 21, 23]. We refer, for instance, to [24] and references therein for a concise description of recent progress in this field. It is important to notice that all the above works correspond to a “Hilbert” framework whereas our results are done in a “\(L^{p}\)-\(L^{q}\)” framework. Working in such a framework allows us to extend the result obtained in the “Hilbert” framework, but it should be noticed that several questions on fluid-structure interaction systems, in the “Hilbert” framework, have been handled by considering a “\(L^{p}\)-\(L^{q}\)” framework: for instance, the uniqueness of weak solutions (see [8, 20]), the asymptotic behavior for large time (see [16]), the asymptotic behavior for small structures (see [31]), etc.

For this approach, several recent results have been obtained for fluid systems, with or without structure. For instance, one can quote [19] (viscous incompressible fluid), [15], (viscous compressible fluid), [27, 28] (viscous compressible fluid with rigid bodies), [18, 35] (incompressible viscous fluid and rigid bodies). Here we consider an incompressible viscous fluid coupled with a structure satisfying an infinite-dimensional system and we thus need to go beyond the theory developed for instance in [35].

Our approach to prove Theorems 1.1 and 1.2 is quite classical. Since the fluid domain \({\mathcal {F}}(\eta (t))\) depends on the structure displacement \(\eta ,\) we first reformulate the problem in a fixed domain. This is achieved by “geometric” change of variables. Next we associate the original nonlinear problem to a linear one. The linear system preserves the fluid-structure coupling. A crucial step here is to establish the \(L^{p}\)-\(L^{q}\) regularity property in the infinite time horizon. This is done by showing that the associate linear operator is \({\mathcal {R}}\)-sectorial and generates an exponentially stable semigroup. We then use the Banach fixed point theorem to prove existence and uniqueness results. Note that for Theorem 1.2, we assume the same conditions on (pq) than for Theorem 1.1 but the result should be also true for \(\frac{1}{p}+\frac{3}{2q}=\frac{3}{2}.\) However to deal with this case one needs some precise results on the interpolation of Besov spaces (see for instance Lemma 2.1).

Let us also remark that this work could also be done in the corresponding 2D/1D model, that is \({\mathcal {F}}\) a regular bounded domain in \({\mathbb {R}}^2\) such that \(\partial {\mathcal {F}}\) contains a flat part \(\Gamma _S={\mathcal {S}}\times \{0\}\), where \({\mathcal {S}}\) is an open bounded interval of \({\mathbb {R}}\). In that case, we would obtain the same result as in Theorem 1.1 and in Theorem 1.2 but with the following condition on pq:

$$\begin{aligned} \frac{1}{p} + \frac{1}{2q} \ne 1, \quad \frac{1}{p} + \frac{1}{2q} \ne \frac{1}{2} \quad \text {and} \quad \frac{1}{p}+\frac{1}{q}<\frac{3}{2}. \end{aligned}$$

The plan of the paper is as follows. In the next section, we use a change of variables to rewrite the governing equations in a cylindrical domain and we also restate our result after change of variables. Then, in Sect. 3, we recall several important results about maximal \(L^{p}\) regularity for Cauchy problems and in particular how to use the \({\mathcal {R}}\)-sectoriality property. We use these results to study in Sect. 4 the linearized system. Finally in Sect. 5 and in Sect. 6, we estimate the nonlinear terms which allows us to prove the main results with a fixed point argument.

2 Change of Variables

In order to prove Theorem 1.2, we first rewrite the system (1.11) in the cylindrical domain \((0,\infty ) \times {\mathcal {F}}\) by constructing an invertible mapping \(X(t,\cdot )\) from the reference configuration \({\mathcal {F}}\) onto \({\mathcal {F}}(\eta (t)).\) More generally, for any \(\eta \in C^{1}(\overline{{\mathcal {S}}})\) satisfying (1.1) and a smallness condition

$$\begin{aligned} \Vert \eta \Vert _{L^\infty ({\mathcal {S}})} \leqslant c_{0} \end{aligned}$$
(2.1)

that ensures in particular (1.2), we can construct a diffeomorphism \(X_\eta : {\mathcal {F}}\rightarrow {\mathcal {F}}(\eta ).\) To do this, we follow the approach of [5]: there exists \(\alpha >0\) such that

$$\begin{aligned} {\mathcal {V}}_{-\alpha } := {\mathcal {S}}\times (-\alpha , 0) \subset {\mathcal {F}}, \quad {\mathcal {V}}_{\alpha } := {\mathcal {S}}\times (0,\alpha ) \subset {\mathbb {R}}^3\setminus {\mathcal {F}}. \end{aligned}$$
(2.2)

Notice that, \(\partial {\mathcal {V}}_{\alpha } \cap \partial {\mathcal {F}}= \Gamma _{S}.\) We consider \(\psi \in C_{c}^{\infty }({\mathbb {R}})\) such that

$$\begin{aligned} \psi = 1 \text{ in } (-\alpha /2,\alpha /2), \quad \psi = 0 \text{ in } {\mathbb {R}} \setminus (-\alpha , \alpha ), \quad 0 \leqslant \psi \leqslant 1. \end{aligned}$$
(2.3)

Let us extend \(\eta \) by 0 in \({\mathbb {R}}^2\setminus {\mathcal {S}}\) so that \(\eta \in C^{1}_c({\mathbb {R}}^2)\) and let us define \(X_\eta \) by

$$\begin{aligned} X_\eta \left( \begin{bmatrix} y_1\\ y_2\\ y_3 \end{bmatrix}\right) =\left( \begin{bmatrix} y_1\\ y_2\\ y_3 +\psi (y_3)\eta (y_1,y_2) \end{bmatrix}\right) \quad \left( y=\begin{bmatrix} y_1\\ y_2\\ y_3 \end{bmatrix}\in {\mathbb {R}}^3\right) . \end{aligned}$$
(2.4)

If we choose \(c_0\) in (2.1) as

$$\begin{aligned} c_{0} := \frac{1}{2\Vert \psi '\Vert _{L^{\infty }({\mathbb {R}})}} \end{aligned}$$
(2.5)

then \(X_\eta \) is a \(C^{1}\)-diffeomorphism from \({\mathcal {F}}\) onto \({\mathcal {F}}(\eta )\) with \(X_\eta (\Gamma _{S}) = \Gamma _{S}(\eta )\). Note that (2.1) and (2.5) yield that \(|\eta |\leqslant {\alpha }/{2}\) in \({\mathcal {S}}\).

Let us assume now that \(\eta \) depends also on time and satisfies for all t relation (2.1) with \(c_0\) given by (2.5). We can define

$$\begin{aligned} X(t,\cdot ) := X_{\eta (t)}. \end{aligned}$$
(2.6)

In particular, \(X(t,\cdot )\) is a \(C^{1}\)-diffeomorphism from \({\mathcal {F}}\) onto \({\mathcal {F}}(\eta (t)).\) For each \(t \geqslant 0,\) we denote by \(Y(t,\cdot ) = X(t,\cdot )^{-1},\) the inverse of \(X(t,\cdot ).\) We have \(X \in C_{b}^0([0,\infty ); C^{1}(\overline{{\mathcal {F}}}))\) and for all \(t \in (0,\infty )\), \(y=[y_1 \ y_2 \ y_3]^{\top } \in {\mathcal {S}}\times (-\alpha /2,\alpha /2),\)

$$\begin{aligned} \det \nabla X (t,y) = 1, \qquad {{\,\mathrm{Cof}\,}}( \nabla X )(t,y) = \begin{bmatrix} 1 &{} 0 &{} - \partial _{y_1} \eta (t,y_{1},y_{2}) \\ 0 &{} 1 &{} - \partial _{y_2} \eta (t,y_{1},y_{2}) \\ 0 &{} 0 &{}1 \end{bmatrix}. \end{aligned}$$
(2.7)

We consider the following change of unknowns

$$\begin{aligned} v(t,y) = {{\,\mathrm{Cof}\,}}\nabla X^{\top }(t,y) \widetilde{v}(t, X(t,y)), \qquad \pi (t,y) = \widetilde{\pi }(t,X(t,y)), \qquad (t,y) \in (0,\infty ) \times {\mathcal {F}}. \end{aligned}$$
(2.8)

The system (1.11) can be rewritten in the form

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _{t} v - \mathrm{div}\,{\mathbb {T}}( v, \pi ) = F( v, \pi , \eta )&{} t> 0, y \in {\mathcal {F}}, \\ \mathrm{div}\,v = 0 &{} t> 0, y \in {\mathcal {F}}, \\ v(t, s, 0) = \partial _{t} \eta (t,s) e_{3} &{} t> 0, s \in {\mathcal {S}}, \\ v = 0 &{} t> 0, y \in \Gamma _{0}, \\ \partial _{tt} \eta + P_{m}\left( \Delta _{s}^{2} \eta \right) - \Delta _{s} \partial _{t}\eta &{}\\ \qquad \qquad = - P_{m} \Big ( {\mathbb {T}}( v, \pi )|_{\Gamma _{S}} e_{3} \cdot e_{3} \Big ) + P_{m}\Big ( {H}( v, \pi , \eta ) \Big ) \qquad &{} t> 0, s \in {\mathcal {S}},\\ \eta = \nabla _s \eta \cdot n_{{\mathcal {S}}} = 0 &{} t > 0, s \in \partial {\mathcal {S}},\\ \eta (0,\cdot ) = \eta _{1}^{0} \text{ in } {\mathcal {S}}, \quad \partial _{t}\eta (0,\cdot ) = \eta _{2}^{0} \text{ in } {\mathcal {S}}, \quad v(0,\cdot ) = v^{0} \text{ in } {\mathcal {F}}, \end{array}\right. } \end{aligned}$$
(2.9)

where

$$\begin{aligned} v^{0}(y) := {{\,\mathrm{Cof}\,}}\nabla X^{\top }(0,y)\widetilde{v}^{0}(X(0,y)) ={{\,\mathrm{Cof}\,}}\nabla X_{\eta _{1}^0}^{\top }(y) \widetilde{v}^{0}(X_{\eta _{1}^{0}}(y)). \end{aligned}$$
(2.10)

Let us write

$$\begin{aligned} a:={{\,\mathrm{Cof}\,}}(\nabla Y)^{\top }, \quad b:={{\,\mathrm{Cof}\,}}(\nabla X)^{\top } \end{aligned}$$
(2.11)

so that

$$\begin{aligned} v(t,y)=b(t,y)\widetilde{v}(t,X(t,y)), \quad \widetilde{v}(t,x)=a(t,x) v(t,Y(t,x)). \end{aligned}$$
(2.12)

After some standard calculation, we find that in (2.9), the expressions of \(F=\left( F_\alpha \right) _{\alpha =1,2,3}\) and H are

$$\begin{aligned} {F}_{\alpha }(v, \pi , \eta )&= \nu \sum _{i,j,k} b_{\alpha i} \frac{\partial ^2 a_{ik}}{\partial x_j^2}(X) v_k +2\nu \sum _{i,j,k,\ell } b_{\alpha i} \frac{\partial a_{ik}}{\partial x_j}(X) \frac{\partial v_k}{\partial y_\ell } \frac{\partial Y_\ell }{\partial x_j}(X)\nonumber \\&\quad +\nu \sum _{j,\ell ,m} \frac{\partial ^2 v_\alpha }{\partial y_\ell \partial y_m} \left( \frac{\partial Y_\ell }{\partial x_j}(X)\frac{\partial Y_m}{\partial x_j}(X) -\delta _{\ell ,j}\delta _{m,j}\right) +\nu \sum _{j,\ell } \frac{\partial v_\alpha }{\partial y_\ell } \frac{\partial ^2 Y_\ell }{\partial x_j^2}(X)\nonumber \\&\quad -\sum _{k,i} \frac{\partial \pi }{\partial y_k} \left( \det (\nabla X) \frac{\partial Y_\alpha }{\partial x_i}(X) \frac{\partial Y_k}{\partial x_i}(X)-\delta _{\alpha ,i}\delta _{k,i}\right) \nonumber \\&-\sum _{i,j,k,m} b_{\alpha i} \frac{\partial a_{ik}}{\partial x_j}(X) a_{j m}(X) v_k v_m -\frac{1}{\det (\nabla X)} \left[ (v\cdot \nabla ) v\right] _\alpha \nonumber \\&\quad -\left[ b(\partial _t a)(X) v\right] _\alpha -\left[ (\nabla v)(\partial _tY)(X)\right] _\alpha , \end{aligned}$$
(2.13)
$$\begin{aligned} {H}(v, \pi , \eta )&=\nu \Bigg \{\sum _{k=1}^3 \left( \sum _{i=1}^2 \partial _{s_i} \eta \left[ \frac{\partial a_{ik}}{\partial x_3}(X) +\frac{\partial a_{3k}}{\partial x_i}(X)\right] -2\frac{\partial a_{3k}}{\partial x_3}(X) \right) v_k\nonumber \\&\quad + \sum _{k=1}^3 \left( \sum _{i=1}^2 \partial _{s_i} \eta \left[ a_{ik}(X)\frac{\partial Y_\ell }{\partial x_3}(X) +a_{3k}(X)\frac{\partial Y_\ell }{\partial x_i}(X)\right] \right. \nonumber \\&\quad \left. -2\left[ a_{3k}(X)\frac{\partial Y_\ell }{\partial x_3}(X) - \delta _{2,k}\delta _{2,\ell }\right] \right) \frac{\partial v_k}{\partial y_\ell }\Bigg \}(t,s,1). \end{aligned}$$
(2.14)

We prove the following result

Lemma 2.1

Let \(1< p, q < \infty \) such that

$$\begin{aligned} \frac{1}{p} + \frac{3}{2q} < \frac{3}{2}, \end{aligned}$$
(2.15)

and \((\eta _{1}^{0}, \widetilde{v}^{0})\) satisfies (1.12). Then \(v^{0}\) defined by (2.10) satisfies \(v^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {F}}).\)

Proof

By using (2.15), we deduce that \(\eta _1^0\in C^1(\overline{{\mathcal {S}}})\). In particular, the map

$$\begin{aligned} \widetilde{v}^{0} \mapsto \widehat{v}^0=\widetilde{v}^{0}\circ X_{\eta _{1}^{0}} \end{aligned}$$
(2.16)

is linear and continuous from \(L^q({\mathcal {F}}(\eta _1^0))\) into \(L^q({\mathcal {F}})\). Let us show that it is also continuous from \(W^{2,q}({\mathcal {F}}(\eta _1^0))\) into \(W^{2,q}({\mathcal {F}})\): Some computation yields

$$\begin{aligned} \frac{\partial ^{2} \widehat{v}^{0}}{\partial y_{i} \partial y_{j}}(y) = \sum _{k, \ell } \frac{\partial ^{2} \widetilde{v}^{0}}{\partial x_{\ell } \partial x_{k}} (X_{\eta _{1}^{0}}(y)) \frac{\partial X_{\eta _{1}^{0},\ell }}{\partial y_{j}}(y) \frac{\partial X_{\eta _{1}^{0},k}}{\partial y_{i}}(y) +\frac{\partial \widetilde{v}^{0}}{\partial x_{k}}(X_{\eta _{1}^{0}}(y)) \frac{\partial ^{2} X_{\eta _{1}^{0},k}}{\partial y_{i} \partial y_{j}} (y). \end{aligned}$$
(2.17)

Using that \(\eta _1^0\in C^1(\overline{{\mathcal {S}}})\), we deduce that the first term in the right-hand side of the above relation belongs to \(L^{q}({\mathcal {F}}).\) For the second term, we first note that \(\displaystyle \frac{\partial \widetilde{v}^{0}}{\partial x_{k}}(X(\cdot )) \in W^{1,q}({\mathcal {F}})\) and \(\displaystyle \frac{\partial ^{2} X_{\eta _{1}^{0},k}}{\partial y_{i} \partial y_{j}} \in B^{2(1-1/p)}_{q,p}({\mathcal {F}}).\) Therefore by [44, Theorem(i), page 196], \(\displaystyle \frac{\partial ^{2} X_{\eta _{1}^{0},k}}{\partial y_{i} \partial y_{j}} \in W^{s_{1},q}({\mathcal {F}})\) for any \(s_{1} < 2(1-1/p).\) Applying standard result on the product of Sobolev spaces we conclude that the second term in (2.17) also belongs to \(L^{q}({\mathcal {F}}).\)

Then by interpolation, we deduce that the map (2.16) is linear continuous from \(B^{2(1-1/p)}_{q,p}({\mathcal {F}}(\eta _1^0))\) into \(B^{2(1-1/p)}_{q,p}({\mathcal {F}})\). Therefore, if \(v^0\in B^{2(1-1/p)}_{q,p}({\mathcal {F}}(\eta _1^0))\), we have

$$\begin{aligned} {{\,\mathrm{Cof}\,}}\nabla X_{\eta _{1}^0}^{\top }\in B^{1+2(1-1/p)}_{q,p} ({\mathcal {F}}), \quad \widehat{v}^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {F}}) \end{aligned}$$

and we deduce that the product \(v^0\in B^{2(1-1/p)}_{q,p}({\mathcal {F}})\) by using [41, Theorem 2, pp.191-192, relation (17)]. \(\square \)

Using the above lemma and the definition of X defined in (2.6), the hypotheses (1.12), (1.13), (1.14) on the initial conditions are transformed into the following conditions:

$$\begin{aligned}&\eta _{1}^{0} \in B^{2(2-1/p)}_{q,p}({\mathcal {S}}),\quad \eta _{2}^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {S}}), \quad v^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {F}}), \end{aligned}$$
(2.18)
$$\begin{aligned}&\eta _{1}^{0} = \nabla _s \eta _{1}^{0}\cdot n_{{\mathcal {S}}} = 0 \quad \text{ on } \ \partial {\mathcal {S}}, \quad \Gamma _{0} \cap \Gamma _{S}(\eta _{1}^{0}) = \emptyset , \quad \int _{{\mathcal {S}}} \eta _{1}^{0} \ ds = 0, \quad \int _{{\mathcal {S}}} \eta _{2}^{0} \ ds = 0, \nonumber \\&\mathrm{div}(v^{0}) = 0 \quad \text {in} \ {\mathcal {F}}, \end{aligned}$$
(2.19)
$$\begin{aligned}&{\left\{ \begin{array}{ll} v^{0}(s,0)\cdot e_3= \eta _{2}^{0}(s) \quad s \in {\mathcal {S}}, \quad v^{0} \cdot {n} = 0 \quad \text{ on } \ \Gamma _{0} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q} > 1, \\ v^0(s, 0) = \eta _2^0(s) e_{3} \quad s \in {\mathcal {S}}, \quad v^{0} = 0 \quad \text{ on } \ \Gamma _{0}, \quad \eta _{2}^{0} = 0 \quad \text{ on } \ \partial {\mathcal {S}} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q}< 1, \\ \nabla _{s} \eta _{2}^{0} \cdot n_{S} = 0 \text{ on } \partial {\mathcal {S}} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q} < \frac{1}{2}. \end{array}\right. } \end{aligned}$$
(2.20)

Here n is the unit normal to \(\partial {\mathcal {F}}\) outward \({\mathcal {F}}\) and in particular on \(\Gamma _S\), \(n=e_3\).

Using the above change of variables Theorem 1.1 and Theorem 1.2 can be rephrased as

Theorem 2.2

Let \(p,q\in (1,\infty )\) satisfying the condition (1.15). Let us assume that \(\eta _{1}^{0} = 0\) and \((\eta _2^0,v^0)\) satisfies (2.18), (2.19), (2.20). Then there exists \(T > 0,\) depending only on \((\eta _2^0,v^0),\) such that the system (2.9) admits a unique strong solution \((v, \pi , \eta )\) in the class of functions satisfying

$$\begin{aligned} v \in W^{1,2}_{p,q}((0,T) ; {\mathcal {F}}), \quad \pi \in L^{p}(0,T;W_{m}^{1,q}({\mathcal {F}})), \quad \eta \in W^{2,4}_{p,q}((0,T) ; {\mathcal {S}}) \end{aligned}$$

Moreover, \(\eta \) satisfies (2.1) and \(X(t,\cdot ) : {\mathcal {F}}\rightarrow {\mathcal {F}}(\eta (t))\) is a \(C^{1}\)-diffeomorphism for all \(t \in [0,T].\)

Theorem 2.3

Let \(p,q\in (1,\infty )\) satisfying the condition (1.15). Then there exists \(\beta _{0} > 0\) such that, for all \(\beta \in [0,\beta _{0}],\) there exist \(\varepsilon _{0}\) and \(C > 0,\) such that for any \((\eta _1^0,\eta _2^0,{v}^0)\) satisfying (2.18), (2.19), (2.20) and

$$\begin{aligned} \Vert \eta _{1}^{0}\Vert _{B^{2(2-1/p)}_{q,p}({\mathcal {S}})} +\Vert \eta _{2}^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {S}})} +\Vert v^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {F}})} < \varepsilon _0, \end{aligned}$$
(2.21)

the system (2.9) admits a unique strong solution \((v, \pi , \eta )\) in the class of functions satisfying

$$\begin{aligned} v \in W^{1,2}_{p,q,\beta }((0,\infty ) ; {\mathcal {F}}), \quad \pi \in L^{p}_\beta (0,\infty ;W_{m}^{1,q}({\mathcal {F}})), \quad \eta \in W^{2,4}_{p,q,\beta }((0,\infty ) ; {\mathcal {S}}) \end{aligned}$$

Moreover, \(\eta \) satisfies (2.1) and \(X(t,\cdot ) : {\mathcal {F}}\rightarrow {\mathcal {F}}(\eta (t))\) is a \(C^{1}\)-diffeomorphism for all \(t \in [0,\infty )\).

3 Some Background on \({\mathcal {R}}\)-sectorial Operators

In this section, we recall some important facts on \({\mathcal {R}}\)-sectorial operators. This notion is associated with the property of \({\mathcal {R}}\)-boundedness (\({\mathcal {R}}\) for Randomized) for a family of operators that we recall here (see, for instance, [10, 11, 30, 45]):

Definition 3.1

Let \({\mathcal {X}}\) and \({\mathcal {Y}}\) be Banach spaces. A family of operators \({\mathcal {E}} \subset {\mathcal {L}}({\mathcal {X}},{\mathcal {Y}})\) is called \({\mathcal {R}}-\)bounded if there exist \(p\in [1,\infty )\) and a constant \(C>0\), such that for any integer \(N \geqslant 1\), any \(T_1, \ldots T_N \in {\mathcal {E}}\), any independent Rademacher random variables \(r_1, \ldots , r_N\), and any \(x_1, \ldots , x_N \in {\mathcal {X}}\),

$$\begin{aligned} \left( {\mathbb {E}} \left\| \sum _{j=1}^{N} r_{j} T_{j} x_{j}\right\| _{\mathcal {Y}}^p\right) ^{1/p} \leqslant C \left( {\mathbb {E}} \left\| \sum _{j=1}^{N} r_{j} x_{j}\right\| _{\mathcal {X}}^p\right) ^{1/p}. \end{aligned}$$

The smallest constant C in the above inequality is called the \({\mathcal {R}}_p\)-bound of \({\mathcal {E}}\) on \({\mathcal {L}}({\mathcal {X}},{\mathcal {Y}})\) and is denoted by \({\mathcal {R}}_p({\mathcal {E}})\).

In the above definition, we denote by \({\mathbb {E}}\) the expectation and a Rademacher random variable is a symmetric random variables with value in \(\{-1,1\}\). It is proved in [11, p.26] that this definition is independent of \(p\in [1,\infty )\).

We have the following useful properties (see Proposition 3.4 in [11]):

$$\begin{aligned} {\mathcal {R}}_p({\mathcal {E}}_1 + {\mathcal {E}}_2) \leqslant {\mathcal {R}}_p({\mathcal {E}}_1 ) + {\mathcal {R}}_p( {\mathcal {E}}_2), \quad {\mathcal {R}}_p({\mathcal {E}}_1 {\mathcal {E}}_2 ) \leqslant {\mathcal {R}}_p({\mathcal {E}}_1 ) {\mathcal {R}}_p( {\mathcal {E}}_2). \end{aligned}$$
(3.1)

For any \(\beta \in (0,\pi )\), we write

$$\begin{aligned} \Sigma _{\beta } = \{ \lambda \in {\mathbb {C}} \setminus \{0\} \ ; \ |\arg (\lambda )| < \beta \}. \end{aligned}$$

We recall the following definition:

Definition 3.2

(sectorial and \({\mathcal {R}}\)-sectorial operators). Let A be a densely defined closed linear operator on a Banach space \({\mathcal {X}}\) with domain \({\mathcal {D}}(A)\). We say that A is a (\({\mathcal {R}}\))-sectorial operator of angle \(\beta \in (0, \pi )\) if

$$\begin{aligned} \Sigma _{\beta } \subset \rho (A) \end{aligned}$$

and if the set

$$\begin{aligned} R_\beta = \left\{ \lambda (\lambda - A)^{-1} \ ; \ \lambda \in \Sigma _{\beta } \right\} \end{aligned}$$

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

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

$$\begin{aligned} \widetilde{R_\beta }= \left\{ A(\lambda - A)^{-1} \ ; \ \lambda \in \Sigma _{\beta } \right\} . \end{aligned}$$

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

This notion of \({\mathcal {R}}\)-sectorial operators is related to the maximal regularity of type \(L^p\) by the following result due to [45] (see also [11, p.45]).

Theorem 3.3

Let \({\mathcal {X}}\) be a UMD Banach space and A a densely defined, closed linear operator on \({\mathcal {X}}\). Then the following assertions are equivalent:

  1. 1.

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

    $$\begin{aligned} u' = A u + f \quad \text {in} \quad (0,T), \quad u(0) = 0 \end{aligned}$$
    (3.2)

    admits a unique solution u with \(u', Au\in L^{p}(0,T;{\mathcal {X}})\) and there exists a constant \(C>0\) such that

    $$\begin{aligned} \Vert u'\Vert _{L^{p}(0,T;{\mathcal {X}})}+\Vert Au\Vert _{L^{p}(0,T;{\mathcal {X}})}\leqslant C\Vert f\Vert _{L^{p}(0,T;{\mathcal {X}})}. \end{aligned}$$
  2. 2.

    A is \({\mathcal {R}}\)-sectorial of angle \(> \frac{\pi }{2}\).

We recall that \({\mathcal {X}}\) is a UMD Banach space if the Hilbert transform is bounded in \(L^p({\mathbb {R}};{\mathcal {X}})\) for \(p\in (1,\infty )\). In particular, the closed subspaces of \(L^q(\Omega )\) for \(q\in (1,\infty )\) are UMD Banach spaces. We refer the reader to [2, pp.141–147] for more information on UMD spaces.

Combining the above theorem with [13, Theorem 2.4] and [43, Theorem 1.8.2], we can deduce the following result on the system

$$\begin{aligned} u' = A u + f \quad \text {in}\quad (0,\infty ), \quad u(0) = u_{0}. \end{aligned}$$
(3.3)

Corollary 3.4

Let \({\mathcal {X}}\) be a UMD Banach space, \(1< p < \infty \) and let A be a closed, densely defined operator in \({\mathcal {X}}\) with domain \({\mathcal {D}}(A).\) Let us assume that A is a \({\mathcal {R}}\)-sectorial operator of angle \( > \frac{\pi }{2}\) and that the semigroup generated by A has negative exponential type. Then for every \(u_{0} \in ({\mathcal {X}}, {\mathcal {D}}(A))_{1-1/p,p}\) and for every \(f \in L^{p}(0,\infty ;{\mathcal {X}}),\) the system (3.3) admits a unique solution in \(L^{p}(0,\infty ;{\mathcal {D}}(A)) \cap W^{1,p}(0,\infty ;{\mathcal {X}}).\)

Let us also mention, the following useful result on the perturbation theory of \({\mathcal {R}}\)-sectoriality, obtained in [29, Corollary 2].

Proposition 3.5

Let A be a \({\mathcal {R}}\)-sectorial operator of angle \(\beta \) on a Banach space \({\mathcal {X}}\). Let \(B : {\mathcal {D}}(B) \rightarrow {\mathcal {X}}\) be a linear operator such that \({\mathcal {D}}(A) \subset {\mathcal {D}}(B)\) and such that there exist \(a,b\geqslant 0\) satisfying

$$\begin{aligned} \Vert B x \Vert _{{\mathcal {X}}} \leqslant a \Vert Ax \Vert _{{\mathcal {X}}} + b \Vert x \Vert _{{\mathcal {X}}} \quad (x\in {\mathcal {D}}(A)). \end{aligned}$$
(3.4)

If

$$\begin{aligned} a < \frac{1}{\widetilde{M}_\beta (A) \widetilde{{\mathcal {R}}_{\beta }}(A)} \quad \text {and} \quad \lambda > \frac{b M_\beta (A) \widetilde{{\mathcal {R}}_{\beta }}(A)}{1-a \widetilde{M}_\beta (A) \widetilde{{\mathcal {R}}_{\beta }}(A)}, \end{aligned}$$

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

4 Linearized System

In order to study the system (2.9), we linearized it and use the theory of the previous section. To this aim, we introduce the operator \({\mathcal {T}} : L^2({\mathcal {S}}) \rightarrow {L}^2(\partial {\mathcal {F}})\) defined by

$$\begin{aligned} ({\mathcal {T}}\eta )(y)&= \left( P_{m} \eta (s)\right) e_3\quad \text{ if } \quad y=(s,0)\in \Gamma _{S},\nonumber \\ ({\mathcal {T}}\eta )(y)&=0\quad \quad \quad \, \text{ if } \quad y\in \Gamma _{0}. \end{aligned}$$
(4.1)

We consider the following linear system

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _{t} v - \mathrm{div}\,{\mathbb {T}}( v, \pi ) = f&{} \text {in} \ (0,\infty )\times {\mathcal {F}}, \\ \mathrm{div}\,v = 0 &{} \text {in} \ (0,\infty )\times {\mathcal {F}}, \\ v = {\mathcal {T}}\eta _2 &{} \text {on} \ (0,\infty )\times \partial {\mathcal {F}}\\ \partial _t \eta _1 = \eta _2 &{} \text {in} \ (0,\infty )\times {\mathcal {S}},\\ \partial _{t} \eta _2 + P_{m}\left( \Delta _{s}^{2} \eta _1\right) -\Delta _{s} \eta _2 =- P_{m} \Big ( {\mathbb {T}}( v, \pi )|_{\Gamma _{S}} e_{3} \cdot e_{3} \Big ) +P_m h &{} \text {in} \ (0,\infty )\times {\mathcal {S}},\\ \eta _1= \nabla _s \eta _1 \cdot n_{{\mathcal {S}}} = 0 &{} \text {on} \ (0,\infty )\times \partial {\mathcal {S}},\\ \eta _1(0,\cdot ) = \eta _{1}^{0} \text{ in } {\mathcal {S}}, \quad \eta _2(0,\cdot ) = \eta _{2}^{0} \text{ in } {\mathcal {S}}, \quad v(0,\cdot ) = v^{0} \text{ in } {\mathcal {F}}. \end{array}\right. } \end{aligned}$$
(4.2)

One can simplify the system (4.2): using that \(\mathrm{div}\,v = 0\) in \({\mathcal {F}}\) and \(v_{1}=v_2 = 0\) on \(\Gamma _{S}\) we deduce that \((Dv)|_{\Gamma _S}e_3\cdot e_3 = 0\). Thus

$$\begin{aligned} - P_{m} \Big ( {\mathbb {T}}( v, \pi )|_{\Gamma _{S}} e_{3} \cdot e_{3} \Big )= \gamma _m \pi , \end{aligned}$$

where \(\gamma _m\) is the following modified trace operator:

$$\begin{aligned} \gamma _{m} f := P_{m} (f|_{\Gamma _{S}}) = f(\cdot ,0) -\frac{1}{|{\mathcal {S}}|}\int _{{\mathcal {S}}} f(s',0) \ ds' \quad (f \in W^{r, q}({\mathcal {F}}) \text{ with } r > 1/q). \end{aligned}$$
(4.3)

This cancelation plays no role in our result and is only used to simplify the calculation.

4.1 The Fluid Operator

Here we recall some results on the Stokes operator in the \(L^q\) framework. Let us introduce the Banach space

$$\begin{aligned} W^q_{\mathrm{div}} ({\mathcal {F}}) = \left\{ \varphi \in L^{q}({\mathcal {F}}) \ ; \ \mathrm{div}\,\varphi \in L^{q}({\mathcal {F}})\right\} , \end{aligned}$$

equipped with the norm

$$\begin{aligned} \Vert \varphi \Vert _{W^q_{\mathrm{div}} ({\mathcal {F}})} :=\Vert \varphi \Vert _{L^{q}({\mathcal {F}})} + \Vert \mathrm{div}\,\varphi \Vert _{L^{q}({\mathcal {F}})}. \end{aligned}$$

We recall (see, for instance, [17, Lemma 1]) that the normal trace can be extended as a continuous and surjective map

$$\begin{aligned} \gamma _{n} : W^q_{\mathrm{div}} ({\mathcal {F}})&\rightarrow W^{-1/q,q}(\partial {\mathcal {F}}),\\ \varphi&\mapsto \varphi \cdot n. \end{aligned}$$

In particular, we can define

$$\begin{aligned} L^{q}_{\sigma }({\mathcal {F}}) = \left\{ \varphi \in {L}^q({\mathcal {F}}) \ ; \ \mathrm{div}\,\varphi = 0 \quad \text {in} \ {\mathcal {F}}, \ \varphi \cdot {n} = 0 \text{ on } \partial {\mathcal {F}}\right\} . \end{aligned}$$

We have the following Helmholtz-Weyl decomposition (see, for instance Section 3 and Theorem 2 of [17]):

$$\begin{aligned} {L}^q({\mathcal {F}}) = L^{q}_{\sigma }({\mathcal {F}}) \oplus G^{q}({\mathcal {F}}), \quad \text {where} \quad G^{q}({\mathcal {F}}) = \left\{ \nabla \varphi \ ; \ \varphi \in W^{1,q}({\mathcal {F}})\right\} . \end{aligned}$$

The corresponding projection operator \({\mathcal {P}}\) from \({ L}^q({\mathcal {F}})\) onto \(L^{q}_{\sigma }({\mathcal {F}})\) can be obtained as

$$\begin{aligned} {\mathcal {P}} f = f - \nabla \varphi , \end{aligned}$$
(4.4)

where \(\varphi \in W^{1,q}({\mathcal {F}})\) is a solution of the following Neumann problem

$$\begin{aligned} \Delta \varphi = \mathrm{div}\,f \quad \text{ in } \ {\mathcal {F}}, \quad \frac{\partial \varphi }{\partial n} = f \cdot n \quad \text{ on } \ \partial {\mathcal {F}}, \end{aligned}$$
(4.5)

that is a solution of

$$\begin{aligned} \int _{{\mathcal {F}}} \nabla \varphi \cdot \nabla \psi \ dy=\int _{{\mathcal {F}}} f \cdot \nabla \psi \ dy \quad (\psi \in W^{1,q'}({\mathcal {F}})), \end{aligned}$$

where \(q'\) is the conjugate exponent of q.

Let us denote by \(A_{F} = {\mathcal {P}} \Delta ,\) the Stokes operator in \(L^{q}_{\sigma }({\mathcal {F}})\) with domain

$$\begin{aligned} {\mathcal {D}}(A_{F}) = W^{2,q}({\mathcal {F}}) \cap W^{1,q}_{0}({\mathcal {F}})\cap L^{q}_{\sigma }({\mathcal {F}}). \end{aligned}$$

Theorem 4.1

Assume \(1< q< \infty .\) Then the Stokes operator \(A_{F}\) generates a \(C^{0}\)-semigroup of negative type. Moreover \(A_{F}\) is an \({\mathcal {R}}\)-sectorial operator in \(L^{q}_{\sigma }({\mathcal {F}})\) of angle \(\beta \) for any \(\beta \in (0,\pi )\).

For the proof, we refer to Corollary 1.2 and Theorem 1.4 in [19].

4.2 The Structure Operator

Let us set

$$\begin{aligned} {\mathcal {X}}_{S} = W^{2,q}_{0,m}({\mathcal {S}}) \times L^{q}_{m}({\mathcal {S}}) \end{aligned}$$

and let us consider the operator \(A_{S} : {\mathcal {D}}(A_{S}) \rightarrow {\mathcal {X}}_{S}\) defined by

$$\begin{aligned} {\mathcal {D}}(A_{S}) = \left( W^{4,q}({\mathcal {S}}) \cap W^{2,q}_{0,m}({\mathcal {S}})\right) \times W^{2,q}_{0,m}({\mathcal {S}}), \quad A_{S} = \begin{pmatrix} 0 &{} \mathrm{Id}\\ -P_{m} \Delta ^{2} &{} \Delta \end{pmatrix}, \end{aligned}$$

where \(P_m\) is defined by (1.8).

Theorem 4.2

Let us assume that \(1< q< \infty .\) Then there exists \(\gamma _{1} > 0\) such that \(A_{S} - \gamma _{1}\) is an \({\mathcal {R}}\)-sectorial operator on \({\mathcal {X}}_{S}\) of angle \(\beta _{1} > \pi /2.\)

Proof

We first consider

$$\begin{aligned} {\mathcal {X}}_{S}^{0} := W^{2,q}_{0}({\mathcal {S}}) \times L^{q}({\mathcal {S}}) \end{aligned}$$

and the operator \(A_{S}^{0}\) defined by

$$\begin{aligned} {\mathcal {D}}(A_{S}^{0}) = \left( W^{4,q}({\mathcal {S}}) \cap W^{2,q}_{0}({\mathcal {S}}) \right) \times W^{2,q}_{0}({\mathcal {S}}), \quad A_{S}^{0} = \begin{pmatrix} 0 &{} \mathrm{Id}\\ -\Delta ^{2} &{} \Delta \end{pmatrix}. \end{aligned}$$

Applying Theorem 5.1 in [12], we have that \(A_{S}^0\) is \({\mathcal {R}}\)-sectorial in \({\mathcal {X}}_{S}^{0}\) of angle \(\beta _{0} > \pi /2.\)

Now we can extend \(A_S\) on \({\mathcal {D}}(A_{S}^{0})\) by \(\widetilde{A}_S = A_{S}^{0}+B_{S}\) where

$$\begin{aligned} B_{S} = \begin{pmatrix} 0 &{} 0 \\ (\mathrm{Id}- P_{m}) \Delta ^{2} &{} 0 \end{pmatrix}, \quad (\mathrm{Id}- P_{m}) \Delta ^{2} \eta _1 =\frac{1}{|{\mathcal {S}}|} \int _{\partial {\mathcal {S}}} \left( \nabla \Delta \eta _1\right) \cdot n_{{\mathcal {S}}} \ ds. \end{aligned}$$

Using standard result on the trace operator, we see that \(B_S\) satisfies the hypotheses of Proposition 3.5 and in particular for any \(a>0\) there exists \(b>0\) such that (3.4) holds. Therefore, there exists \(\gamma _{1} > 0\) such that \(\widetilde{A}_{S} - \gamma _{1}\) is an \({\mathcal {R}}\)-sectorial operator on \({\mathcal {X}}_{S}^{0}\) of angle \(\beta _0.\)

Let \(\lambda \ne 0\), \((g_{1}, g_{2}) \in {\mathcal {X}}_{S}\) and \((\eta _1,\eta _2)\in {\mathcal {D}}(A_{S}^{0})\) such that

$$\begin{aligned} (\lambda -\widetilde{A}_S)\begin{bmatrix} \eta _1 \\ \eta _2 \end{bmatrix} =\begin{bmatrix} g_1 \\ g_2 \end{bmatrix}. \end{aligned}$$

We can write this equation as

$$\begin{aligned}&\lambda \eta _{1} - \eta _{2} = g_{1} \qquad \text{ in } {\mathcal {S}}, \\&\lambda \eta _{2} + P_{m} \Delta ^{2} \eta _{1} - \Delta \eta _{2} =g_{2}, \quad \text{ in } {\mathcal {S}}, \\&\eta _1 = \nabla _s \eta _1 \cdot n_{{\mathcal {S}}} = \eta _2 = \nabla _s \eta _2 \cdot n_{{\mathcal {S}}}= 0 \qquad \text {on} \ \partial {\mathcal {S}}. \end{aligned}$$

Integrating the first two equations over \({\mathcal {S}}\) we find that \((\eta _1,\eta _2)\in {\mathcal {D}}(A_{S})\). Thus

$$\begin{aligned} \left[ (\lambda -\widetilde{A}_S)^{-1}\right] _{|{\mathcal {X}}_S} =(\lambda -A_S)^{-1}. \end{aligned}$$

Using basic properties on \({\mathcal {R}}\)-boundedness, we deduce the result. \(\square \)

4.3 The Fluid-Structure Operator

In this subsection we rewrite (4.2) in a suitable operator form. The idea is to eliminate the pressure from both the fluid and the structure equations. To eliminate the pressure from the fluid equation we use the Leray projector \({\mathcal {P}}\) defined in Eq. (4.4). Following [39], we first decompose the fluid velocity into two parts \({\mathcal {P}}v\) and \((\mathrm{Id}-{\mathcal {P}})v.\) Next, we split the pressure into two parts, one which depends on \({\mathcal {P}} v\) and another part which depends on \(\eta _{2}.\) This will lead us to an equation of evolution for \(({\mathcal {P}} v, \eta _{1}, \eta _{2})\) and an algebraic equation for \((\mathrm{Id}- {\mathcal {P}})v.\)

The advantage of this formulation is that the \({\mathcal {R}}\)-boundedness of the fluid-structure operator can be obtained just by using the fact that the operators \(A_{F}\) and \(A_{S}\) are \({\mathcal {R}}\)-sectorial and a perturbation argument. This idea has been used in several fluid-solid interaction problems in the Hilbert space setting as well as in \(L^{q}\)-setting (see, for instance, [27, 34, 36, 40] and the references therein).

Let us consider the following problem :

$$\begin{aligned} \left\{ \begin{array}{rl} - \mathrm{div}\,{\mathbb {T}}(w, \psi ) = f &{}\quad \text{ in } \ {\mathcal {F}}, \\ \mathrm{div}\,w = 0 &{}\quad \text{ in } \ {\mathcal {F}}, \\ w = {\mathcal {T}} g&{}\quad \text{ on } \ \partial {\mathcal {F}}, \\ \displaystyle \int _{{\mathcal {F}}} \psi \ dx= 0. \end{array}\right. \end{aligned}$$
(4.6)

From [42, Proposition 2.3, p. 35], we have the following result:

Lemma 4.3

Assume \(1< q < \infty \). For any \(f \in L^{q}({\mathcal {F}})\) and \(g \in W^{2,q}_{0,m}({\mathcal {S}})\), the system (4.6) admits a unique solution \((w,\psi )\in W^{2,q}({\mathcal {F}}) \times W_{m}^{1,q}({\mathcal {F}}).\)

This allows us to introduce the following operators: we consider

$$\begin{aligned} D_{\text {v}} \in {\mathcal {L}}(W^{2,q}_{0,m}({\mathcal {S}}),W^{2,q}({\mathcal {F}})) \quad \text {and} \quad D_{\text {p}} \in {\mathcal {L}}(W^{2,q}_{0,m}({\mathcal {S}}),W^{1,q}_{m}({\mathcal {F}})) \end{aligned}$$
(4.7)

defined by

$$\begin{aligned} D_{\text {v}}g = w, \quad D_{\text {p}}g = \psi , \end{aligned}$$
(4.8)

where \((w,\psi )\) is the solution to the problem (4.6) associated with g and in the case \(f = 0.\)

Second, we consider the Neumann problem

$$\begin{aligned} \Delta \varphi = 0 \text{ in } {\mathcal {F}}, \quad \frac{\partial \varphi }{\partial n} = h \text{ on } \partial {\mathcal {F}}, \quad \int _{{\mathcal {F}}} \varphi \ dx= 0. \end{aligned}$$
(4.9)

Let us denote by N the operator defined by

$$\begin{aligned} N h = \varphi . \end{aligned}$$
(4.10)

Using classical results (see for instance Theorem 4.2 and Theorem 4.3 of [33]), we have the following properties of N:

$$\begin{aligned}&N \in {\mathcal {L}}(W^{1-1/q,q}_{m}(\partial {\mathcal {F}}), W^{2,q}_{m}({\mathcal {F}})), \quad N \in {\mathcal {L}}(W^{-1/q,q}_{m}(\partial {\mathcal {F}}), W^{1,q}_{m}({\mathcal {F}})), \nonumber \\&N \in {\mathcal {L}}(L^{q}_{m}(\partial {\mathcal {F}}), W^{1+1/q -\varepsilon ,q}_{m}({\mathcal {F}})), \end{aligned}$$
(4.11)

for any \(\varepsilon > 0.\) We recall that \(W^{-1/q,q}_{m}(\partial {\mathcal {F}})\) is defined as follows:

$$\begin{aligned} W^{-1/q,q}_{m}(\partial {\mathcal {F}})=\left\{ h \in W^{-1/q,q}(\partial {\mathcal {F}}) \ ; \ \left\langle h, 1 \right\rangle _{W^{-1/q,q},W^{1-1/q',q'}} = 0\right\} , \end{aligned}$$
(4.12)

where \(q'\) the conjugate exponent of q.

We also define

$$\begin{aligned} N_S g = N h\quad \text {with}\quad h(y)={\left\{ \begin{array}{ll} g(s) &{}\text{ if } \quad y=(s,0)\in \Gamma _{S},\\ 0 &{} \text{ if } \quad y\in \Gamma _{0}. \end{array}\right. } \end{aligned}$$
(4.13)

From the above properties of N, we deduce that

$$\begin{aligned} N_{S}\in {\mathcal {L}}(L^{q}_m({\mathcal {S}}), W^{1+1/q - \varepsilon ,q}_{m}({\mathcal {F}})), \end{aligned}$$
(4.14)

for any \(\varepsilon > 0.\)

Finally, we introduce the operator \(N_{HW} \in {\mathcal {L}}(L^{q}({\mathcal {F}}), W^{1,q}_{m}({\mathcal {F}}))\) defined by

$$\begin{aligned} N_{HW} f = \varphi , \end{aligned}$$
(4.15)

where \(\varphi \) solves (4.5).

Using the above operators, we can obtain the following proposition. The proof is similar to the proof of [36, Proposition 3.7]. For the sake of completeness, we provide a short proof here.

Proposition 4.4

Let \(1<p,q<\infty .\) Assume

$$\begin{aligned}&v \in W^{1,2}_{p,q}((0,\infty ) ; {\mathcal {F}}), \quad \pi \in L^{p}(0,\infty ;W^{1,q}_{m}({\mathcal {F}})), \\&\eta _{1} \in W^{2,4}_{p,q}((0,\infty ) ; {\mathcal {S}}), \quad \eta _{2} \in W^{1,2}_{p,q}((0,\infty ) ; {\mathcal {S}}). \end{aligned}$$

Then \((v,\pi , \eta _{1}, \eta _{2})\) is a solution of (4.2) if and only if

$$\begin{aligned} {\left\{ \begin{array}{ll} {\mathcal {P}} v' = A_{F} {\mathcal {P}} v - A_{F} {\mathcal {P}} D_{\mathrm{v}} \eta _{2} +{\mathcal {P}} f&{} \text {in} \ (0,\infty ),\\ \partial _t \eta _1 = \eta _2 &{} \text {in} \ (0,\infty ),\\ (\mathrm{Id}+ \gamma _{m}N_{S})\partial _t \eta _2 + P_{m}\Delta ^2 \eta _1 -\Delta \eta _2 = \gamma _{m} N (\nu \Delta {\mathcal {P}} v \cdot n) +P_m h +\gamma _m N_{HW} f &{} \text {in} \ (0,\infty ),\\ {[}{\mathcal {P}}v, \eta _{1}, \eta _{2}]^{\top }(0,\cdot ) =[{\mathcal {P}}v^{0},\eta _{1}^{0},\eta _{2}^{0}]^{\top } &{} \\ (\mathrm{Id}- {\mathcal {P}} )v = (\mathrm{Id}- {\mathcal {P}} )D_{\mathrm{v}} \eta _{2} &{} \text {in} \ (0,\infty ),\\ \pi = N (\nu \Delta {\mathcal {P}} v \cdot n) - N_{S} \partial _{t} \eta _{2}+N_{HW} f&{} \text {in} \ (0,\infty ).\\ \end{array}\right. } \end{aligned}$$
(4.16)

Proof

Considering the equation satisfied by \((v - D_{\mathrm{v}} g,\pi - D_{\mathrm{p}} g)\), we obtain (4.16)\(_{1}\) and (4.16)\(_{5}.\) Using (4.4) and (4.5), it follows that \(\Delta (\mathrm{Id}- {\mathcal {P}} ) v = 0\) in \({\mathcal {F}}.\) Thus applying the divergence and normal trace operators to (4.6), we infer that

$$\begin{aligned} \Delta \psi = \mathrm {div} f \quad \text{ in } {\mathcal {F}}, \quad \quad \frac{\partial \psi }{\partial n} = f \cdot n + \nu \Delta {\mathcal {P}} v \cdot n -{\mathcal {T}}\partial _t \eta _2 \cdot n \text{ on } \partial {\mathcal {F}}. \end{aligned}$$
(4.17)

Note that \(\mathrm{div}\,\Delta {\mathcal {P}} v = 0\) and therefore \(\Delta {\mathcal {P}} v \cdot n\) belongs to \(W^{-1/q,q}_{m}(\partial {\mathcal {F}}).\) The expression of \(\psi \) then follows from the definition of the operators N, \(N_S\) and \(N_{HW}\) defined in (4.10), (4.13) and (4.15) respectively. Finally, using the expression of the pressure \(\pi \) we can rewrite the equation satisfied by \(\eta _{2}\) as in (4.16)\(_{3}.\) \(\square \)

In the literature, the operator

$$\begin{aligned} M_S:=\mathrm{Id}+ \gamma _{m}N_{S} \end{aligned}$$

is known as the added mass operator. We are going to show that it is invertible.

Lemma 4.5

The operator \(M_{S} = \mathrm{Id}+ \gamma _{m} N_{S} \in {\mathcal {L}}(L^{q}_{m}({\mathcal {S}}))\) is an automorphism in \(W^{s,q}_{m}({\mathcal {S}})\) for any \(s \in [0,1).\) Moreover, \(M_{S}^{-1} - \mathrm{Id}\in {\mathcal {L}}(L^{q}_{m}({\mathcal {S}}), W^{s,q}_{m}({\mathcal {S}})),\) for any \(s \in [0,1).\) In particular, \(M_{S}^{-1} - \mathrm{Id}\) is a compact operator on \(L^{q}_{m}({\mathcal {S}}).\)

Proof

At first, we show that \(M_{S}\) is an invertible operator on \(L^{q}_{m}({\mathcal {S}}).\) Since

$$\begin{aligned} \gamma _{m}N_{S} \in {\mathcal {L}}(L^{q}_{m}({\mathcal {S}}), W^{1-\varepsilon ,q}_{m}({\mathcal {S}})), \end{aligned}$$

for any \(\varepsilon \in (0,1]\), it is sufficient to show that the kernel of \(M_S\) is reduced to \(\{0\}\): assume

$$\begin{aligned} (\mathrm{Id}+ \gamma _{m}N_{S}) f = 0. \end{aligned}$$
(4.18)

Then \(f\in W^{1-\varepsilon ,q}_{m}({\mathcal {S}})\subset L^{2}_{m}({\mathcal {S}})\) for \(\varepsilon \) small enough. In particular (see (4.13)), \(\vartheta = N_{S} f \in H^1({\mathcal {F}})\) is the weak solution of

$$\begin{aligned} \Delta \vartheta = 0 \text{ in } {\mathcal {F}}, \quad \frac{\partial \vartheta }{\partial n} = f \text{ on } \Gamma _{S}, \quad \frac{\partial \vartheta }{\partial n} = 0 \text{ on } \Gamma _{0}. \end{aligned}$$

Multiplying (4.18) by f and using the above system, we deduce after integration by parts,

$$\begin{aligned} \int _{{\mathcal {S}}} \left[ (\mathrm{Id}+ \gamma _{m}N_{S}) f\right] f\ ds =\int _{{\mathcal {S}}} f^{2} \ ds+ \int _{{\mathcal {F}}} |\nabla \vartheta |^{2} \ dy = 0. \end{aligned}$$

Thus \(f = 0\) and \(M_{S}\) is an invertible operator on \(L^{q}_{m}({\mathcal {S}}).\) Let \(s\in [0,1)\) and \(f_{0} \in W^{s,q}_{m}({\mathcal {S}}).\) By the above argument, there exists a unique \(f \in L^{q}_{m}({\mathcal {S}})\) such that

$$\begin{aligned} (\mathrm{Id}+ \gamma _{m}N_{S}) f = f_{0}. \end{aligned}$$

As \(\gamma _{m}N_{S} f\in W^{s.q}_{m}({\mathcal {S}})\) we conclude that \(f \in W^{s,q}_{m}({\mathcal {S}}).\) Thus \(M_{S}\) is an invertible operator on \(W^{s,q}_{m}({\mathcal {S}}).\) Finally, the compactness of the operator \(M_{S}^{-1} - \mathrm{Id}\) follows from the following identity

$$\begin{aligned} M_{S}^{-1} - \mathrm{Id}= M_{S}^{-1} - M_{S}^{-1} M_{S} = - M_{S}^{-1} \gamma _{m} N_{S}. \end{aligned}$$

\(\square \)

We are now in a position to rewrite the system (4.2) in a suitable operator form. Let us set

$$\begin{aligned} {\mathcal {X}} = L^{q}_\sigma ({\mathcal {F}}) \times {\mathcal {X}}_{S} \end{aligned}$$
(4.19)

and consider the operator \({\mathcal {A}}_{FS} : {\mathcal {D}}({\mathcal {A}}_{FS}) \rightarrow {\mathcal {X}}\) defined by

$$\begin{aligned} {\mathcal {D}}({\mathcal {A}}_{FS}) = \Big \{ [v, \eta _{1}, \eta _{2}]^{\top } \in \left[ W^{2,q}({\mathcal {F}})\cap L^{q}_\sigma ({\mathcal {F}})\right] \times {\mathcal {D}}(A_{S}) \ ; \ v-{\mathcal {P}}D_{\mathrm{v}}\eta _{2} \in {\mathcal {D}}({\mathcal {A}}_{F})\Big \}, \end{aligned}$$

and

$$\begin{aligned} {\mathcal {A}}_{FS} = {\mathcal {A}}_{FS}^{0} + {\mathcal {B}}_{FS}, \end{aligned}$$

with

$$\begin{aligned} {\mathcal {A}}_{FS}^{0}:=\begin{bmatrix} A_F &{} 0 &{} -A_F {\mathcal {P}}D_{\text {v}} \\ 0 &{} 0 &{} \mathrm{Id}\\ 0 &{} - P_{m} \Delta ^2 &{} \Delta \end{bmatrix} \end{aligned}$$
(4.20)

and

$$\begin{aligned} {\mathcal {B}}_{FS} = \begin{bmatrix} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0\\ M_S^{-1} \gamma _m N (\nu \Delta (\cdot )\cdot n) &{} -(M_S^{-1} - \mathrm{Id})P_{m} \Delta ^2 &{} (M_S^{-1}-\mathrm{Id}) \Delta \end{bmatrix}. \end{aligned}$$
(4.21)

Combining Proposition 4.4 and Lemma 4.5, we can rewrite the system (4.2) as

$$\begin{aligned}&\frac{d}{dt} \begin{bmatrix} {\mathcal {P}} v \\ \eta _{1} \\ \eta _{2} \end{bmatrix} = {\mathcal {A}}_{FS} \begin{bmatrix} {\mathcal {P}} v \\ \eta _{1} \\ \eta _{2} \end{bmatrix} + \begin{bmatrix} {\mathcal {P}} f \\ 0 \\ \overline{h} \end{bmatrix}, \quad \begin{bmatrix} {\mathcal {P}} v \\ \eta _{1} \\ \eta _{2} \end{bmatrix} (0) = \begin{bmatrix} {\mathcal {P}} v^{0} \\ \eta _{1}^{0} \\ \eta _{2}^{0} \end{bmatrix}, \end{aligned}$$
(4.22)
$$\begin{aligned}&(\mathrm{Id}- {\mathcal {P}} )v =(\mathrm{Id}- {\mathcal {P}} )D_{\mathrm{v}} \eta _{2}, \end{aligned}$$
(4.23)
$$\begin{aligned}&\pi = N (\nu \Delta {\mathcal {P}} v \cdot n) - N_{S} \partial _{t} \eta _{2} + N_{HW} f, \end{aligned}$$
(4.24)

where

$$\begin{aligned} \overline{h} = M_{S}^{-1} P_{m} h + M_{S}^{-1} \gamma _{m} N_{HW} f. \end{aligned}$$
(4.25)

4.4 \({\mathcal {R}}\)-Sectoriality of the Operator \({\mathcal {A}}_{FS}\)

In this subsection we prove the following theorem

Theorem 4.6

Let \(1< q < \infty .\) There exists \(\gamma _{2} > 0\) such that \({\mathcal {A}}_{FS} - \gamma _{2}\) is an \({\mathcal {R}}\)-sectorial operator in \({\mathcal {X}}\) of angle \(>\pi /2.\) Moreover the operator \({\mathcal {A}}_{FS}\) generates an exponentially stable semigroup on \({\mathcal {X}}:\) there exist constants \(C> 0\) and \(\beta _{0} > 0\) such that

$$\begin{aligned} \left\| e^{t{\mathcal {A}}_{FS}} (v^{0}, \eta _{1}^{0}, \eta _{2}^{0})^{\top } \right\| _{{\mathcal {X}}} \leqslant C e^{-\beta _{0}t} \left\| (v^{0}, \eta _{1}^{0}, \eta _{2}^{0})^{\top } \right\| _{{\mathcal {X}}} \quad (t\geqslant 0). \end{aligned}$$
(4.26)

Proof

Observe that

$$\begin{aligned} \lambda \left( \lambda - {\mathcal {A}}_{FS}^{0}\right) ^{-1} =\begin{bmatrix} \lambda (\lambda - A_{F})^{-1} &{} -A_{F}(\lambda - A_{F})^{-1} {\mathcal {P}}\widetilde{D}_{\mathrm{v}} \lambda (\lambda - A_{S})^{-1} \\ 0 &{} \lambda (\lambda - A_{S})^{-1} \end{bmatrix}, \end{aligned}$$

where \(\widetilde{D}_{\mathrm{v}} \left[ f_{1}, f_{2} \right] ^{\top } =D_{\mathrm{v}} {f}_{2}.\) Using a standard transposition method and Lemma 4.3, we see that

$$\begin{aligned} D_{\mathrm{v}} \in {\mathcal {L}}(L^q_m({\mathcal {S}}),L^q({\mathcal {F}})). \end{aligned}$$
(4.27)

Therefore by Theorems 4.1 and 4.2, there exists \(\gamma > 0\) such that \({\mathcal {A}}_{FS}^{0}-\gamma \) is \({\mathcal {R}}\)-sectorial operator in \({\mathcal {X}}\) of angle \(> \pi /2.\)

Next, we want to show \({\mathcal {B}}_{FS} \in {\mathcal {L}}({\mathcal {D}}({\mathcal {A}}_{FS}), {\mathcal {X}})\) is a compact operator. Assume \([v,\eta _1,\eta _2]^{\top }\in {\mathcal {D}}({\mathcal {A}}_{FS})\). Then \(\Delta v\in L^{q}({\mathcal {F}})\) and \(\mathrm{div}\,\Delta v= 0\) and thus from the trace result recalled in Sect. 4.1,

$$\begin{aligned} (\Delta v) \cdot n\in W^{-1/q,q}_m(\partial {\mathcal {F}}). \end{aligned}$$

This yields \(N((\Delta v) \cdot n) \in W^{1,q}_{m}({\mathcal {F}})\), \(\gamma _m N((\Delta v) \cdot n) \in W_m^{1-1/q,q}({\mathcal {S}})\) and, using Lemma 4.5,

$$\begin{aligned} M_{S}^{-1}\gamma _{m}N((\Delta v)\cdot n) \in W^{1-1/q,q}_m({\mathcal {S}}). \end{aligned}$$

On the other hand, using again Lemma 4.5, we deduce

$$\begin{aligned} (M_S^{-1} - \mathrm{Id})P_{m} \Delta ^2 \in {\mathcal {L}}(W^{4,q}({\mathcal {S}}), W^{1-\varepsilon ,q}_{m}({\mathcal {S}})), \quad (M_S^{-1}-\mathrm{Id}) \Delta \in {\mathcal {L}}(W^{2,q}_{m}({\mathcal {S}}), W^{1-\varepsilon ,q}_{m}({\mathcal {S}})) \end{aligned}$$

for any \(\varepsilon >0\). Therefore, \({\mathcal {B}}_{FS} \in {\mathcal {L}}({\mathcal {D}}({\mathcal {A}}_{FS}), {\mathcal {X}})\) is a compact operator and by [14, Chapter III, Lemma 2.16], \({\mathcal {B}}_{FS}\) is a \({\mathcal {A}}_{FS}^0\)-bounded operator with relative bound 0. Finally, using Proposition 3.5 we conclude the first part of the theorem. In particular \({\mathcal {A}}_{FS}\) generates an analytic semigroup and to show the second part of the theorem, it is sufficient to show that

$$\begin{aligned} {\mathbb {C}}^{+}=\displaystyle \left\{ \lambda \in {\mathbb {C}} \ ; \ \mathrm {Re} \lambda \geqslant 0\right\} \subset \rho ({\mathcal {A}}_{FS}). \end{aligned}$$

Moreover, using that \({\mathcal {A}}_{FS}\) has a compact resolvent and the Fredholm alternative theorem, we can show the above relation by proving that \(\ker (\lambda - {\mathcal {A}}_{FS})=\{0\}\) for \(\lambda \in {\mathbb {C}}^{+}\). Assume \(\lambda \in {\mathbb {C}}^{+}\) and

$$\begin{aligned} (v, \pi , \eta _{1}, \eta _{2}) \in W^{2,q}({\mathcal {F}}) \times W^{1,q}_{m}({\mathcal {F}}) \times W^{4,q}_m({\mathcal {S}}) \times W^{2,q}_{m}({\mathcal {S}}) \end{aligned}$$

satisfy

$$\begin{aligned} {\left\{ \begin{array}{ll} \lambda v - \mathrm{div}\,{\mathbb {T}}(v, \pi ) = 0 &{} \text {in} \ {\mathcal {F}},\\ \mathrm{div}\,v = 0&{} \text {in} \ {\mathcal {F}},\\ v = {\mathcal {T}}\eta _{2}&{} \text{ on } \ \partial {\mathcal {F}},\\ \lambda \eta _{1} - \eta _{2} = 0 &{} \text{ in }\ {\mathcal {S}}, \\ \lambda \eta _{2} + P_{m} \Delta ^{2} \eta _{1} - \Delta \eta _{2} = \gamma _{m} \pi &{} \text{ in }\ {\mathcal {S}}, \\ \eta _1 = \nabla _s \eta _1\cdot n_{{\mathcal {S}}} = 0 &{} \text{ on } \ \partial {\mathcal {S}}. \end{array}\right. } \end{aligned}$$
(4.28)

First we notice that

$$\begin{aligned} (v, \pi , \eta _{1}, \eta _{2}) \in W^{2,2}({\mathcal {F}}) \times W^{1,2}_{m}({\mathcal {F}}) \times W^{4,2}_m({\mathcal {S}}) \times W^{2,2}_{m}({\mathcal {S}}). \end{aligned}$$
(4.29)

If \(q \geqslant 2\) then it is a consequence of Hölder’s inequality. Let us assume that \(1< q< 2\) and let us take \(\lambda _0 \in \rho ({\mathcal {A}}_{FS})\) (see Theorem 4.6). We have

$$\begin{aligned} (\lambda _0 - {\mathcal {A}}_{FS})[v, \eta _{1}, \eta _{2}]^{\top } =(\lambda _0-\lambda )[v, \eta _{1}, \eta _{2}]^{\top } \end{aligned}$$

By following the calculation done in Sect. 4.3, we see that the system (4.28) can be written as

$$\begin{aligned} {\left\{ \begin{array}{ll} \left( \lambda _0 - {\mathcal {A}}_{FS} \right) \begin{bmatrix} {\mathcal {P}} v \\ \eta _{1} \\ \eta _{2} \end{bmatrix} = (\lambda _0 - \lambda ) \begin{bmatrix} {\mathcal {P}} v \\ \eta _{1} \\ \eta _{2} \end{bmatrix}, \\ (\mathrm{Id}- {\mathcal {P}} )v = (\mathrm{Id}- {\mathcal {P}} )D_{\mathrm{v}} \eta _{2}, \\ \pi = N (\nu \Delta {\mathcal {P}} v \cdot n) - \lambda N_{S} \eta _{2}. \end{array}\right. } \end{aligned}$$

Since \(W^{2,q}({\mathcal {F}}) \subset L^{2}({\mathcal {F}}),\) \(W^{2,q}({\mathcal {S}}) \subset L^{2}({\mathcal {S}})\) and \((\lambda _0 - {\mathcal {A}}_{FS})\) is invertible, we deduce (4.29).

Using (4.29), we can multiply (4.28)\(_{1}\) by \(\overline{v}\) and (4.28)\(_{5}\) by \(\overline{\eta _{2}},\) and we obtain after integration by parts:

$$\begin{aligned} \lambda \int _{{\mathcal {F}}} |v|^{2} \ dy + 2\nu \int _{{\mathcal {F}}} |D(v)|^2 \ dy + \lambda \int _{{\mathcal {S}}} |\eta _2|^2 \ ds +\overline{\lambda }\int _{{\mathcal {S}}} |\Delta _s \eta _{1}|^{2} \ ds+ \int _{{\mathcal {S}}} |\nabla _s \eta _{2}|^{2} \ ds = 0. \end{aligned}$$

Since \(\mathrm {Re} \lambda \geqslant 0,\) from the above equality and using the boundary conditions we obtain that \(v = \pi = \eta _{1} =\eta _{2} = 0.\) This completes the proof of the theorem. \(\square \)

In order to obtain a result of well-posedness on the system (4.2), we need to impose some compatibility conditions on the data:

$$\begin{aligned} \eta _{1}^{0} = \nabla _s \eta _{1}^{0} \cdot n_{{\mathcal {S}}}= 0 \quad \text{ on } \ \partial {\mathcal {S}}, \quad \int _{{\mathcal {S}}} \eta _{1}^{0} \ ds = 0, \quad \int _{{\mathcal {S}}} \eta _{2}^{0} \ ds = 0, \quad \mathrm{div}\,v^{0} = 0 \text{ in } {\mathcal {F}}, \end{aligned}$$
(4.30)

and

$$\begin{aligned} {\left\{ \begin{array}{ll} v^{0}\cdot n = {\mathcal {T}}\eta _2\cdot n\quad \text{ on } \ \partial {\mathcal {F}}&{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q} > 1, \\ v^{0}= {\mathcal {T}}\eta _2\quad \text{ on } \ \partial {\mathcal {F}}, \quad \eta _{2}^{0} = 0 \quad \text{ on } \ \partial {\mathcal {S}} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q}< 1, \\ \nabla _{s} \eta _{2}^{0} \cdot n_{S} = 0 \text{ on } \partial {\mathcal {S}} &{} \text{ if } \quad \frac{1}{p} + \frac{1}{2q} < \frac{1}{2}. \end{array}\right. } \end{aligned}$$
(4.31)

We deduce from Theorem 4.6 the following result

Corollary 4.7

Let \(p,q\in (1,\infty )\) with \(\displaystyle \frac{1}{p} + \frac{1}{2q} \ne \frac{1}{2},\) \(\displaystyle \frac{1}{p} + \frac{1}{2q} \ne 1\) and let \(\beta \in [0,\beta _{0}],\) where \(\beta _{0}\) is the constant in Theorem 4.6. Assume

$$\begin{aligned}&v^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {F}}),\quad \eta _{1}^{0} \in B^{2(2-1/p)}_{q,p}({\mathcal {S}}), \quad \eta _{2}^{0} \in B^{2(1-1/p)}_{q,p}({\mathcal {S}}),\nonumber \\&f \in L^{p}_\beta (0,\infty ;L^{q}({\mathcal {F}})), \quad h \in L^{p}_\beta (0,\infty ;L^{q}({\mathcal {S}})) \end{aligned}$$
(4.32)

satisfy the compatibility conditions (4.30) and (4.31). Then the system (4.2) admits a unique strong solution

$$\begin{aligned}&v \in W^{1,2}_{p,q,\beta }((0,\infty ) ; {\mathcal {F}}), \quad \pi \in L^{p}_\beta (0,\infty ;W_{m}^{1,q}({\mathcal {F}})), \\&\eta _{1} \in W^{2,4}_{p,q,\beta }((0,\infty ) ; {\mathcal {S}}) \cap L^{p}(0,\infty ;L^{q}_{m}({\mathcal {S}})), \\&\eta _{2} \in W^{1,2}_{p,q,\beta }((0,\infty ) ; {\mathcal {S}}) \cap L^{p}(0,\infty ;L^{q}_{m}({\mathcal {S}})). \end{aligned}$$

Moreover, there exists a constant \(C_{L}\) depending on pq and the geometry such that

$$\begin{aligned}&\left\| v\right\| _{W^{1,2}_{p,q,\beta }((0,\infty ) ; {\mathcal {F}})} + \left\| \pi \right\| _{L^{p}_\beta (0,\infty ;W_{m}^{1,q}({\mathcal {F}}))} + \left\| \eta _{1}\right\| _{W^{2,4}_{p,q,\beta }((0,\infty ) ; {\mathcal {S}})} + \left\| \eta _{2}\right\| _{W^{1,2}_{p,q,\beta } ((0,\infty ) ; {\mathcal {S}})} \nonumber \\&\quad \leqslant C_{L} \Big ( \Vert v^{0}\Vert _{B^{2(1-1/p)}_{q,p} ({\mathcal {F}})} + \Vert \eta _{1}^{0}\Vert _{B^{2(2-1/p)}_{q,p} ({\mathcal {S}})} + \Vert \eta _{2}^{0} \Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {S}})} \nonumber \\&\quad \qquad + \left\| f\right\| _{L^{p}_\beta (0,\infty ;L^{q} ({\mathcal {F}}))} + \left\| h\right\| _{L^{p}_\beta (0,\infty ;L^{q}({\mathcal {S}}))} \Big ). \end{aligned}$$
(4.33)

Proof

Let us first consider the case \(\beta =0.\) Using (4.25), (4.15) and Lemma 4.5 we can also verify that \(\overline{h} \in L^{p}(0,\infty ;L^{q}_{m}({\mathcal {S}})).\)

The compatibility conditions (4.30), (4.31) and the interpolation results [3, Theorem 3.4] and [4, Theorem 4.9.1 and Example 4.9.3]) yield

$$\begin{aligned} \begin{bmatrix} {\mathcal {P}} v^{0}, \eta _{1}^{0}, \eta _{2}^{0} \end{bmatrix}^{\top } \in \left( {\mathcal {X}}, {\mathcal {D}}({\mathcal {A}}_{FS}) \right) _{1-1/p, p} \end{aligned}$$

and

$$\begin{aligned} \begin{bmatrix} {\mathcal {P}} f, 0, \overline{h} \end{bmatrix}^{\top } \in L^p(0,\infty ;{\mathcal {X}}). \end{aligned}$$

From Theorem 4.6, we know that \({\mathcal {A}}_{FS}\) generates an analytic exponentially stable semigroup on \({\mathcal {X}}\) and is a \({\mathcal {R}}\)-sectorial operator on \({\mathcal {X}}.\) Therefore by Corollary 3.4

$$\begin{aligned} ({\mathcal {P}}v, \eta _{1}, \eta _{2}) \in L^{p}(0,\infty ;{\mathcal {D}}({\mathcal {A}}_{FS})) \cap W^{1,p}(0,\infty ;{\mathcal {X}}). \end{aligned}$$

We deduce from (4.23), (4.7) and (4.27) that \(v \in W^{1,2}_{p,q}((0,\infty ) ; {\mathcal {F}})\) and next using relations (4.11), (4.14) and (4.15), we obtain \(\pi \in L^{p}(0,\infty ;W^{1,q}_{m}({\mathcal {F}})).\)

The case \(\beta > 0\) can be reduced to the previous case by multiplying all the functions by \(e^{\beta t}\) and using the fact that \({\mathcal {A}}_{FS} + \beta \) is a \({\mathcal {R}}\)-sectorial operator and generates an exponentially stable semigroup. \(\square \)

5 Local in Time Existence

The aim of this section is to prove Theorems 1.1 and 2.2. Throughout this section we assume the following

Assumption 5.1

\(\eta _{1}^{0} = 0\), \((p,q) \in (1,\infty )\) satisfies (1.15) and \((\eta _2^0,v^0)\) satisfies (2.18), (2.19), (2.20).

For \(T>0\) and \(R>0\), we define \({\mathbb {S}}_{T,R}\) as follows

$$\begin{aligned} {\mathbb {S}}_{T,R} := \Big \{ (f, h) \in L^{p}(0,T;L^{q}({\mathcal {F}}))\times L^{p}(0,T;L^{q}({\mathcal {S}})) \ ; \ \Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}\leqslant R \Big \}. \end{aligned}$$
(5.1)

In order to prove Theorem 2.2, we show that for R fixed and for T small, we can define the map

$$\begin{aligned} {\mathcal {N}}_{T,R}: {\mathbb {S}}_{T,R} \longrightarrow {\mathbb {S}}_{T,R} \quad (f, h) \longmapsto (F(v, \pi , \eta ), H(v, \pi , \eta )), \end{aligned}$$
(5.2)

where \((v, \pi , \eta )\) is the solution to the system (4.2) in \((0,T) \times {\mathcal {F}}\) (see Corollary 4.7) and where F and H are given by (2.13)–(2.14). Then we show that for T small enough and R fixed \({\mathcal {N}}_{T,R}({\mathbb {S}}_{T,R}) \subset {\mathbb {S}}_{T,R}\) (see Proposition 5.2 below) and that, \({\mathcal {N}}_{T,R}|_{{\mathbb {S}}_{T,R}}\) is a strict contraction (see Proposition 5.3 below). This shows that \({\mathcal {N}}_{T,R}\) admits a unique fixed point and allows us to deduce Theorem 2.2.

First, we deduce from Corollary 4.7 that

$$\begin{aligned} \left\| v\right\| _{W^{1,2}_{p,q}((0,T) ; {\mathcal {F}})} + \left\| \pi \right\| _{L^{p}(0,T;W_{m}^{1,q}({\mathcal {F}}))} + \left\| \eta \right\| _{W^{2,4}_{p,q}((0,T) ; {\mathcal {S}})} \leqslant C(\Vert v^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {F}})} +\Vert \eta _{2}^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {S}})}+R). \end{aligned}$$
(5.3)

We take in what follows

$$\begin{aligned} R:=\Vert v^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {F}})} + \Vert \eta _{2}^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {S}})} \end{aligned}$$

and the constants below may depend on R, but not on T. In order to simplify the computation, we also assume that \(T\in (0,1)\).

With these conventions, by using [44, (7), p.196], we have that for any \(s_{1}\in (0,2(1-1/p))\), with \(s_{1}\) not an integer,

$$\begin{aligned} \Vert \eta \Vert _{L^{\infty }(0,T;W^{2+s_{1},q}({\mathcal {S}}))} +\Vert \eta \Vert _{W^{1,\infty }(0,T;W^{s_{1},q}({\mathcal {S}}))} +\Vert v\Vert _{L^{\infty }(0,T;W^{s_{1},q}({\mathcal {F}}))} \leqslant C. \end{aligned}$$
(5.4)

Since \(\eta (0,\cdot ) = 0,\) we have

$$\begin{aligned} \Vert \eta \Vert _{L^{\infty }(0,T;W^{2,q}({\mathcal {S}}))} \leqslant C T^{1/p'} \Vert \partial _{t} \eta \Vert _{L^{p}(0,T;W^{2,q}({\mathcal {S}}))} \leqslant C T^{1/p'}. \end{aligned}$$
(5.5)

Thus, by interpolation between (5.4) and (5.5) ([43, Theorem 2, p. 317]), we deduce that for any \(s_{1}\in (0,2(1-1/p))\), there exists \(\varepsilon =\varepsilon (s_1)>0\) such that

$$\begin{aligned} \Vert \eta \Vert _{L^{\infty }(0,T;W^{2+s_{1},q}({\mathcal {S}}))} \leqslant C T^\varepsilon . \end{aligned}$$
(5.6)

From (1.15), there exists \(s_{1}\in (0,2(1-1/p))\), such that \(s_1+1>3/q\) and thus with the Sobolev embeddings, we deduce that

$$\begin{aligned} \Vert \eta \Vert _{L^{\infty }(0,T; C^{1}(\overline{{\mathcal {S}}}))} \leqslant C T^{\varepsilon }. \end{aligned}$$
(5.7)

Therefore, for T small enough, \(\eta (t,\cdot )\) satisfies (2.1) for all \(t\in [0,T]\) where \(c_{0}\) is defined in (2.5). We can thus construct X by (2.6) so that \(X(t,\cdot )\) is a \(C^{1}\)-diffeomorphism from \({\mathcal {F}}\) onto \({\mathcal {F}}(\eta (t)).\) We can also consider \(F(v, \pi , \eta )\) and \(H(v, \pi , \eta ))\) given by (2.13)–(2.14). In order to estimate these expressions, we also note that by (real or complex) interpolation ([43, Theorem 2, p. 317]) for \(\theta \in (0,1)\),

$$\begin{aligned} \Vert v (t,\cdot ) \Vert _{W^{s_2,q}({\mathcal {F}})} \leqslant C \Vert v(t,\cdot ) \Vert _{W^{s_1,q}({\mathcal {F}})}^{1-\theta }\Vert v (t,\cdot ) \Vert _{W^{2,q}({\mathcal {F}})}^{\theta }, \quad s_2=2\theta +(1-\theta )s_1, \end{aligned}$$

if \(s_2\) is not an integer. We can find \(\theta \in (0,1/3)\) and \(s_{1}\in (0,2(1-1/p))\) such that \(s_2 \geqslant 2/q\) so that by Sobolev embeddings,

$$\begin{aligned} \Vert v \Vert _{L^{3p}(0,T;L^{3q}({\mathcal {F}}))} \leqslant CT^\varepsilon \Vert v \Vert _{L^\infty (0,T;W^{s_1,q}({\mathcal {F}}))}^{1-\theta }\Vert v \Vert _{L^p(0,T;W^{2,q}({\mathcal {F}}))}^{\theta } \leqslant C T^\varepsilon \end{aligned}$$
(5.8)

and similarly,

$$\begin{aligned}&\Vert \nabla ^2_s \eta \Vert _{L^{3p}(0,T;L^{3q}({\mathcal {S}}))} +\Vert \partial _t \eta \Vert _{L^{3p}(0,T;L^{3q}({\mathcal {S}}))} \leqslant CT^\varepsilon , \end{aligned}$$
(5.9)
$$\begin{aligned}&\Vert \nabla v \Vert _{L^{3p/2}(0,T;L^{3q/2}({\mathcal {F}}))} +\Vert \nabla ^3_s \eta \Vert _{L^{3p/2}(0,T;L^{3q/2}({\mathcal {S}}))} +\Vert \nabla _s \partial _t \eta \Vert _{L^{3p/2}(0,T;L^{3q/2}({\mathcal {S}}))} \leqslant CT^\varepsilon . \end{aligned}$$
(5.10)

We are now in position to prove the following result:

Proposition 5.2

With the above assumptions (in particular Assumption 5.1), there exists \(T>0\) small enough such that the map \({\mathcal {N}}_{T,R}\) (see (5.2)) is well-defined and satisfies \({\mathcal {N}}_{T,R}({\mathbb {S}}_{T,R}) \subset {\mathbb {S}}_{T,R}\).

Proof

From (2.4) and (5.7), we deduce that for \(T>0\) small enough

$$\begin{aligned}&\Vert \nabla X - I_3\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} +\Vert \nabla Y(X) - I_3\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} \nonumber \\&\quad +\Vert a(X) - I_3\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} +\Vert b - I_3\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} \leqslant C T^{\varepsilon }, \end{aligned}$$
(5.11)
$$\begin{aligned}&\Vert \det (\nabla X) - 1\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} \leqslant C T^{\varepsilon }, \quad \frac{1}{2} \leqslant \Vert \det \nabla X \Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} \leqslant \frac{3}{2}, \end{aligned}$$
(5.12)
$$\begin{aligned}&\Vert \nabla X \Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} +\Vert \nabla Y(X)\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} +\Vert a(X)\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} +\Vert b\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} \leqslant C. \end{aligned}$$
(5.13)

We recall that a and b are defined by (2.11).

By using standard properties of linear algebra, we have that

$$\begin{aligned} a(X) = \frac{\nabla X}{\det (\nabla X)} \end{aligned}$$
(5.14)

and thus for all ijk,

$$\begin{aligned}&\left| \frac{\partial a_{ik}}{\partial x_j}(X) \right| \leqslant C \left| \nabla ^2 X \right| \leqslant C \left( |\eta |+|\nabla _s \eta |+ |\nabla _s^2 \eta | \right) , \end{aligned}$$
(5.15)
$$\begin{aligned}&\left| \frac{\partial ^2 a_{ik}}{\partial x_j^2}(X) \right| \leqslant C\left( \left| \nabla ^2 X \right| ^2 +\left| \nabla ^3 X \right| \right) \leqslant C \left( \left( |\eta |+|\nabla _s \eta |+ |\nabla _s^2 \eta | \right) ^2 + |\nabla _s^3 \eta | \right) , \end{aligned}$$
(5.16)
$$\begin{aligned}&\left| \partial _t a(X) \right| \leqslant C\left( \left| \nabla ^2 X \right| +\left| \nabla \partial _t X \right| \right) \leqslant C \left( |\eta |+|\nabla _s \eta |+ |\nabla _s^2 \eta | +|\nabla _s \partial _t \eta | \right) . \end{aligned}$$
(5.17)

We also have

$$\begin{aligned}&\left| \partial _t Y(X) \right| \leqslant C\left| \partial _t X \right| \leqslant C|\partial _t \eta |, \end{aligned}$$
(5.18)
$$\begin{aligned}&\left| \frac{\partial ^2 Y_\ell }{\partial x_j^2}(X)\right| \leqslant C\left| \nabla ^2 X \right| \leqslant C \left( |\eta |+|\nabla _s \eta |+ |\nabla _s^2 \eta | \right) . \end{aligned}$$
(5.19)

Combining the above estimates with (5.8), (5.9) and (5.10), we deduce that F defined by (2.13) satisfies

$$\begin{aligned} \left\| {F}(v, \pi , \eta ) \right\| _{L^{p}(0,T;L^{q}({\mathcal {F}}))} \leqslant CT^\varepsilon . \end{aligned}$$
(5.20)

Using trace theorems, we deduce from (5.3) and from (5.10) that

$$\begin{aligned} \Vert \nabla v \Vert _{L^p(0,T;L^q(\partial {\mathcal {F}}))} \leqslant C, \quad \Vert v \Vert _{L^{3p/2}(0,T;L^{3q/2}(\partial {\mathcal {F}}))} \leqslant CT^\varepsilon . \end{aligned}$$

From this relation, the above estimates and (5.9), (5.10), we deduce that H defined by (2.14) satisfies

$$\begin{aligned} \left\| {H}(v, \pi , \eta ) \right\| _{L^{p}(0,T;L^{q}({\mathcal {S}}))} \leqslant CT^\varepsilon . \end{aligned}$$
(5.21)

Relations (5.20) and (5.21) yield that \({\mathcal {N}}({\mathcal {B}}_{T,R}) \subset {\mathcal {B}}_{T,R}\) for T small enough. \(\square \)

Proposition 5.3

With the above assumptions (in particular Assumption 5.1), there exists \(T>0\) small enough such that the map \({\mathcal {N}}_{T,R}\) (see (5.2)) is a strict contraction on \({\mathbb {S}}_{T,R}\).

Proof

The proof is similar to the proof of Proposition 5.2, we only give the main ideas and omit the details. We consider \((f^{(i)}, h^{(i)})\), \(i=1,2\). We have

$$\begin{aligned}&{\mathcal {N}}_{T,R}(f^{(1)}, h^{(1)}) -{\mathcal {N}}_{T,R}(f^{(2)}, h^{(2)}) \nonumber \\&\quad =(F(v^{(1)}, \pi ^{(1)}, \eta ^{(1)})-F(v^{(2)}, \pi ^{(2)}, \eta ^{(2)}), H(v^{(1)}, \pi ^{(1)}, \eta ^{(1)})-H(v^{(2)}, \pi ^{(2)}, \eta ^{(2)})), \end{aligned}$$
(5.22)

where \((v^{(i)}, \pi ^{(i)}, \eta ^{(i)})\) is the solution to the system (4.2) in \((0,T) \times {\mathcal {F}}\) (see Corollary 4.7) associated with \((f^{(i)}, h^{(i)})\), \(i=1,2\) and where F and H are given by (2.13)-(2.14). By taking T as Proposition 5.2, we have for each i that \((v^{(i)}, \pi ^{(i)}, \eta ^{(i)})\) satisfies the same property obtained in the proof of Proposition 5.2 and in particular, \(X^{(i)}\), \(Y^{(i)}\), \(a^{(i)}\), \(b^{(i)}\) defined by (2.6) and (2.11) satisfy also the same properties obtained in the proof of Proposition 5.2.

We write

$$\begin{aligned} v=v^{(1)}-v^{(2)}, \quad pi=\pi ^{(1)}-\pi ^{(2)}, \quad \eta =\eta ^{(1)}-\eta ^{(2)}, \quad f=f^{(1)}-f^{(2)}, \quad g=g^{(1)}-g^{(2)}, \end{aligned}$$

Applying Corollary 4.7, we first obtain

$$\begin{aligned} \left\| v\right\| _{W^{1,2}_{p,q}((0,T) ; {\mathcal {F}})} + \left\| \pi \right\| _{L^{p}(0,T;W_{m}^{1,q}({\mathcal {F}}))} + \left\| \eta \right\| _{W^{2,4}_{p,q}((0,T) ; {\mathcal {S}})} \leqslant C(\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} +\Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}). \end{aligned}$$
(5.23)

As in the proof of Proposition 5.2, the constants below may depend on R, but not on T and we assume \(T\in (0,1)\) to simplify. Following the proof of (5.7), we can obtain

$$\begin{aligned} \Vert \eta \Vert _{L^{\infty }(0,T; C^{1}(\overline{{\mathcal {S}}}))} \leqslant C T^{\varepsilon }(\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} +\Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}) \end{aligned}$$
(5.24)

and following the proof of (5.8), (5.9) and (5.10), we deduce

$$\begin{aligned}&\Vert v \Vert _{L^{3p}(0,T;L^{3q}({\mathcal {F}}))} +\Vert \nabla ^2_s \eta \Vert _{L^{3p}(0,T;L^{3q}({\mathcal {S}}))} +\Vert \partial _t \eta \Vert _{L^{3p}(0,T;L^{3q}({\mathcal {S}}))}\nonumber \\&\quad \leqslant CT^\varepsilon (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p} (0,T;L^{q}({\mathcal {S}}))}) \end{aligned}$$
(5.25)

and

$$\begin{aligned}&\Vert \nabla v \Vert _{L^{3p/2}(0,T;L^{3q/2}({\mathcal {F}}))} +\Vert \nabla ^3_s \eta \Vert _{L^{3p/2}(0,T;L^{3q/2}({\mathcal {S}}))} +\Vert \nabla _s \partial _t \eta \Vert _{L^{3p/2}(0,T;L^{3q/2}({\mathcal {S}}))}\nonumber \\&\quad \leqslant CT^\varepsilon (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p} (0,T;L^{q}({\mathcal {S}}))}). \end{aligned}$$
(5.26)

Using trace theorems, we deduce from the above estimates that

$$\begin{aligned}&\Vert \nabla v \Vert _{L^p(0,T;L^q(\partial {\mathcal {F}}))} \leqslant C(\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}),\\&\Vert v \Vert _{L^{3p/2}(0,T;L^{3q/2}(\partial {\mathcal {F}}))} \leqslant CT^\varepsilon (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}). \end{aligned}$$

We also deduce from the above estimate and from (2.6) that

$$\begin{aligned}&\Vert \nabla X^{(1)} -\nabla X^{(2)}\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} +\Vert \nabla Y^{(1)}(X^{(1)}) - \nabla Y^{(2)}(X^{(2)})\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))}\nonumber \\&\quad +\Vert a^{(1)}(X^{(1)}) - a^{(2)}(X^{(2)})\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} +\Vert b^{(1)} - b^{(2)}\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))}\nonumber \\&\quad +\Vert \det (\nabla X^{(1)}) - \det (\nabla X^{(2)})\Vert _{L^{\infty }(0,T;C^0(\overline{{\mathcal {F}}}))} \leqslant C T^{\varepsilon } (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} +\Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}). \end{aligned}$$
(5.27)

From (5.14) and from the above estimates, we obtain for all ijk,

$$\begin{aligned}&\left| \frac{\partial a_{ik}^{(1)}}{\partial x_j}(X^{(1)})-\frac{\partial a_{ik}^{(2)}}{\partial x_j}(X^{(2)}) \right| \leqslant C |\nabla _s^2 \eta | +C T^{\varepsilon } (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}) \left( |\nabla _s^2 \eta ^{(1)}|+|\nabla _s^2 \eta ^{(2)}|\right) , \end{aligned}$$
(5.28)
$$\begin{aligned}&\left| \frac{\partial ^2 a_{ik}^{(1)}}{\partial x_j^2}(X^{(1)}) -\frac{\partial ^2 a_{ik}^{(2)}}{\partial x_j^2}(X^{(2)}) \right| \leqslant C\left( \left( |\nabla _s^2 \eta ^{(1)}|+|\nabla _s^2 \eta ^{(2)}|\right) |\nabla _s^2 \eta | + |\nabla _s^3 \eta |\right) ,\nonumber \\&\quad +C T^{\varepsilon } (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p}(0,T;L^{q} ({\mathcal {S}}))}) \left( 1+ |\nabla _s^2 \eta ^{(1)}|^2 +|\nabla _s^3 \eta ^{(1)}|+ |\nabla _s^2 \eta ^{(2)}|^2 + |\nabla _s^3 \eta ^{(2)}| \right) , \end{aligned}$$
(5.29)
$$\begin{aligned}&\left| \partial _t a^{(1)}(X^{(1)})-\partial _t a^{(2)}(X^{(2)}) \right| \leqslant C \left( |\nabla _s^2 \eta | + |\nabla _s \partial _t \eta | \right) \nonumber \\&\quad +C T^{\varepsilon } (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} +\Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}) \left( 1+ |\nabla _s^2 \eta ^{(1)}| + |\nabla _s \partial _t \eta ^{(1)}| \right) , \end{aligned}$$
(5.30)
$$\begin{aligned}&\left| \partial _t Y^{(1)}(X^{(1)})-\partial _t Y^{(2)}(X^{(2)}) \right| \leqslant C|\partial _t \eta | +C T^{\varepsilon } (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} + \Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))})\left( |\partial _t \eta ^{(1)}|+|\partial _t \eta ^{(2)}|\right) , \end{aligned}$$
(5.31)
$$\begin{aligned}&\left| \frac{\partial ^2 Y^{(1)}_\ell }{\partial x_j^2}(X^{(1)})-\frac{\partial ^2 Y^{(2)}_\ell }{\partial x_j^2}(X^{(2)})\right| \leqslant C |\nabla _s^2 \eta |\nonumber \\&\quad +C T^{\varepsilon } (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} +\Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}) \left( 1+ |\nabla _s^2\eta ^{(1)}|+ |\nabla _s^2 \eta ^{(2)}| \right) . \end{aligned}$$
(5.32)

Combining the above estimates with (5.11)–(5.19), with (5.8)–(5.10) and with (5.25)–(5.26), we deduce that

$$\begin{aligned}&\left\| {\mathcal {N}}_{T,R}(f^{(1)}, h^{(1)}) -{\mathcal {N}}_{T,R}(f^{(2)}, h^{(2)}) \right\| _{L^{p}(0,T;L^{q}({\mathcal {F}})) \times L^{p}(0,T;L^{q}({\mathcal {S}}))} \nonumber \\&\quad \leqslant C T^{\varepsilon } (\Vert f\Vert _{L^{p}(0,T;L^{q}({\mathcal {F}}))} +\Vert h\Vert _{L^{p}(0,T;L^{q}({\mathcal {S}}))}). \end{aligned}$$
(5.33)

Thus for T small enough, we deduce the result. \(\square \)

6 Global in Time Existence

The aim of this section is to prove Theorem 1.2 and Theorem 2.3. The proof is similar to the proof of Theorem 1.1 and Theorem 2.2 given in Sect. 5. Throughout this section we assume the following

Assumption 6.1

\((p,q) \in (1,\infty )\) satisfies (1.15) and \((\eta _1^0,\eta _2^0,v^0)\) satisfies (2.18), (2.19), (2.20).

Let us fix \(\beta \in [0, \beta _{0}],\) where \(\beta _{0}\) is introduced in Corollary 4.7 and for \(R>0\), we define \({\mathbb {S}}_{R}\) as follows

$$\begin{aligned} {\mathbb {S}}_{R} := \Big \{ (f, h) \in L^{p}_\beta (0,\infty ;L^{q}({\mathcal {F}}))\times L^{p}_\beta (0,\infty ;L^{q}({\mathcal {S}})) \ ; \ \Vert f\Vert _{L^{p}_\beta (0,\infty ;L^{q}({\mathcal {F}}))} +\Vert h\Vert _{L^{p}_\beta (0,\infty ;L^{q}({\mathcal {S}}))}\leqslant R \Big \}. \end{aligned}$$
(6.1)

We take in what follows

$$\begin{aligned} R:=\Vert v^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {F}})} +\Vert \eta _{2}^{0}\Vert _{B^{2(1-1/p)}_{q,p}({\mathcal {S}})} +\Vert \eta _{1}^{0}\Vert _{B^{2(2-1/p)}_{q,p}({\mathcal {S}})} \end{aligned}$$

and to simplify the computation, we assume that \(R\in (0,1)\).

In order to prove Theorem 2.3, we show that for R small, we can define the map

$$\begin{aligned} {\mathcal {N}}_{R}: {\mathbb {S}}_{R} \longrightarrow {\mathbb {S}}_{R} \quad (f, h) \longmapsto (F(v, \pi , \eta ), H(v, \pi , \eta )), \end{aligned}$$
(6.2)

where \((v, \pi , \eta )\) is the solution to the system (4.2) in \((0,\infty ) \times {\mathcal {F}}\) (see Corollary 4.7) and where F and H are given by (2.13)–(2.14). Then we show that for R small enough \({\mathcal {N}}_R({\mathbb {S}}_{R}) \subset {\mathbb {S}}_{R}\) (see Proposition 6.2 below) and that, \({\mathcal {N}}_R|_{{\mathbb {S}}_{R}}\) is a strict contraction (see Proposition 6.3 below). This shows that \({\mathcal {N}}_{R}\) admits a unique fixed point and allows us to deduce Theorem 2.3.

First, we deduce from Corollary 4.7 that

$$\begin{aligned} \left\| v\right\| _{W^{1,2}_{p,q,\beta }((0,\infty ) ; {\mathcal {F}})} + \left\| \pi \right\| _{L^{p}_\beta (0,\infty ;W_{m}^{1,q}({\mathcal {F}}))} +\left\| \eta \right\| _{W^{2,4}_{p,q,\beta }((0,\infty ) ; {\mathcal {S}})} \leqslant CR. \end{aligned}$$
(6.3)

By using [44, (7), p.196] and the Sobolev embeddings, we deduce from the above estimate

$$\begin{aligned} \Vert \eta \Vert _{L^{\infty }_\beta (0,\infty ; C^{1}(\overline{{\mathcal {S}}}))} \leqslant C R. \end{aligned}$$
(6.4)

Therefore, for R small enough, \(\eta (t,\cdot )\) satisfies (2.1) for all \(t\in [0,\infty )\) where \(c_{0}\) is defined in (2.5). We can thus construct X by (2.6) so that \(X(t,\cdot )\) is a \(C^{1}\)-diffeomorphism from \({\mathcal {F}}\) onto \({\mathcal {F}}(\eta (t)).\) We can also consider \(F(v, \pi , \eta )\) and \(H(v, \pi , \eta ))\) given by (2.13)–(2.14).

As in the previous section, we use (real or complex) interpolation results ([43, Theorem 2, p. 317]) to deduce that

$$\begin{aligned} \Vert v (t,\cdot ) \Vert _{W^{s_2,q}({\mathcal {F}})} \leqslant C \Vert v(t,\cdot ) \Vert _{W^{s_1,q}({\mathcal {F}})}^{2/3}\Vert v (t,\cdot ) \Vert _{W^{2,q}({\mathcal {F}})}^{1/3}, \end{aligned}$$

for any \(s_2<2(1+s_1)/3\). Using (1.15), there exists \(s_{1}\in (0,2(1-1/p))\) such that \(s_2 \geqslant 2/q\) so that by Sobolev embeddings,

$$\begin{aligned} \Vert v \Vert _{L^{3p}_\beta (0,\infty ;L^{3q}({\mathcal {F}}))} \leqslant C \Vert v \Vert _{L^\infty _\beta (0,\infty ;W^{s_1,q}({\mathcal {F}}))}^{2/3}\Vert v \Vert _{L^p_\beta (0,\infty ;W^{2,q}({\mathcal {F}}))}^{1/3} \leqslant C R. \end{aligned}$$
(6.5)

and similarly,

$$\begin{aligned} \Vert \nabla ^2_s \eta \Vert _{L^{3p}_\beta (0,\infty ;L^{3q}({\mathcal {S}}))} +\Vert \partial _t \eta \Vert _{L^{3p}_\beta (0,\infty ;L^{3q}({\mathcal {S}}))} \leqslant C R \end{aligned}$$
(6.6)

and

$$\begin{aligned} \Vert \nabla v \Vert _{L^{3p/2}_\beta (0,\infty ;L^{3q/2}({\mathcal {F}}))} +\Vert \nabla ^3_s \eta \Vert _{L^{3p/2}_\beta (0,\infty ;L^{3q/2}({\mathcal {S}}))} +\Vert \nabla _s \partial _t \eta \Vert _{L^{3p/2}_\beta (0,\infty ;L^{3q/2} ({\mathcal {S}}))} \leqslant CR. \end{aligned}$$
(6.7)

Using trace theorems, we deduce from (6.3) and from (6.7) that

$$\begin{aligned} \Vert \nabla v \Vert _{L^p_\beta (0,\infty ;L^q(\partial {\mathcal {F}}))} \leqslant CR, \quad \Vert v \Vert _{L^{3p/2}_\beta (0,\infty ;L^{3q/2} (\partial {\mathcal {F}}))} \leqslant CR. \end{aligned}$$
(6.8)

We are now in position to prove the following result:

Proposition 6.2

With the above assumptions (in particular Assumption 6.1), there exists \(R>0\) small enough such that the map \({\mathcal {N}}_{R}\) (see (6.2)) is well-defined and satisfies \({\mathcal {N}}_{R}({\mathbb {S}}_{R}) \subset {\mathbb {S}}_{R}\).

Proof

From (2.4) and (6.4), we deduce that for \(T>0\) small enough

$$\begin{aligned}&\Vert \nabla X - I_3\Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} +\Vert \nabla Y(X) - I_3\Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))}\nonumber \\&\quad +\Vert a(X) - I_3\Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} +\Vert b - I_3\Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} \leqslant C R, \end{aligned}$$
(6.9)
$$\begin{aligned}&\Vert \det (\nabla X) - 1\Vert _{L^{\infty }_\beta (0,\infty ;C^0 (\overline{{\mathcal {F}}}))} \leqslant C R, \quad \frac{1}{2} \leqslant \Vert \det \nabla X \Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} \leqslant \frac{3}{2}, \end{aligned}$$
(6.10)
$$\begin{aligned}&\Vert \nabla X \Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} +\Vert \nabla Y(X)\Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} +\Vert a(X)\Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} +\Vert b\Vert _{L^{\infty }_\beta (0,\infty ;C^0(\overline{{\mathcal {F}}}))} \leqslant C. \end{aligned}$$
(6.11)

We recall that a and b are defined by (2.11).

Using the above estimates, relations (5.15)–(5.19), (6.3), (6.5), (6.6), (6.7) and (6.8) we deduce that F and H defined by (2.13), (2.14) satisfy

$$\begin{aligned} \left\| {F}(v, \pi , \eta ) \right\| _{L^{p}_\beta (0,\infty ;L^{q}({\mathcal {F}}))} +\left\| {H}(v, \pi , \eta ) \right\| _{L^{p}_\beta (0,\infty ;L^{q}({\mathcal {S}}))} \leqslant CR^2, \end{aligned}$$
(6.12)

which yields that \({\mathcal {N}}_R({\mathbb {S}}_{R}) \subset {\mathcal {S}}_{R}\) for R small enough. \(\square \)

We can also prove the following result by following the method used to prove Proposition 5.3 (we omit the proof).

Proposition 6.3

With the above assumptions (in particular Assumption 6.1), there exists \(R>0\) small enough such that the map \({\mathcal {N}}_{R}\) (see (6.2)) is a strict contraction on \({\mathbb {S}}_{R}\).

By combining Propositions 6.2 and 6.3, we deduce Theorem 2.3.