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

Existence and Regularity Results for a Nonlinear Fluid-Structure Interaction Problem with Three-Dimensional Structural Displacement

Sunčica Čanić1 and Boris Muha2 and Krutika Tawri1
1 Department of Mathematics, University of California Berkeley, CA, USA.
2 Department of Mathematics, University of Zagreb, Croatia.
Abstract.

In this paper we investigate a nonlinear fluid-structure interaction (FSI) problem involving the Navier-Stokes equations, which describe the flow of an incompressible, viscous fluid in a 3D domain interacting with a thin viscoelastic lateral wall. The wall’s elastodynamics is modeled by a two-dimensional plate equation with fractional damping, accounting for displacement in all three directions. The system is nonlinearly coupled through kinematic and dynamic conditions imposed at the time-varying fluid-structure interface, whose location is not known a priori.

We establish three key results, particularly significant for FSI problems that account for vector displacements of thin structures. Specifically, we first establish a hidden spatial regularity for the structure displacement, which forms the basis for proving that self-contact of the structure will not occur within a finite time interval. Secondly, we demonstrate temporal regularity for both the structure and fluid velocities, which enables a new compactness result for three-dimensional structural displacements. Finally, building on these regularity results, we prove the existence of a local-in-time weak solution to the FSI problem. This is done through a constructive proof using time discretization via the Lie operator splitting method.

These results are significant because they address the well-known issues associated with the analysis of nonlinearly coupled FSI problems capturing vector displacements of elastic/viscoelastic structures in 3D, such as spatial and temporal regularity of weak solutions and their well-posedness.

1. Introduction

We study a fluid-structure interaction (FSI) problem involving an incompressible, viscous fluid flowing within a three-dimensional domain, bounded by thin compliant lateral walls. The fluid dynamics is governed by the three-dimensional Navier-Stokes equations, while the structural dynamics is modeled by a linear viscoelastic plate equation incorporating fractional damping.

The interaction between the fluid and the structure is characterized by a fully coupled system, with kinematic and dynamic coupling conditions that enforce the continuity of velocities and contact forces at the dynamic fluid-structure interface. This coupling introduces a significant geometric nonlinearity to the problem, as the fluid domain’s location is not known a priori and is instead one of the unknowns in the problem.

The field of fluid-structure interaction (FSI) analysis has seen tremendous progress over the past two decades (see, e.g., [2, 16, 3] and references therein). In this paper, we focus on the interaction between fluid flow and a plate structure, so our brief literature review emphasizes the analysis of moving boundary FSI problems where the structural dynamics is described by lower-dimensional models.

Most existing works involving lower-dimensional models interacting with viscous incompressible fluids consider the case of scalar displacement, where the structure deforms only in a fixed direction typically normal to the reference configuration. The theory of weak solutions in this context is well developed, see [4, 11, 22, 33, 27, 19] and references therein. Strong solutions have also been studied in this context, as can be found in e.g. [18, 12, 13, 20] and refences within.

The case of three-dimensional (3D3D) structural displacement, where the structure can deform in all three spatial directions (vector displacement), is less well-studied, with only a few works addressing weak solutions. In [34], the authors investigated an FSI problem involving a 3D3D fluid flow interacting with a two-dimensional (2D2D) cylindrical shell supported by a mesh of elastic rods. They proved the existence of a weak solution under additional assumptions that ensured the structure’s displacement remained Lipschitz continuous in space at all times. In the 2D2D fluid and 1D1D structure scenario, several results have been obtained. The local-in-time existence of weak solutions to FSI problems where 2D2D Navier-Stokes equations are coupled with 1D1D plate or shell equations via the Navier slip boundary condition is established in [24]. More recently, [17] considered an FSI problem where the structure is described by a nonlinear beam equation with a term that penalizes compression, preventing domain degeneracy. Additionally, recent work [14] has established the existence of local-in-time strong solutions for an FSI problem where the structure is modeled as a linear plate.

To the best of our knowledge, the present work is the first to establish the existence of weak (finite energy) solutions for a moving boundary FSI problem where a 3D3D fluid is coupled with a 2D2D plate with 3D3D vector displacement.

The primary challenge in developing a theory for FSI problems involving structure equations accounting for 3D vector displacements is managing the difficulties associated with self-contact. Specifically, proving existence results requires ruling out fluid domain degeneracy, i.e., preventing self-contact of the structure over the time interval where the solution is defined. In particular, in the case of 3D3D displacement, the standard energy estimates do not provide sufficient regularity of the structure to analyze issues with self-contact.

Another challenge in developing a theory for FSI problems with vector displacements and with the geometrically nonlinear coupling is designing suitable compactness arguments for the fluid and structure velocities whose energy-based regularity estimates are insufficient to deduce compactness.

In this manuscript we address both of those challenges by proving two “hidden” regularity results for weak solutions of such problems. The first regularity result improves the spatial regularity of structure displacement over the “basic” regularity provided by the energy estimate, and the second regularity result improves the temporal regularity of fluid and structure velocities over that provided by the energy estimates. The first is used in ensuring non-degeneracy of the fluid domain, while the second is used in establishing compactness arguments for the fluid and structure velocities in this class of nonlinear moving boundary problems. Finally, building on these regularity results we prove the existence of a weak solution to a FSI involving 3D Navier-Stokes equations coupled to the 2D plate equation with fractional damping accounting for 3D vector displacements. Thus, the main results of this paper are three-pronged: (1) We provide a hidden regularity result for 2D plates with fractional damping allowing 3D vectoral displacements, (2) We provide a hidden temporal regularity result for fluid and structure velocities in a nonlinearly coupled 3D fluid-2D plate FSI problem with fractional damping and 3D vector displacements, and (3) We prove a well-posedness result for weak solutions of the nonlinearly coupled 3D fluid-2D plate FSI problem with fractional damping and 3D vector displacements.

More precisely, in terms of spatial regularity of structure displacement, in Section 3.1 we prove that the structure displacement belongs to the space LtHx2+δL_{t}^{\infty}H^{2+\delta}_{x} for a sufficiently small δ>0\delta>0, which is crucial for establishing that the structure displacement is Lipschitz continuous in space at any given time, ensuring injectivity of the maps that map the reference configuration of the fluid domain onto the “current” location of the moving domain. This is generally one of the key issues in the analysis of nonlinearly-coupled moving boundary problems with 3D3D (vector) structure displacements. The main ideas behind the proof of this hidden spatial regularity result rely on constructing appropriate test functions for the structure variable and their solenoidal extensions to the fluid domain, which satisfy the kinematic coupling condition. A key step is to formulate a suitable non-homogeneous time-dependent Stokes problem whose solution is used to construct the desirable test functions. This approach generalizes the approach presented in [21] to vector displacements. The technique developed here can be applied to other settings, including nonlinear structure operators that are coercive in H2H^{2}, and different boundary conditions, including the time-dependent inlet/outlet boundary data.

In terms of temporal hidden regularity result for the fluid and structure velocities, in Section 3.2 we prove that the fractional time derivative of order 1/81/8 of the fluid and structure velocities can be uniformly bounded in Lt2Lx2L_{t}^{2}L_{x}^{2}, i.e., we obtain uniform bounds for the fluid and structure velocities in Ntα,2Lx2N_{t}^{\alpha,2}L_{x}^{2}, where Nα,pN^{\alpha,p} is Nikolski space. The key idea is to construct appropriate test functions for the coupled FSI problem by utilizing a time-regularized (averaged) modification of the structure and fluid velocities, similar to the approaches used in [4, 11]. The construction of these time-regularized (averaged) test functions presents several challenges, arising from the motion of the fluid domain, the non-zero longitudinal displacement of the structure, and the mismatch in spatial regularity between the structure velocity and its corresponding test function. Additionally, the test functions for the fluid and structure must satisfy the kinematic coupling condition at the moving boundary, with the fluid test function also needing to satisfy the divergence-free condition within the moving fluid domain. To enforce these conditions, we construct a Bogovskii-type operator on a time-varying domain with a Lipschitz boundary. The construction of the Bogovskii-type operator presented here holds significant potential for applications to analyzing general incompressible flow problems on moving domains involving Lipschitz boundaries.

Finally, in Section 4, we present a constructive proof of the existence of a local-in-time weak solution to a FSI problem between the 3D flow of an incompressible, viscous fluid modeled by the Navier-Stokes equations and a 2D plate with fractional damping modeling elastodynamics of a plate with 3D vector displacements. We employ a Lie operator splitting method, first utilized in the context of FSI in [22] (and further developed in [23, 34] for different FSI settings). The coupled problem is discretized in time and split into a structure subproblem and a fluid subproblem along the dynamic coupling condition. This time discretization via Lie operator splitting yields a sequence of approximate solutions, which is shown to converge, up to a subsequence and in an appropriate sense, to the desired solution.

The structural regularity result from Section 3.1 is crucial in the construction of approximate solutions and the limiting solution, allowing us to obtain the desired solution up to a strictly positive time TT determined by self-intersection of the fluid domain boundary. The temporal regularity result obtained in Section 3.2 is crucial to achieve the compactness of the sequence of approximations of the fluid and structure velocities. This ensures that a subsequence of the approximate solutions converges strongly in the relevant topologies as the time step approaches zero, which allowed us to pass to the limit in the approximate weak formulations to prove that the limits satisfy the continuous weak formulation of the original problem.

In this final step of taking the limit in the approximate weak formations, one needs to deal with one last difficulty associated with general problems on moving domains – the fact that the test functions in weak formulations depend on the fluid domain motion, and thus on the time-discretization step, via structure displacements, in a nontrivial way (through the divergence-free condition). This is a classical problem in FSI problems with nonlinear coupling, see e.g., [22, 23, 25, 24]. Taking the limit in approximate weak formulations requires constructing appropriate test functions which would converge, as the time-discretization step converges to zero, to the test functions of the continuous problem in the norm strong enough to pass to the limit. Indeed, in Section 4.4 we construct such test functions and take the limit in approximate weak formulations to show that the approximate solutions constructed here converge to a weak solution of the continuous problem.

2. Problem setup

We consider the flow a fluid in a periodic channel interacting with a complaint structure that sits atop the fluid domain. See Figure 1. We assume that the structure displacement is periodic, with the reference domain for the structure equations given by

Γ={(x,y,z)3:(x,y)𝕋2,z=1},\Gamma=\{(x,y,z)\in\mathbb{R}^{3}:(x,y)\in\mathbb{T}^{2},z=1\},

where 𝕋2\mathbb{T}^{2} is the 2D torus.

The fluid reference domain, is then given by

𝒪=Γ×(0,1).\mathcal{O}=\Gamma\times(0,1).

We denote by Γr=𝒪Γ\Gamma_{r}=\partial\mathcal{O}\setminus\Gamma the rigid part of the boundary of the fluid reference domain 𝒪\mathcal{O}.

Refer to caption
Figure 1. The fluid domain

In this work we assume that the displacement of the compliant structure, denoted by 𝜼{\boldsymbol{\eta}}, is a vector function with all three components of displacement satisfying a vector equation for a plate with fractional damping, thereby allowing all three components of displacement to be different from zero.

The fluid domain deforms as a result of the interaction between the fluid and the structure. The time-dependent fluid domain in 3D, whose displacement is not known a priori is then given by

𝒪𝜼(t)=𝐀𝜼(t,𝒪),\mathcal{O}_{{{\boldsymbol{\eta}}}}(t)={{\bf A}_{\boldsymbol{\eta}}}(t,\mathcal{O}),

whereas its deformable interface is given by

Γ𝜼(t)=𝐀𝜼(t,Γ),\Gamma_{{\boldsymbol{\eta}}}(t)={{\bf A}_{\boldsymbol{\eta}}}(t,\Gamma),

where 𝐀𝜼{\bf A}_{\boldsymbol{\eta}} is a family of C1C^{1} diffeomorphisms parametrized by time t[0,T]t\in[0,T], such that

(1) 𝐀𝜼(t)=𝐢𝐝+𝜼(t) on Γ,𝐀𝜼(t)|Γ=𝐢𝐝,det𝐀𝜼(t,x,y,z)>0.{\bf A}_{\boldsymbol{\eta}}(t)={\bf id}+{\boldsymbol{\eta}}(t)\text{ on }\Gamma,\quad{{\bf A}_{\boldsymbol{\eta}}}(t)|_{\Gamma}={\bf id},\quad\text{det}\nabla{{\bf A}_{\boldsymbol{\eta}}}(t,x,y,z)>0.

We will now describe the fluid and the structure equations and the two-way coupling that describe the interactions that take place between them.

The fluid subproblem: The fluid flow is modeled by the incompressible Navier-Stokes equations in the 3D time-dependent domains 𝒪𝜼(t)3\mathcal{O}_{{\boldsymbol{\eta}}}(t)\subset\mathbb{R}^{3} :

(2) t𝐮+(𝐮)𝐮=σ𝐮=0}in𝒪𝜼(t)×(0,T),\left.\begin{split}\partial_{t}{\bf u}+({\bf u}\cdot\nabla){\bf u}&=\nabla\cdot\sigma\\ \nabla\cdot{\bf u}&=0\end{split}\right\}\quad{\rm in}\ \mathcal{O}_{{\boldsymbol{\eta}}}(t)\times(0,T),

where 𝐮=(ux,uy,uz){\bf u}=(u_{x},u_{y},u_{z}) is the fluid velocity. The Cauchy stress tensor is given by σ=pI+2ν𝐃(𝐮)\sigma=-pI+2\nu{\bf D}({\bf u}) where pp is the fluid pressure, ν>0\nu>0 is the kinematic viscosity coefficient and 𝐃(𝐮)=12(𝐮+(𝐮)T){\bf D}({\bf u})=\frac{1}{2}(\nabla{\bf u}+(\nabla{\bf u})^{T}) is the symmetrized gradient of fluid velocity. Finally, on the rigid part of the boundary we prescribe the no-slip boundary conditions:

𝐮=0 on Γr.\displaystyle{\bf u}=0\quad\text{ on }\Gamma_{r}.

The structure subproblem: The elastodynamics problem is given by the linearly visco-elastic plate equations describing the displacement of the structure in three spatial directions. The plate is displaced from its reference domain Γ\Gamma by 𝜼=(ηx,ηy,ηz){\boldsymbol{\eta}}=(\eta_{x},\eta_{y},\eta_{z}), which satisfies the following equation for some 0<s10<s\leq 1, see e.g., [5, 31]:

(3) t2𝜼+Δ2𝜼+γΛ2+2st𝜼=F𝜼 in Γ,whereΛ=(Δ)12.\displaystyle\partial^{2}_{t}{\boldsymbol{\eta}}+\Delta^{2}{\boldsymbol{\eta}}+\gamma\Lambda^{2+2s}\partial_{t}{\boldsymbol{\eta}}=F_{\boldsymbol{\eta}}\quad\text{ in }\Gamma,\quad{\rm where}\ \Lambda=(-\Delta)^{\frac{1}{2}}.

Here, F𝜼F_{{\boldsymbol{\eta}}} denotes the total force experienced by the structure. Assuming that the external forcing on the structure is 0, this force F𝜼F_{\boldsymbol{\eta}} in the coupled problem results from the jump in the normal stress (traction) across the structure. With the assumption that the external force is zero, F𝜼F_{\boldsymbol{\eta}} comes entirelly from the fluid load felt by the structure (see (5)).

Since we work on the torus, the square root of negative Laplatian, denoted here by Λ\Lambda, along with its powers can be defined via Fourier transform.

The damped plate model given by equation (3) has been extensively studied in the literature, see e.g., [5, 31].

Remark 1.

In the classical work [5], s=0s=0 is identified as a critical parameter for which the semigroup, defined by the spatial differential operator, becomes analytic. In our work, we use the dissipation term to derive a priori estimates, which necessitates that s>0s>0.

We stress here that while the fluid equations are posed on time-dependent domains, in Eulerian framework, the structure equations are defined in Lagrangian coordinates on the fixed reference domain Γ\Gamma.

The non-linear fluid-structure coupling: The coupling between the structure and the fluid takes place across the ”current” location of the fluid-structure interface. We consider a two-way coupling described by the so-called kinematic and dynamic coupling conditions that describe continuity of velocity and continuity of normal stress at the fluid-structure interface, respectively.

  • The kinematic coupling condition which describes the continuity of velocities at the interface is the no-slip boundary condition, which in the case of moving boundary, reads:

    (4) t𝜼(t)=(𝐮𝐀𝜼(t))|Γ.\displaystyle\partial_{t}{\boldsymbol{\eta}}(t)=({\bf u}\circ{\bf A}_{\boldsymbol{\eta}}(t))|_{\Gamma}.
  • The dynamic coupling condition specifies the load F𝜼F_{\boldsymbol{\eta}} experienced by the structure:

    (5) F𝜼=S𝜼(t)((σ𝐧𝜼)𝐀𝜼(t))|Γ,\displaystyle F_{{\boldsymbol{\eta}}}=-S_{{\boldsymbol{\eta}}}(t)\left((\sigma{\bf n}^{{\boldsymbol{\eta}}})\circ{{\bf A}_{\boldsymbol{\eta}}}(t)\right)|_{\Gamma},

    where 𝐧𝜼{\bf n}^{{\boldsymbol{\eta}}} is the unit outward normal to the boundary of 𝒪𝜼\mathcal{O}_{\boldsymbol{\eta}}, and S𝜼=|cofA𝜼𝐞3|S_{\boldsymbol{\eta}}=|{\rm cof}\nabla A_{{\boldsymbol{\eta}}}{\bf e}_{3}| defines surface measure of Γ𝜼\Gamma_{{{\boldsymbol{\eta}}}}, i.e. dΓ𝜼=S𝜼dΓd\Gamma_{{\boldsymbol{\eta}}}=S_{{\boldsymbol{\eta}}}d\Gamma. This term arises from the transformation between Eulerian and Lagrangian coordinates.

2.1. Energy of the coupled problem

In this section we formally derive the following energy inequality corresponding to the coupled fluid-structure interaction problem:

(6) 𝐮L(0,T;𝐋2(𝒪𝜼()))+𝐃𝐮L2(0,T;𝐋2(𝒪𝜼()))+𝜼L(0,T;𝐇2(Γ))+t𝜼L2(0,T;𝐇1+s(Γ))C\boxed{\|{\bf u}\|_{L^{\infty}(0,T;{\bf L}^{2}(\mathcal{O}_{\boldsymbol{\eta}}(\cdot)))}+\|{\bf D}{\bf u}\|_{L^{2}(0,T;{\bf L}^{2}(\mathcal{O}_{\boldsymbol{\eta}}(\cdot)))}+\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}+\|\partial_{t}{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{1+s}(\Gamma))}\leq C}

where C>0C>0 depends only on the given data 𝐮0,𝐯0,𝜼0{\bf u}_{0},{\bf v}_{0},{\boldsymbol{\eta}}_{0}.

To show the derivation of this energy inequality, we begin by multiplying the fluid equations (2) with 𝐮{\bf u} and then integrate over the moving domain 𝒪𝜼\mathcal{O}_{\boldsymbol{\eta}}. For any t[0,T]t\in[0,T] we obtain,

(7) 𝒪𝜼(t)(t𝐮𝐮+(𝐮)𝐮𝐮)=𝒪𝜼(t)(σ)𝐮.\displaystyle\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}(\partial_{t}{\bf u}\cdot{\bf u}+({\bf u}\cdot\nabla){\bf u}\cdot{\bf u})=\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}(\nabla\cdot\sigma)\cdot{\bf u}.

Thanks to Reynold’s transport theorem, the first half of the left-hand side of (4.9) can be written as

𝒪𝜼(t)t𝐮𝐮=ddt𝒪𝜼(t)12|𝐮|2Γ𝜼(t)12|𝐮|2𝐮𝐧𝜼\displaystyle\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}\partial_{t}{\bf u}\cdot{\bf u}=\frac{d}{dt}\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}\frac{1}{2}|{\bf u}|^{2}-\int_{\Gamma_{\boldsymbol{\eta}}(t)}\frac{1}{2}|{\bf u}|^{2}{\bf u}\cdot{\bf n}^{\boldsymbol{\eta}}

Whereas, the advection term can be treated as follows

𝒪𝜼(t)(𝐮)𝐮𝐮\displaystyle\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}({\bf u}\cdot\nabla){\bf u}\cdot{\bf u} =12Γ𝜼|𝐮|2𝐮𝐧𝜼.\displaystyle=\frac{1}{2}\int_{\Gamma_{{{\boldsymbol{\eta}}}}}|{\bf u}|^{2}{\bf u}\cdot{\bf n}^{{\boldsymbol{\eta}}}.

For the term on the right-hand side of (4.9) we obtain

𝒪𝜼(t)(σ)𝐮\displaystyle\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}(\nabla\cdot\sigma)\cdot{\bf u} =𝒪𝜼(t)σ𝐧𝜼𝐮𝒪𝜼(t)|𝐃𝐮|2.\displaystyle=\int_{\partial\mathcal{O}_{\boldsymbol{\eta}}(t)}\sigma{\bf n}^{\boldsymbol{\eta}}\cdot{\bf u}-\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}|{\bf D}{\bf u}|^{2}.

Now, by applying the kinematic and dynamic coupling conditions at the fluid-structure interface and by using (3) and the fact that 𝜼{\boldsymbol{\eta}} is periodic in xx and yy, we obtain that

Γ𝜼(t)σ𝐧𝜼𝐮\displaystyle\int_{\Gamma_{\boldsymbol{\eta}}(t)}\sigma{\bf n}^{\boldsymbol{\eta}}\cdot{\bf u} =ΓS𝜼(t)((σ𝐧𝜼)𝐀𝜼(t)|Γ)(𝐮𝐀𝜼(t))|Γ=ΓF𝜼t𝜼\displaystyle=\int_{\Gamma}S_{{\boldsymbol{\eta}}}(t)\left((\sigma{\bf n}^{{\boldsymbol{\eta}}})\circ{{\bf A}_{\boldsymbol{\eta}}}(t)|_{\Gamma}\right)({\bf u}\circ{{\bf A}_{\boldsymbol{\eta}}}(t))|_{\Gamma}=\int_{\Gamma}F_{{\boldsymbol{\eta}}}\partial_{t}{\boldsymbol{\eta}}
=Γ(t2𝜼+Δ2𝜼γΛ2+2st𝜼)t𝜼\displaystyle=\int_{\Gamma}(\partial^{2}_{t}{\boldsymbol{\eta}}+\Delta^{2}{\boldsymbol{\eta}}-\gamma\Lambda^{2+2s}\partial_{t}{\boldsymbol{\eta}})\cdot\partial_{t}{\boldsymbol{\eta}}
=ddtΓ(|t𝜼|2+|Δ𝜼|2)+γΓ|Λ1+st𝜼|2.\displaystyle=\frac{d}{dt}\int_{\Gamma}(|\partial_{t}{\boldsymbol{\eta}}|^{2}+|\Delta{\boldsymbol{\eta}}|^{2})+\gamma\int_{\Gamma}|\Lambda^{{1+s}}\partial_{t}{\boldsymbol{\eta}}|^{2}.

Hence, gathering all the equations above we obtain:

(8) ddt𝒪𝜼(t)12|𝐮|2+𝒪𝜼(t)|𝐃𝐮|2+ddtΓ(|t𝜼|2+|Δ𝜼|2)+γΓ|Λ1+st𝜼|2.\begin{split}\frac{d}{dt}\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}\frac{1}{2}|{\bf u}|^{2}+\int_{\mathcal{O}_{\boldsymbol{\eta}}(t)}|{\bf D}{\bf u}|^{2}+\frac{d}{dt}\int_{\Gamma}(|\partial_{t}{\boldsymbol{\eta}}|^{2}&+|\Delta{\boldsymbol{\eta}}|^{2})+\gamma\int_{\Gamma}|\Lambda^{{1+s}}\partial_{t}{\boldsymbol{\eta}}|^{2}.\end{split}

Integration with respect to time implies the energy inequality (6).

2.2. Weak formulation on moving domains

Before we derive the weak formulation of the deterministic system described in the previous subsection, we define the following function spaces for the fluid velocity, the structure, and the coupled FSI problem:

𝒱~F(t)={𝐮𝐇1(𝒪𝜼(t)):𝐮=0,𝐮=0 on Γr},\displaystyle\tilde{\mathscr{V}}_{F}(t)=\{{\bf u}\in{\bf H}^{1}(\mathcal{O}_{{\boldsymbol{\eta}}}(t)):\nabla\cdot{\bf u}=0,{\bf u}=0\text{ on }\Gamma_{r}\},
𝒲~F(0,T)=L(0,T;𝐋2(𝒪𝜼()))L2(0,T;𝒱~F()),\displaystyle\tilde{\mathscr{W}}_{F}(0,T)=L^{\infty}(0,T;{\bf L}^{2}(\mathcal{O}_{\boldsymbol{\eta}}(\cdot)))\cap L^{2}(0,T;\tilde{\mathscr{V}}_{F}(\cdot)),
𝒱S=𝐇2(Γ)\displaystyle\mathscr{V}_{S}={\bf H}^{2}(\Gamma)
𝒲S(0,T)=W1,(0,T;𝐋2(Γ))L(0,T;𝒱S)H1(0,T;𝐇1+s(Γ));0<s1,\displaystyle\mathscr{W}_{S}(0,T)=W^{1,\infty}(0,T;{\bf L}^{2}(\Gamma))\cap L^{\infty}(0,T;\mathscr{V}_{S})\cap H^{1}(0,T;{\bf H}^{1+{s}}(\Gamma));\quad 0<s\leq 1,
𝒲~(0,T)={(𝐮,𝜼)𝒲~F(0,T)×𝒲S(0,T):t𝜼(t)=𝐮𝐀𝜼(t) on Γ}.\displaystyle\tilde{\mathscr{W}}(0,T)=\{({\bf u},{\boldsymbol{\eta}})\in\tilde{\mathscr{W}}_{F}(0,T)\times\mathscr{W}_{S}(0,T):\partial_{t}{\boldsymbol{\eta}}(t)={\bf u}\circ{\bf A}_{\boldsymbol{\eta}}(t)\text{ on }\Gamma\}.

Here bold-faced lettered spaces are used for vector valued functions. We will take test functions (𝐪,𝝍)({\bf q},\boldsymbol{\psi}) from the following space:

𝒟~(0,T)={(𝐪,𝝍)C1([0,T);𝒱~F()×𝒱S):𝝍(t)=𝐪𝐀𝜼(t), on Γ}.\tilde{\mathscr{D}}(0,T)=\{({\bf q},\boldsymbol{\psi})\in C^{1}([0,T);\tilde{\mathscr{V}}_{F}(\cdot)\times\mathscr{V}_{S}):\boldsymbol{\psi}(t)={\bf q}\circ{\bf A}_{\boldsymbol{\eta}}(t),\text{ on }\Gamma\}.

Now, we can introduce the weak formulation of our problem on moving domain.

Definition 1.

We say that (𝐮,𝛈)𝒲~(0,T)({\bf u},{\boldsymbol{\eta}})\in\tilde{\mathscr{W}}(0,T), is a weak solution to (2)-(5) if for any test function 𝐐=(𝐪,𝛙)𝒟~(0,T){\bf Q}=({\bf q},\boldsymbol{\psi})\in\tilde{\mathscr{D}}(0,T) the following equality holds:

(9) 0T𝒪𝜼(t)𝐮t𝐪0TΓt𝜼t𝝍0TΓ𝜼(t)(𝐮𝐪)(𝐮𝐧𝜼)+0T𝒪𝜼(t)((𝐮)𝐮𝐪)+2ν0T𝒪𝜼(t)𝐃(𝐮)𝐃(𝐪)+0TΓΔ𝜼Δ𝝍+γ0TΓΛ1+st𝜼:Λ1+s𝝍=𝒪𝜼0𝐮0𝐪(0)+Γ𝐯0𝝍(0).\begin{split}&-\int_{0}^{T}\int_{\mathcal{O}_{{{\boldsymbol{\eta}}}(t)}}{\bf u}\cdot\partial_{t}{\bf q}-\int_{0}^{T}\int_{\Gamma}\partial_{t}{{\boldsymbol{\eta}}}\partial_{t}\boldsymbol{\psi}-\int_{0}^{T}\int_{\Gamma_{{\boldsymbol{\eta}}}(t)}({\bf u}\cdot{\bf q})({\bf u}\cdot{\bf n}^{{\boldsymbol{\eta}}})\\ &+\int_{0}^{T}\int_{\mathcal{O}_{{{\boldsymbol{\eta}}}}(t)}\left(({\bf u}\cdot\nabla){\bf u}\cdot{{\bf q}}\right)+2\nu\int_{0}^{T}\int_{\mathcal{O}_{{{\boldsymbol{\eta}}}(t)}}{\bf D}({\bf u})\cdot{\bf D}({\bf q})\\ &+\int_{0}^{T}\int_{\Gamma}\Delta{\boldsymbol{\eta}}\cdot\Delta\boldsymbol{\psi}+\gamma\int_{0}^{T}\int_{\Gamma}{\Lambda^{{1+s}}\partial_{t}{\boldsymbol{\eta}}:\Lambda^{{1+s}}\boldsymbol{\psi}}=\int_{\mathcal{O}_{{\boldsymbol{\eta}}_{0}}}{\bf u}_{0}{\bf q}(0)+\int_{\Gamma}{\bf v}_{0}\boldsymbol{\psi}(0).\end{split}
Remark 2.

Notice that under the assumption that there exists of a family of C1C^{1} diffeomorphisms 𝐀𝛈{\bf A}_{\boldsymbol{\eta}}, defined in (1), this weak formulation is well-defined. Along with the weak solutions, we will construct the corresponding maps 𝐀𝛈{\bf A}_{\boldsymbol{\eta}} satisfying these assumptions.

We recall that one of our main results in this manuscript is the existence of a solution to the FSI problem (2)-(5) in the sense of Definition 1.

2.3. Arbitrary Lagrangian-Eulerian (ALE) formulation on fixed domain

To deal with the geometric non-linearities resulting from the motion of the fluid domain we transform the fluid equations onto the fixed reference domain 𝒪=Γ×(0,1)\mathcal{O}=\Gamma\times(0,1) and give a weak formulation equivalent to (LABEL:origweakform) posed on this fixed domain. For that purpose, we consider a family of Arbitrary Lagrangian-Eulerian (ALE) mappings that are ubiquitous in the field of computational fluid-structure interaction. The ALE maps, denoted by 𝐀𝜼{\bf A}_{\boldsymbol{\eta}}, constitute a family, parametrized by time t[0,T]t\in[0,T], of diffeomorphisms from the fixed domain 𝒪\mathcal{O} onto the moving domain 𝒪𝜼(t)\mathcal{O}_{{{\boldsymbol{\eta}}}}(t). With the aid of these maps, we will find a relevant weak formulation on 𝒪\mathcal{O} satisfied by 𝐮𝐀𝜼{\bf u}\circ{\bf A}_{\boldsymbol{\eta}}.

In this article, these maps will be obtained by considering harmonic extensions of the structure displacement 𝜼{\boldsymbol{\eta}} in 𝒪\mathcal{O}. That is, the ALE maps solve the following equations:

(10) Δ𝐀𝜼=0, in 𝒪,𝐀𝜼=id+𝜼 on Γ, and 𝐀𝜼=id on 𝒪Γ.\begin{split}\Delta{\bf A}_{\boldsymbol{\eta}}&=0,\quad\text{ in }\mathcal{O},\\ {\bf A}_{\boldsymbol{\eta}}=\textbf{id}+{{\boldsymbol{\eta}}}\text{ on }\Gamma,&\quad\text{ and }\quad{\bf A}_{\boldsymbol{\eta}}=\textbf{id}\text{ on }\partial\mathcal{O}\setminus\Gamma.\end{split}

The existence and uniqueness of the solution to (10) is classical. However, we need to prove that these maps are well-defined, i.e. 𝐀𝜼(t):𝒪𝒪𝜼(t){\bf A}_{\boldsymbol{\eta}}(t):\mathcal{O}\to\mathcal{O}_{\boldsymbol{\eta}}(t) is indeed a C1C^{1}-diffeomorphism for every t[0,T]t\in[0,T] (see Remark 2.1).

Before moving on to analyzing the properties of 𝐀𝜼{\bf A}_{\boldsymbol{\eta}}, we summarize the notation that will be used to simplify the ALE formulation of the problem. First, we denote the Jacobian of the ALE maps by

(11) J𝜼=det 𝐀η.J_{{\boldsymbol{\eta}}}=\text{det }\nabla{\bf A}_{\eta}.

Next, under the transformation given in (10), the transformed gradient and the transformed symmetrized gradient of any function 𝐠𝜼:=𝐠𝐀𝜼{\bf g}^{\boldsymbol{\eta}}:={\bf g}\circ{\bf A}_{\boldsymbol{\eta}} for 𝐠𝐇1(𝒪𝜼){\bf g}\in{\bf H}^{1}(\mathcal{O}_{\boldsymbol{\eta}}) are given by

𝜼𝐠𝜼=𝐠𝐀𝜼=𝐠𝜼(𝐀𝜼)1 and 𝐃η(𝐮)=12(η𝐮+(η)T𝐮).\nabla^{{\boldsymbol{\eta}}}{\bf g}^{{\boldsymbol{\eta}}}=\nabla{\bf g}\circ{\bf A}_{\boldsymbol{\eta}}=\nabla{\bf g}^{{\boldsymbol{\eta}}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}\quad\text{ and }\quad{\bf D}^{\eta}({\bf u})=\frac{1}{2}(\nabla^{\eta}{\bf u}+(\nabla^{\eta})^{T}{\bf u}).

Similarly, the transformed divergence will be denoted by

div𝜼𝐮=𝜼𝐮=tr(𝜼𝐮).\text{div}^{{\boldsymbol{\eta}}}{\bf u}=\nabla^{{\boldsymbol{\eta}}}\cdot{\bf u}=tr(\nabla^{\boldsymbol{\eta}}{\bf u}).

Finally, we use 𝐰η{\bf w}^{\eta} to denote the ALE velocity:

𝐰η=t𝐀𝜼.{\bf w}^{\eta}=\partial_{t}{\bf A}_{\boldsymbol{\eta}}.

Using this notation we will give the definition of function spaces used to describe the fixed domain ALE formulation of our FSI problem. We define,

𝒱F𝜼(t)={𝐮𝐇1(𝒪):𝜼𝐮=0,𝐮=0 on Γr},\displaystyle\mathscr{V}^{{\boldsymbol{\eta}}}_{F}(t)=\{{\bf u}\in{\bf H}^{1}(\mathcal{O}):\nabla^{\boldsymbol{\eta}}\cdot{\bf u}=0,{\bf u}=0\text{ on }\Gamma_{r}\},
𝒲F(0,T)=L(0,T;𝐋2(𝒪))L2(0,T;𝒱F𝜼()),\displaystyle\mathscr{W}_{F}(0,T)=L^{\infty}(0,T;{\bf L}^{2}(\mathcal{O}))\cap L^{2}(0,T;\mathscr{V}^{{\boldsymbol{\eta}}}_{F}(\cdot)),
𝒲(0,T)={(𝐮,𝜼)𝒲F(0,T)×𝒲S(0,T):t𝜼(t)=𝐮|Γ}.\displaystyle\mathscr{W}(0,T)=\{({\bf u},{\boldsymbol{\eta}})\in\mathscr{W}_{F}(0,T)\times\mathscr{W}_{S}(0,T):\partial_{t}{\boldsymbol{\eta}}(t)={\bf u}|_{\Gamma}\}.

The space of test functions is as follows:

𝒟𝜼(0,T)={(𝐪,𝝍)C1([0,T);𝒱F𝜼()H3(𝒪)×𝒱S):𝐪|Γ=𝝍}.\mathscr{D}^{\boldsymbol{\eta}}(0,T)=\{({{\bf q}},\boldsymbol{\psi})\in C^{1}([0,T);\mathscr{V}^{{\boldsymbol{\eta}}}_{F}(\cdot)\cap H^{3}(\mathcal{O})\times\mathscr{V}_{S}):{\bf q}|_{\Gamma}=\boldsymbol{\psi}\}.

Now we will now present a weak formulation on the fixed domain 𝒪\mathcal{O}, derivation of which is the same as given in Section 4.3 [22].

Definition 2.

We say that (𝐮,𝛈)𝒲(0,T)({\bf u},{\boldsymbol{\eta}})\in\mathscr{W}(0,T) is a weak solution of the nonlinearly coupled FSI problem (2)-(5) defined in terms of a fixed domain formulation on 𝒪\mathcal{O} if the following equation holds for any (𝐪,𝛙)𝒟𝛈(0,T)({\bf q},\boldsymbol{\psi})\in\mathscr{D}^{\boldsymbol{\eta}}(0,T):

(12) 0T𝒪J𝜼𝐮t𝐪0TΓt𝜼t𝝍=0T𝒪tJ𝜼𝐮𝐪0T𝒪J𝜼(𝐮𝜼𝐮𝐪𝐰𝜼𝜼𝐮𝐪)2ν0T𝒪J𝜼𝐃𝜼(𝐮):𝐃𝜼(𝐪)+0TΓΔ𝜼Δ𝝍+γ0TΓΛ1+st𝜼:Λ1+s𝝍+𝒪J𝜼0𝐮0𝐪(0)+Γ𝐯0𝝍(0).\begin{split}&-\int_{0}^{T}\int_{\mathcal{O}}{J}_{{\boldsymbol{\eta}}}{\bf u}\cdot\partial_{t}{{\bf q}}-\int_{0}^{T}\int_{\Gamma}\partial_{t}{{\boldsymbol{\eta}}}\,\partial_{t}{\boldsymbol{\psi}}=\int_{0}^{T}\int_{\mathcal{O}}\partial_{t}J_{\boldsymbol{\eta}}{\bf u}\cdot{\bf q}\\ &-\int_{0}^{T}\int_{\mathcal{O}}J_{{\boldsymbol{\eta}}}({\bf u}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q}-{\bf w}^{{\boldsymbol{\eta}}}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q})-2\nu\int_{0}^{T}\int_{\mathcal{O}}{J}_{{\boldsymbol{\eta}}}\,{\bf D}^{{{\boldsymbol{\eta}}}}({\bf u}):{\bf D}^{{{\boldsymbol{\eta}}}}({\bf q})\\ &+\int_{0}^{T}\int_{\Gamma}\Delta{\boldsymbol{\eta}}\cdot\Delta\boldsymbol{\psi}+\gamma\int_{0}^{T}\int_{\Gamma}\Lambda^{{1+s}}\partial_{t}{\boldsymbol{\eta}}:\Lambda^{{1+s}}\boldsymbol{\psi}+\int_{\mathcal{O}}J_{{\boldsymbol{\eta}}_{0}}{\bf u}_{0}{\bf q}(0)+\int_{\Gamma}{\bf v}_{0}\boldsymbol{\psi}(0).\end{split}
Remark 3.

We note that if the ALE map 𝐀𝛈(t):𝒪𝒪𝛈(t){\bf A}_{\boldsymbol{\eta}}(t):\mathcal{O}\mapsto\mathcal{O}_{\boldsymbol{\eta}}(t), defined as the solution to (10), is Lipschitz continuous and bijective, then Definitions 1 and 2 are equivalent. In other words, (𝐮,𝛈)({\bf u},{\boldsymbol{\eta}}) solves (LABEL:weaksol) iff (𝐮~,𝛈)(\tilde{\bf u},{\boldsymbol{\eta}}) where 𝐮~=𝐮𝐀𝛈1\tilde{\bf u}={\bf u}\circ{\bf A}^{-1}_{\boldsymbol{\eta}} solves (LABEL:origweakform) i.e. it is the desired weak solution of our FSI problem in the sense of Definition 1.

Remark 4.

(Notation) Throughout the rest of the manuscript we will be using 𝐮~\tilde{\bf u}, where

𝐮~=𝐮𝐀𝜼1,\tilde{\bf u}={\bf u}\circ{\bf A}^{-1}_{\boldsymbol{\eta}},

to denote the fluid velocity defined on the moving domain 𝒪𝛈\mathcal{O}_{\boldsymbol{\eta}}, to distinguish between the solution 𝐮{\bf u} defined on the fixed domain 𝒪\mathcal{O}, and the solution 𝐮~\tilde{\bf u} defined on the moving domain 𝒪𝛈\mathcal{O}_{\boldsymbol{\eta}}.

Next, we discuss conditions that are sufficient to imply bijectivity of the ALE maps 𝐀𝜼{\bf A}_{\boldsymbol{\eta}}. First, observe that for any k0k\geq 0 the solution to (10) satisfies (see e.g. [15]):

(13) 𝐀𝜼𝐇k+12(𝒪)𝜼𝐇k(Γ).\displaystyle\|{\bf A}_{\boldsymbol{\eta}}\|_{{\bf H}^{k+\frac{1}{2}}(\mathcal{O})}\leq\|{\boldsymbol{\eta}}\|_{{\bf H}^{k}(\Gamma)}.

Observe also that, for any p2p\geq 2, we have the following regularity result for the harmonic extension 𝐀𝜼{\bf A}_{\boldsymbol{\eta}} of the boundary data 𝐢𝐝+𝜼{\bf id}+{\boldsymbol{\eta}} thanks to the discussion presented in Section 5 in [15]:

(14) 𝐀𝜼𝐢𝐝𝐖2,p(𝒪)C𝜼𝐖21p,p(Γ)C𝜼𝐇33p(Γ).\displaystyle\|{\bf A}_{\boldsymbol{\eta}}-{\bf id}\|_{{\bf W}^{2,p}(\mathcal{O})}\leq C\|{{\boldsymbol{\eta}}}\|_{{\bf W}^{2-\frac{1}{p},p}(\Gamma)}\leq C\|{{\boldsymbol{\eta}}}\|_{{\bf H}^{3-\frac{3}{p}}(\Gamma)}.

Hence, Morrey’s inequality (see e.g. Theorem 7.26 in [10]) implies that for some Cp>0C^{*}_{p}>0 the following inequality holds true for p>3p>3,

(15) (𝐀𝜼𝐢𝐝)𝐂0,13p(𝒪¯)(𝐀𝜼𝐢𝐝)𝐖1,p(𝒪¯)Cp𝜼𝐇33p(Γ).\displaystyle\|\nabla({\bf A}_{\boldsymbol{\eta}}-{\bf id})\|_{{\bf C}^{0,1-\frac{3}{p}}(\bar{\mathcal{O}})}\leq\|\nabla({\bf A}_{\boldsymbol{\eta}}-{\bf id})\|_{{\bf W}^{1,p}(\bar{\mathcal{O}})}\leq C^{*}_{p}\|{{\boldsymbol{\eta}}}\|_{{\bf H}^{3-\frac{3}{p}}(\Gamma)}.

Now thanks to Theorem 5.5-1 (B) of [6], for as long as the structure displacement 𝜼{\boldsymbol{\eta}} satisfies

(16) 𝜼𝐇33p(Γ)1Cp,for any p>3,\displaystyle\|{\boldsymbol{\eta}}\|_{{\bf H}^{3-\frac{3}{p}}(\Gamma)}\leq\frac{1}{C^{*}_{p}},\qquad\text{for any }p>3,

the map 𝐀𝜼𝐂1,13p(𝒪¯){\bf A}_{\boldsymbol{\eta}}\in{\bf C}^{1,1-\frac{3}{p}}(\bar{\mathcal{O}}) is injective. Thanks to invariance of domains (see [6]), we infer that 𝐀𝜼{\bf A}_{\boldsymbol{\eta}} is thus a bijection between the domains 𝒪\mathcal{O} and 𝒪𝜼\mathcal{O}_{\boldsymbol{\eta}}.

We have established the following

Proposition 2.1.

If for any δ>0\delta>0, 𝛈L(0,T;𝐇2+δ(Γ)){\boldsymbol{\eta}}\in L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma)) then for some small enough T0>0T_{0}>0, the map 𝐀𝛈L(0,T0;𝐂1,δ(𝒪¯)){\bf A}_{\boldsymbol{\eta}}\in L^{\infty}(0,T_{0};{\bf C}^{1,\delta}(\bar{\mathcal{O}})) solving (10) is bijective and Definitions 1 and 2 are thus equivalent.

2.4. Main results

We are now in a position to state the main results of this article.

In the first two theorems we will state the enhanced spatial regularity of the structure and the enhanced temporal regularity of the fluid and structure velocities.

Before stating these results we recall the definition of Nikolski spaces. Let the translation in time by hh of a function ff be denoted by:

τhf(t,)=f(th,),h.\tau_{h}f(t,\cdot)=f(t-h,\cdot),\quad h\in\mathbb{R}.

Let 𝒴\mathcal{Y} be a Banach space. Then, for any 0<m<10<m<1 and 1p<1\leq p<\infty, the Nikolski space is defined as:

(17) Nm,p(0,T;𝒴)={𝐮Lp(0,T;𝒴):sup0<h<T1hmτh𝐮𝐮Lp(h,T;𝒴)<}.{N^{m,p}(0,T;\mathcal{Y})}=\{{\bf u}\in L^{p}(0,T;\mathcal{Y}):\sup_{0<h<T}\frac{1}{h^{m}}\|\tau_{h}{\bf u}-{\bf u}\|_{L^{p}(h,T;\mathcal{Y})}<\infty\}.

Now we state our a priori estimates that provide additional regularity for the structure displacement and the fluid velocity. These are our first two main results of the manuscript.

Theorem 1.

Let (𝐮~,𝛈)(\tilde{\bf u},{\boldsymbol{\eta}}) be a smooth solution to the FSI problem defined on moving domains, satisfying (LABEL:origweakform). Then the following a priori estimates addressing spatial regularity hold true:

  1. (1)

    The structure displacement 𝜼{\boldsymbol{\eta}} satisfies:

    (18) 𝜼L(0,T0;𝐇2+δ(Γ))+𝜼L2(0,T0;𝐇3(sδ)(Γ))<C,for any  0<δ<s.\displaystyle\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma))}+\|{\boldsymbol{\eta}}\|_{L^{2}(0,T_{0};{\bf H}^{3-(s-\delta)}(\Gamma))}<C,\qquad\text{for any }\,0<\delta<s.
  2. (2)

    Moreover, the ALE maps defined by (10) satisfy

    (19) 𝐀𝜼L(0,T0;𝐂1,δ(𝒪¯))<C.\displaystyle\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf C}^{1,\delta}(\bar{\mathcal{O}}))}<C.

Here, CC depends only on the energy norm of the initial data, as well as the H2+sH^{2+s}-norm of the initial displacement 𝛈0{\boldsymbol{\eta}}_{0}, and on the viscoelasticity coefficient γ>0\gamma>0 and on domain 𝒪\mathcal{O}.

Theorem 2.

Let (𝐮~,𝛈)(\tilde{\bf u},{\boldsymbol{\eta}}) be a smooth solution to the FSI problem defined on moving domains, satisfying (LABEL:origweakform). Then the fluid and structure velocities (𝐮~,t𝛈)(\tilde{\bf u},\partial_{t}{\boldsymbol{\eta}}) satisfy the following a priori estimate addressing temporal regularity property:

(20) 𝐮~𝐀𝜼N18,2(0,T0;𝐋2(𝒪))+t𝜼N18,2(0,T0;𝐋2(Γ))<C.\displaystyle\|\tilde{\bf u}\circ{\bf A}_{\boldsymbol{\eta}}\|_{N^{\frac{1}{8},2}(0,T_{0};{\bf L}^{2}(\mathcal{O}))}+\|\partial_{t}{\boldsymbol{\eta}}\|_{N^{\frac{1}{8},2}(0,T_{0};{\bf L}^{2}(\Gamma))}<C.

Here, CC depends only on the energy norm of the initial data, as well as the H2+sH^{2+s}-norm of the initial displacement 𝛈0{\boldsymbol{\eta}}_{0}, and on domain 𝒪\mathcal{O}.

Our third main result of the manuscript is the existence of a weak solution to the nonlinearly coupled problem, as stated in the following theorem.

Theorem 3.

Let the initial data for structure displacement, structure velocity and fluid velocity be such that 𝛈0𝐇2+s(Γ),𝐯0𝐋2(Γ){\boldsymbol{\eta}}_{0}\in{\bf H}^{2+s}(\Gamma),{\bf v}_{0}\in{\bf L}^{2}(\Gamma) and 𝐮0𝐋2(𝒪𝛈0){\bf u}_{0}\in{\bf L}^{2}(\mathcal{O}_{{\boldsymbol{\eta}}_{0}}). Then there exists T0>0T_{0}>0 and at least one weak solution to the system (2)-(5) on [0,T0][0,T_{0}] in the sense of Definition 1.

In what follows, we will give the proofs of these two theorems. We will start, in Section 3, with the proofs Theorems 1 and 2, and then use these regularity results in Section 4 to construct a weak solution for (2)-(5), thus proving Theorem 3. Specifically, Theorem 1, which states that at any time the structure displacement is Lipschitz continuous in space, is crucial in obtaining a positive time-length during which the fluid domain remains non-degenerate and thus in transforming the fluid equations onto the fixed domain 𝒪\mathcal{O}. It is also used in the construction of the Bogovski-type operator constructed in the proof of Theorem 2. Theorem 2 is used in Section 4 to obtain compactness of the sequence of approximate solutions to prove the existence of a solution to the FSI problem in the sense of Definition 2.

3. Regularity results

3.1. The structure regularity result.

In this section we will prove Theorem 1 showing the a priori regularity result for 𝜼{\boldsymbol{\eta}}. To establish this result we work in the fixed domain setting of Definition 2 and operate under the assumption that 𝜼{\boldsymbol{\eta}} is smooth and that the map 𝐀𝜼(t):𝒪𝒪𝜼(t){\bf A}_{\boldsymbol{\eta}}(t):\mathcal{O}\mapsto\mathcal{O}_{\boldsymbol{\eta}}(t), solving (10) is bijective.

In this case, we make note of the following result that gives us the equivalent of the energy estimate (6) for the fixed domain counterparts.

Lemma 3.1.

Let (𝐮,𝛈)({\bf u},{\boldsymbol{\eta}}) be a weak solution in the sense of the Definition 2 on the fixed domain 𝒪\mathcal{O}. Assume that the ALE maps 𝐀𝛈(t):𝒪𝒪𝛈(t){\bf A}_{\boldsymbol{\eta}}(t):\mathcal{O}\mapsto\mathcal{O}_{\boldsymbol{\eta}}(t), solving (10), are bijective and that for some α>0\alpha>0 their Jacobians satisfy inf𝒪J𝛈>α>0\inf_{\mathcal{O}}J_{\boldsymbol{\eta}}>\alpha>0 for all t[0,T]t\in[0,T]. Then, for some constant K1>0K_{1}>0 depending only on 𝐀𝛈L(0,T;𝐖1,(𝒪)),(𝐀𝛈)1L(0,T;𝐖1,(𝒪))\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))},\|({\bf A}_{\boldsymbol{\eta}})^{-1}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))} and α\alpha we have,

𝐮L(0,T;𝐋2(𝒪))L2(0,T;𝒱F𝜼)<K1.\|{\bf u}\|_{L^{\infty}(0,T;{\bf L}^{2}(\mathcal{O}))\cap L^{2}(0,T;\mathscr{V}_{F}^{{\boldsymbol{\eta}}})}<K_{1}.
Proof of Lemma 3.1.

This Lemma is a consequence of the energy estimate (6). Owing to the assumption that 𝐀𝜼(t):𝒪𝒪𝜼(t){\bf A}_{\boldsymbol{\eta}}(t):\mathcal{O}\mapsto\mathcal{O}_{\boldsymbol{\eta}}(t) is bijective, we can write

𝐮=𝐮~𝐀𝜼,{\bf u}=\tilde{\bf u}\circ{\bf A}_{\boldsymbol{\eta}},

where (𝐮~,𝜼)(\tilde{\bf u},{\boldsymbol{\eta}}) is a solution to the FSI problem in the sense of Definition 1. Then the energy estimate (6) gives us that

supt[0,T]𝒪𝜼|𝐮~|2+0T𝒪𝜼|𝐃(𝐮~)|2C,\sup_{t\in[0,T]}\int_{\mathcal{O}_{\boldsymbol{\eta}}}|\tilde{\bf u}|^{2}+\int_{0}^{T}\int_{\mathcal{O}_{\boldsymbol{\eta}}}|{\bf D}(\tilde{\bf u})|^{2}\leq C,

where CC depends only on the given initial data.

We will use these bounds to obtain the desired estimates for 𝐮{\bf u}. The first bounds are obtained easily by a change of variables as follows,

αsup0tT𝒪|𝐮|2sup0tT𝒪J𝜼|𝐮|2=sup0tT𝒪𝜼|𝐮~|2C.\displaystyle\alpha\sup_{0\leq t\leq T}\int_{\mathcal{O}}|{\bf u}|^{2}\leq\sup_{0\leq t\leq T}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}|{\bf u}|^{2}=\sup_{0\leq t\leq T}\int_{\mathcal{O}_{\boldsymbol{\eta}}}|\tilde{\bf u}|^{2}\leq C.

Next, to bound 𝐮{\bf u} in L2(0,T;𝒱F𝜼)L^{2}(0,T;\mathscr{V}_{F}^{{\boldsymbol{\eta}}}), we must first establish a connection between the gradient and the symmetrized gradient of 𝐮~\tilde{{\bf u}} which is traditionally done with the aid of Korn’s inequality. However, due to our setting that involves time-varying fluid domains, we appeal to Lemma 1 in [36] that gives the existence of a universal Korn constant K>0K>0 which depends only on the reference domain 𝒪\mathcal{O} and the quantities 𝐀𝜼L(0,T;𝐖1,(𝒪)),(𝐀𝜼)1L(0,T;𝐖1,(𝒪))\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))},\|({\bf A}_{\boldsymbol{\eta}})^{-1}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}. For this constant we have that

𝐮~𝐇1(𝒪𝜼)K𝐃(𝐮~)𝐋2(𝒪𝜼).\displaystyle\|\tilde{\bf u}\|_{{\bf H}^{1}(\mathcal{O}_{\boldsymbol{\eta}})}\leq K\|{\bf D}(\tilde{\bf u})\|_{{\bf L}^{2}(\mathcal{O}_{\boldsymbol{\eta}})}.

These bounds do not immediately translate to desired Lt2Hx1L^{2}_{t}H^{1}_{x}-bounds for 𝐮{\bf u}. We observe that on the fixed 𝒪\mathcal{O} we have the following relation between the gradient and the transformed gradient (via ALE maps) of 𝐮{\bf u}:

𝐮=𝜼𝐮𝐀𝜼.\nabla{\bf u}=\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot\nabla{\bf A}_{\boldsymbol{\eta}}.

Hence, we write,

α0T\displaystyle\alpha\int_{0}^{T} 𝒪|𝐮|2𝑑x0T𝒪J𝜼|𝐮|2𝑑x=0T𝒪J𝜼|𝜼𝐮𝐀𝜼|2𝑑x\displaystyle\int_{\mathcal{O}}|\nabla{\bf u}|^{2}dx\leq\int_{0}^{T}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}|\nabla{\bf u}|^{2}dx=\int_{0}^{T}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}|\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot\nabla{\bf A}_{\boldsymbol{\eta}}|^{2}dx
𝐀𝜼L(0,T;𝐖1,(𝒪))20T𝒪J𝜼|𝜼𝐮|2𝑑x=𝐀𝜼L(0,T;𝐖1,(𝒪))20T𝒪𝜼|𝐮~|2𝑑x\displaystyle\leq\|{\bf A}_{\boldsymbol{\eta}}\|^{2}_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}\int_{0}^{T}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}|\nabla^{{\boldsymbol{\eta}}}{\bf u}|^{2}dx=\|{\bf A}_{\boldsymbol{\eta}}\|^{2}_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}\int_{0}^{T}\int_{\mathcal{O}_{\boldsymbol{\eta}}}|\nabla\tilde{\bf u}|^{2}dx
K𝐀𝜼L(0,T;𝐖1,(𝒪))20T𝒪𝜼|𝐃(𝐮~)|2𝑑x\displaystyle\leq K\|{\bf A}_{\boldsymbol{\eta}}\|^{2}_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}\int_{0}^{T}\int_{\mathcal{O}_{\boldsymbol{\eta}}}|{\bf D}(\tilde{\bf u})|^{2}dx
K1(𝐀𝜼L(0,T;𝐖1,(𝒪)),(𝐀𝜼)1L(0,T;𝐖1,(𝒪))).\displaystyle\leq K_{1}(\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))},\|({\bf A}_{\boldsymbol{\eta}})^{-1}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}).

This completes the proof of Lemma 3.1. ∎

Now, we proceed with the proof of Theorem 1. We will assume that the setting of Lemma 3.1 holds true. The main idea behind the proof of Theorem 1, namely obtaining estimate (18), is to consider the ”transformed” weak formulation (2) and use for a test function 𝝍\boldsymbol{\psi} the function 𝝍Λ2κ𝜼\boldsymbol{\psi}\sim\Lambda^{2\kappa}{\boldsymbol{\eta}} for 1s<κ<11-s<\kappa<1. In fact, to obtain precisely (18), we take

(21) κ=1(sδ),for any 0<δ<s.\kappa=1-(s-\delta),\qquad\text{for any $0<\delta<s$}.

Due to the kinematic coupling condition embedded in the test space 𝒟𝜼(0,T)\mathscr{D}^{\boldsymbol{\eta}}(0,T), we will also construct a transformed-divergence (divη)-free extension 𝐪{\bf q} of 𝝍\boldsymbol{\psi} to be used as the fluid test function. In the setting of [21], where tangential interactions between the fluid and the structure are negligible, this extension, in the case of flat reference geometry, is obtained simply by extending the boundary data onto the moving domain by a constant in the direction normal to Γ\Gamma and then composing it with the ALE map. However, constructing an appropriate extension in our setting is not easy. Firstly, Λ2κ𝜼\Lambda^{2\kappa}{\boldsymbol{\eta}} is not guaranteed to have a solenoidal extension in the moving fluid domain 𝒪𝜼\mathcal{O}_{\boldsymbol{\eta}} and thus special care has to be taken in the construction of 𝝍\boldsymbol{\psi} to ensure that it possesses an extension 𝐪{\bf q} in 𝒪\mathcal{O} which is divergence free in terms of the transformed-divergence operator (divη). Secondly, 𝐪{\bf q} and 𝝍\boldsymbol{\psi}, as a pair of test functions for (2), must satisfy appropriate bounds.

Now, due to its complicated form, instead of looking for a transformed-divergence-free extension 𝐪{\bf q} of the function 𝝍\boldsymbol{\psi} directly, we will first find a solenoidal extension of a modification of 𝝍\boldsymbol{\psi}, denoted by 𝝋\boldsymbol{\varphi}, on 𝒪\mathcal{O}, and then transform this function appropriately to obtain the desired 𝐪{\bf q}. That is, we will find a function 𝝋\boldsymbol{\varphi} such that it satisfies div𝝋=0\boldsymbol{\varphi}=0 on 𝒪\mathcal{O} and then we will transform 𝝋\boldsymbol{\varphi} into 𝐪{\bf q} in a way that guarantees that div𝐪𝜼=0{}^{\boldsymbol{\eta}}{\bf q}=0. This transformation will be obtained by multiplying 𝝋\boldsymbol{\varphi} with the inverse of the cofactor matrix of 𝐀𝜼{\bf A}_{\boldsymbol{\eta}} and using the Piola identity (see Theorem 1.7-1 in [6]) to obtain (see (28)):

(22) 𝐪\displaystyle{\bf q} =J𝜼1(𝐀𝜼)𝝋.\displaystyle=J^{-1}_{{\boldsymbol{\eta}}}(\nabla{\bf A}_{\boldsymbol{\eta}})\boldsymbol{\varphi}.

At this point we only have 𝐪{\bf q} written in terms of 𝝋\boldsymbol{\varphi}, but we still do not have 𝝋\boldsymbol{\varphi} defined, and we still do not have 𝝍\boldsymbol{\psi}. Next, we work on constructing the test function 𝝍\boldsymbol{\psi} that ”behaves” like Λ2κ𝜼\Lambda^{2\kappa}{\boldsymbol{\eta}} and satisfies the kinematic coupling condition with 𝐪{\bf q}, and has the additional property that its appropriate modification has a divergence-free extension 𝝋\boldsymbol{\varphi} in 𝒪\mathcal{O}.

Naturally, this modification must account for the transformation of 𝝋\boldsymbol{\varphi} into 𝐪{\bf q} as given in (22). Hence, we define

(23) 𝝍:=Λ2κ𝜼c𝝃\displaystyle\boldsymbol{\psi}:=\Lambda^{2\kappa}{\boldsymbol{\eta}}-c\boldsymbol{\xi}

where c𝝃c\boldsymbol{\xi} is a correction term that allows us to transform 𝝍\boldsymbol{\psi} so that its transformation possesses a divergence free extension in 𝒪\mathcal{O}. More precisely, we let

(24) c=Γ𝜼×Λ2κ𝜼Γ𝜼×𝝃,\displaystyle{c}=\frac{\int_{\Gamma}\nabla{\boldsymbol{\eta}}\times\Lambda^{2\kappa}{\boldsymbol{\eta}}}{\int_{\Gamma}\nabla{\boldsymbol{\eta}}\times\boldsymbol{\xi}},

where 𝝃𝐂0([0,T]×Γ)\boldsymbol{\xi}\in{\bf C}^{\infty}_{0}([0,T]\times\Gamma) is such that the denominator in the definition of the constant cc is non-zero. In fact, we choose 𝝃\boldsymbol{\xi} such that 𝝃(t)𝐂2(Γ)=1\|\boldsymbol{\xi}(t)\|_{{\bf C}^{2}(\Gamma)}=1 and 𝜼×𝝃(t)=1\nabla{\boldsymbol{\eta}}\times\boldsymbol{\xi}(t)=1 for every t[0,T]t\in[0,T]. Note that for this choice of 𝝃\boldsymbol{\xi} we have

sup0tT|c(t)|sup0tT𝐧𝜼(t)𝐋2(Γ)Λ2κ𝜼(t)𝐋2(Γ).\displaystyle\sup_{0\leq t\leq T}|c(t)|\leq\sup_{0\leq t\leq T}\|{\bf n}_{\boldsymbol{\eta}}(t)\|_{{\bf L}^{2}(\Gamma)}\|\Lambda^{2\kappa}{\boldsymbol{\eta}}(t)\|_{{\bf L}^{2}(\Gamma)}.

Note, due to the periodic boundary conditions imposed on the structure displacement, 𝝍\boldsymbol{\psi} is indeed a valid structure test function.

Now for the solenoidal function 𝝋\boldsymbol{\varphi} in (22), we define it to be the solution of a time-dependent Stokes problem with non-homogeneous boundary data defined as follows.

For any fixed δ\delta such that 0<δ<s0<\delta<s, let κ\kappa be as defined in (21), namely κ=1(sδ)\kappa=1-(s-\delta). Then we choose the solenoidal function 𝝋\boldsymbol{\varphi} in (22) to be the solution of

(25) 𝝋tΔ𝝋+p=0 in 𝒪,div 𝝋=0 in 𝒪,𝝋=J𝜼(𝐀𝜼)1|Γ(Λ2κ𝜼c𝝃) on Γ,𝝋(t=0)=𝝋0,\begin{split}\boldsymbol{\varphi}_{t}-\Delta\boldsymbol{\varphi}+\nabla p&=0\qquad\text{ in }\mathcal{O},\\ \text{div }\boldsymbol{\varphi}&=0\qquad\text{ in }\mathcal{O},\\ \boldsymbol{\varphi}&=J_{{\boldsymbol{\eta}}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}|_{\Gamma}\left(\Lambda^{2\kappa}{\boldsymbol{\eta}}-c\boldsymbol{\xi}\right)\qquad\text{ on }\Gamma,\\ \boldsymbol{\varphi}(t=0)&=\boldsymbol{\varphi}_{0},\end{split}

such that initial condition satisfies,

𝝋0|Γ\displaystyle\boldsymbol{\varphi}_{0}|_{\Gamma} =J𝜼0(A𝜼0)1|Γ(Λ2κ𝜼0c𝝃(0)).\displaystyle=J_{{\boldsymbol{\eta}}_{0}}(\nabla A_{{\boldsymbol{\eta}}_{0}})^{-1}|_{\Gamma}\left(\Lambda^{2\kappa}{\boldsymbol{\eta}}_{0}-c\boldsymbol{\xi}(0)\right).

Indeed, observe that this is the right choice of boundary value since it satisfies the following compatibility condition:

Γ𝝋(0,0,1)=Γ𝜼×(Λ2κ𝜼c𝝃)=0.\int_{\Gamma}\boldsymbol{\varphi}\cdot(0,0,1)=\int_{\Gamma}\nabla{\boldsymbol{\eta}}\times(\Lambda^{2\kappa}{\boldsymbol{\eta}}-c\boldsymbol{\xi})=0.

Now, for an appropriate choice of the trace space 𝒢m(Γ×(0,T))\mathcal{G}^{m}(\Gamma\times(0,T)), Theorem 6.1 in [7] guarantees the existence of a unique solution (𝝋,p)(\boldsymbol{\varphi},p) to (25) that satisfies

𝝋L2(0,T;𝐇m(𝒪))+t𝝋L2(0,T;𝐇m2(𝒪))\displaystyle\|\boldsymbol{\varphi}\|_{L^{2}(0,T;{\bf H}^{m}(\mathcal{O}))}+\|\partial_{t}\boldsymbol{\varphi}\|_{L^{2}(0,T;{\bf H}^{m-2}(\mathcal{O}))} +pL2(0,T;𝐇m2(𝒪))\displaystyle+\|\nabla p\|_{L^{2}(0,T;{\bf H}^{m-2}(\mathcal{O}))}
(26) J𝜼(𝐀𝜼)1|Γ(Λ2κ𝜼c𝝃)𝒢m(Γ×(0,T)).\displaystyle\leq\|J_{{\boldsymbol{\eta}}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}|_{\Gamma}(\Lambda^{2\kappa}{\boldsymbol{\eta}}-c\boldsymbol{\xi})\|_{\mathcal{G}^{m}(\Gamma\times(0,T))}.

We will consider 𝒢m(ΓT)\mathcal{G}^{m}(\Gamma_{T}) with 32<m<2\frac{3}{2}<m<2. This choice of mm balances the following two considerations: the chosen mm has to be large enough to bound the time derivative of the test function 𝐪{\bf q} in an appropriate dual space, which will be discussed later in estimate (42) (see the remark following the estimate), while still ensuring that mm is not too large in order to capture the limited regularity of the boundary data in (25).

For any m>32m>\frac{3}{2} the trace space 𝒢m\mathcal{G}_{m} is endowed with the following norm,

ϕ𝒢m(Γ×(0,T)):=ϕL2(0,T;𝐇m12(Γ))+ϕ𝐧H1(0,T;𝐇m52(Γ))+ϕ𝝉H2m12m(0,T;𝐇(12m)(m12)(Γ)),\displaystyle\|\phi\|_{\mathcal{G}^{m}(\Gamma\times(0,T))}:=\|\phi\|_{L^{2}(0,T;{\bf H}^{m-\frac{1}{2}}(\Gamma))}+\|\phi\cdot{\bf n}\|_{H^{1}(0,T;{\bf H}^{m-\frac{5}{2}}(\Gamma))}+\|\phi_{\boldsymbol{\tau}}\|_{H^{\frac{2m-1}{2m}}(0,T;{\bf H}^{(1-\frac{2}{m})(m-\frac{1}{2})}(\Gamma))},

where 𝐧=(0,0,1){\bf n}=(0,0,1) is the unit normal to Γ\Gamma and ϕ𝝉\phi_{\boldsymbol{\tau}} is the projection of ϕ\phi onto the tangent space of Γ\Gamma.

Next we comment on the validity of the choice of such test functions (𝐪,𝝍)({\bf q},\boldsymbol{\psi}). In summary, we have defined

(27) 𝐪=J𝜼1𝐀𝜼𝝋,𝝍=(Λ2κ𝜼c𝝃),\displaystyle{\bf q}=-J_{{\boldsymbol{\eta}}}^{-1}\nabla{\bf A}_{\boldsymbol{\eta}}\,\boldsymbol{\varphi},\qquad\boldsymbol{\psi}=-(\Lambda^{2\kappa}{\boldsymbol{\eta}}-c\boldsymbol{\xi}),

where cc is given in (24) and 𝝋\boldsymbol{\varphi} solves (25). As mentioned earlier, due to the properties of the Piola transform (see e.g. Theorem 1.7-1 in [6]), we have

(28) div𝜼𝐪=J𝜼1(div𝝋)=0.\displaystyle\text{div}^{{\boldsymbol{\eta}}}{\bf q}=J_{{\boldsymbol{\eta}}}^{-1}(\text{div}\boldsymbol{\varphi})=0.

Moreover, it is also true that 𝐪|Γ=𝝍{\bf q}|_{\Gamma}=\boldsymbol{\psi} on (0,T)×Γ(0,T)\times\Gamma. Hence, we conclude that this pair (𝐪,𝝍)({\bf q},\boldsymbol{\psi}) is a valid test function for (LABEL:weaksol).

We proceed with the proof of Theorem 1 by replacing the test function 𝝍\boldsymbol{\psi} in the weak formulation (LABEL:weaksol) with the above-constructed 𝝍=(Λ2κ𝜼c𝝃)\boldsymbol{\psi}=-(\Lambda^{2\kappa}{\boldsymbol{\eta}}-c\boldsymbol{\xi}), and then express the terms containing 𝜼{\boldsymbol{\eta}} that we want to estimate, using the remaining terms from the weak formulation. We obtain:

(29) γ20TΓddt|Λ1+κ+s𝜼|2+0TΓ|Λ2+κ𝜼|2=0TcΓ|tΛκ𝜼|2+0TΓt𝜼t𝝃+0TcΓΔ𝜼Δ𝝃γ0TcΓΛ1+st𝜼:Λ1+s𝝃+0T𝒪J𝜼𝐮t𝐪0T𝒪J𝜼(𝐮𝜼𝐮𝐪𝐰𝜼𝐮𝐪)+0T𝒪tJ𝜼𝐮𝐪2ν0T𝒪J𝜼𝐃𝜼(𝐮)𝐃𝜼(𝐪)+𝒪J𝜼0𝐮0𝐪(0)+Γ𝐯0𝝍(0),:=I1++I11.\begin{split}&\frac{\gamma}{2}\int_{0}^{T}\int_{\Gamma}\frac{d}{dt}|\Lambda^{1+\kappa+s}{\boldsymbol{\eta}}|^{2}+\int_{0}^{T}\int_{\Gamma}|\Lambda^{2+{\kappa}}{\boldsymbol{\eta}}|^{2}=\int_{0}^{T}c\int_{\Gamma}|\partial_{t}\Lambda^{\kappa}{{\boldsymbol{\eta}}}|^{2}\\ &+\int_{0}^{T}\int_{\Gamma}\partial_{t}{{\boldsymbol{\eta}}}\,\partial_{t}{\boldsymbol{\xi}}+\int_{0}^{T}c\int_{\Gamma}\Delta{\boldsymbol{\eta}}\cdot\Delta\boldsymbol{\xi}-\gamma\int_{0}^{T}c\int_{\Gamma}\Lambda^{1+s}\partial_{t}{\boldsymbol{\eta}}:\Lambda^{1+s}\boldsymbol{\xi}\\ &+\int_{0}^{T}\int_{\mathcal{O}}{J}_{{\boldsymbol{\eta}}}{\bf u}\cdot\partial_{t}{{\bf q}}-\int_{0}^{T}\int_{\mathcal{O}}J_{{\boldsymbol{\eta}}}({\bf u}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q}-{\bf w}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q})\\ &+\int_{0}^{T}\int_{\mathcal{O}}\partial_{t}J_{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q}-2\nu\int_{0}^{T}\int_{\mathcal{O}}{J}_{{\boldsymbol{\eta}}}\,{\bf D}^{{{\boldsymbol{\eta}}}}({\bf u})\cdot{\bf D}^{{{\boldsymbol{\eta}}}}({\bf q})\\ &+\int_{\mathcal{O}}J_{{\boldsymbol{\eta}}_{0}}{\bf u}_{0}{\bf q}(0)+\int_{\Gamma}{\bf v}_{0}\boldsymbol{\psi}(0),\\ &:=I_{1}+...+I_{11}.\end{split}

In the rest of this proof we will estimate the terms Ij,1j11I_{j},1\leq j\leq 11 to get the desired final estimate.

However, before estimating each term Ij,1j11I_{j},1\leq j\leq 11, we plan to obtain bounds for 𝝋\boldsymbol{\varphi} that will result in appropriate bounds for the test function 𝐪{\bf q} (see (27)), which will require bounds for 𝝋|Γ=J𝜼(𝐀𝜼)1|Γ(Λ2κ𝜼c𝝃)\boldsymbol{\varphi}|_{\Gamma}=J_{{\boldsymbol{\eta}}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}|_{\Gamma}\left(\Lambda^{2\kappa}{\boldsymbol{\eta}}-c\boldsymbol{\xi}\right) in the trace space 𝒢m(Γ×(0,T))\mathcal{G}^{m}(\Gamma\times(0,T)) for a well-chosen m>32m>\frac{3}{2}. More precisely, we plan to use (26) to show the following estimate of 𝝋|Γ\boldsymbol{\varphi}|_{\Gamma}:

Proposition 3.2.

For any 0<δ<s0<\delta<s let m=32+εm=\frac{3}{2}+\varepsilon where 0<ε<min{δ/4,sδ}0<\varepsilon<\min\{\delta/4,s-\delta\}. Then, the function 𝛗\boldsymbol{\varphi} defined to be the solution of (25), satisfies the following trace estimate:

(30) 𝝋|Γ𝒢m(Γ×(0,T))C(1+𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ))),\displaystyle\|\boldsymbol{\varphi}|_{\Gamma}\|_{\mathcal{G}^{m}(\Gamma\times(0,T))}\leq C(1+\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}),

where κ\kappa is defined in (21).

Proof.

(Proof of Proposition 3.2) First, we write m=32+εm=\frac{3}{2}+\varepsilon. Now, observe that for any 0<ε<1κ=sδ0<\varepsilon<1-\kappa=s-\delta, the following estimate holds true:

Λ2κ𝜼L2(0,T;𝐇m12(Γ))=Λ2κ𝜼L2(0,T;𝐇1+ε(Γ))=𝜼L2(0,T;𝐇2κ+1+ε(Γ))𝜼L2(0,T;𝐇2+κ(Γ)).\displaystyle\|\Lambda^{2\kappa}{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{m-\frac{1}{2}}(\Gamma))}=\|\Lambda^{2\kappa}{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{1+\varepsilon}(\Gamma))}=\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2\kappa+1+\varepsilon}(\Gamma))}\leq\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}.

Due to the trace theorem, the fact that H32+ε(𝒪)H^{\frac{3}{2}+\varepsilon}(\mathcal{O}) is a Banach algebra, and using the Sobolev estimate for harmonic extensions (13), we have

J𝜼(𝐀𝜼)1|Γ𝐇1+ε(Γ)CJ𝜼(𝐀𝜼)1𝐇32+ε(𝒪)C𝐀𝜼𝐇52+ε(𝒪)2C𝜼𝐇2+ε(Γ)2.\displaystyle\|J_{\boldsymbol{\eta}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}|_{\Gamma}\|_{{\bf H}^{1+\varepsilon}(\Gamma)}\leq C\|J_{\boldsymbol{\eta}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}\|_{{\bf H}^{\frac{3}{2}+\varepsilon}(\mathcal{O})}\leq C\|{\bf A}_{\boldsymbol{\eta}}\|^{2}_{{\bf H}^{\frac{5}{2}+\varepsilon}(\mathcal{O})}\leq C\|{\boldsymbol{\eta}}\|^{2}_{{\bf H}^{2+\varepsilon}(\Gamma)}.

We now interpolate the right-hand side as follows,

J𝜼(𝐀𝜼)1|ΓL(0,T;𝐇1+ε(Γ))\displaystyle\|J_{\boldsymbol{\eta}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}|_{\Gamma}\|_{L^{\infty}(0,T;{\bf H}^{1+\varepsilon}(\Gamma))} C𝜼L(0,T;𝐇2+ε(Γ))2\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{2}_{L^{\infty}(0,T;{\bf H}^{2+\varepsilon}(\Gamma))}
C𝜼L(0,T;𝐇2(Γ))32𝜼L(0,T;𝐇2+4ε(Γ))12,\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{\frac{3}{2}}_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+4\varepsilon}(\Gamma))},
(31) C𝜼L(0,T;𝐇2+4ε(Γ))12.\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+4\varepsilon}(\Gamma))}.

Hence, choosing

(32) 0<ε<min{δ4,1κ}0<\varepsilon<\min\{\frac{\delta}{4},1-\kappa\}

and noting that for s>1s>1, Hs(Γ)H^{s}(\Gamma) is a Banach algebra, we obtain

𝝋|ΓL2(0,T;𝐇m12(Γ))\displaystyle\|\boldsymbol{\varphi}|_{\Gamma}\|_{L^{2}(0,T;{\bf H}^{m-\frac{1}{2}}(\Gamma))} CJ𝜼(𝐀𝜼)1Λ2κ𝜼|ΓL2(0,T;𝐇1+ε(Γ))\displaystyle\leq C\|J_{\boldsymbol{\eta}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}\Lambda^{2\kappa}{\boldsymbol{\eta}}|_{\Gamma}\|_{L^{2}(0,T;{\bf H}^{1+\varepsilon}(\Gamma))}
CJ𝜼(𝐀𝜼)1|ΓL(0,T;𝐇1+ε(Γ))Λ2κ𝜼L2(0,T;𝐇1+ε(Γ))\displaystyle\leq C\|J_{\boldsymbol{\eta}}(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}|_{\Gamma}\|_{L^{\infty}(0,T;{\bf H}^{1+\varepsilon}(\Gamma))}\|\Lambda^{2\kappa}{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{1+\varepsilon}(\Gamma))}
C𝜼L(0,T;𝐇2+4ε(Γ))12𝜼L2(0,T;𝐇2+κ(Γ))\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+4\varepsilon}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}
(33) C𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ)).\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}.

This gives an estimate on the first term in the definition of 𝒢m(Γ×(0,T))\mathcal{G}_{m}(\Gamma\times(0,T)). Now we focus on the second term. We observe that

Λ2κ𝜼H1(0,T;𝐇m52(Γ))C𝜼H1(0,T;𝐇1+ε+2κ(Γ))C𝜼H1(0,T;𝐇1+s(Γ))<C.\displaystyle\|\Lambda^{2\kappa}{\boldsymbol{\eta}}\|_{H^{1}(0,T;{\bf H}^{m-\frac{5}{2}}(\Gamma))}\leq C\|{\boldsymbol{\eta}}\|_{H^{1}(0,T;{\bf H}^{-1+\varepsilon+2\kappa}(\Gamma))}\leq C\|{\boldsymbol{\eta}}\|_{H^{1}(0,T;{\bf H}^{1+s}(\Gamma))}<C.

Hence, by combining this observation with (31) and by applying Theorem 8.2 in [1], we find

t(J𝜼𝐀𝜼1Λ2κ𝜼|Γ)L2(0,T;𝐇m52(Γ))\displaystyle\|\partial_{t}(J_{\boldsymbol{\eta}}\nabla{\bf A}_{\boldsymbol{\eta}}^{-1}\Lambda^{2\kappa}{\boldsymbol{\eta}}|_{\Gamma})\|_{L^{2}(0,T;{\bf H}^{m-\frac{5}{2}}(\Gamma))} (J𝜼A𝜼1)|ΓL(0,T;𝐇1+ε(Γ))Λ2κ𝜼H1(0,T;𝐇m52(Γ))\displaystyle\leq\|(J_{\boldsymbol{\eta}}\nabla A_{\boldsymbol{\eta}}^{-1})|_{\Gamma}\|_{L^{\infty}(0,T;{\bf H}^{1+\varepsilon}(\Gamma))}\|\Lambda^{2\kappa}{\boldsymbol{\eta}}\|_{H^{1}(0,T;{\bf H}^{m-\frac{5}{2}}(\Gamma))}
+(J𝜼A𝜼1)|ΓH1(0,T;𝐇ε(Γ))Λ2κ𝜼L(0,T;𝐇m32(Γ))\displaystyle+\|(J_{\boldsymbol{\eta}}\nabla A_{\boldsymbol{\eta}}^{-1})|_{\Gamma}\|_{H^{1}(0,T;{\bf H}^{\varepsilon}(\Gamma))}\|\Lambda^{2\kappa}{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{m-\frac{3}{2}}(\Gamma))}
C𝜼L(0,T;𝐇2+4ε(Γ))12𝜼H1(0,T;𝐇1+s(Γ))\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+4\varepsilon}(\Gamma))}\|{\boldsymbol{\eta}}\|_{H^{1}(0,T;{\bf H}^{1+s}(\Gamma))}
+Ct𝜼L2(0,T;𝐇1+s(Γ))2𝜼L(0,T;𝐇ε+2κ(Γ)).\displaystyle+C\|\partial_{t}{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{1+s}(\Gamma))}^{2}\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{\varepsilon+2\kappa}(\Gamma))}.

Since we took ε<min{δ4,1κ}\varepsilon<\min\{\frac{\delta}{4},1-\kappa\} and κ<1\kappa<1, we see that 𝜼L(0,T;𝐇ε+2κ(Γ))𝜼L(0,T;𝐇2(Γ))C\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{\varepsilon+2\kappa}(\Gamma))}\leq\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}\leq C. Hence we conclude that,

(34) 𝝋|ΓH1(0,T;𝐇m52(Γ))\displaystyle\|\boldsymbol{\varphi}|_{\Gamma}\|_{H^{1}(0,T;{\bf H}^{m-\frac{5}{2}}(\Gamma))} C+C𝜼L(0,T;𝐇2+δ(Γ))12.\displaystyle\leq C+C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}.

Finally, interpolating between the spaces in (33) and (34), we obtain

𝝋|ΓH2m12m(0,T;𝐇(12m)(m12)(Γ))\displaystyle\|\boldsymbol{\varphi}|_{\Gamma}\|_{H^{\frac{2m-1}{2m}}(0,T;{\bf H}^{(1-\frac{2}{m})(m-\frac{1}{2})}(\Gamma))} 𝝋|ΓH1(0,T;𝐇m52(Γ))2m12m𝝋|ΓL2(0,T;𝐇m12(Γ))12m\displaystyle\leq\|\boldsymbol{\varphi}|_{\Gamma}\|^{\frac{2m-1}{2m}}_{H^{1}(0,T;{\bf H}^{m-\frac{5}{2}}(\Gamma))}\|\boldsymbol{\varphi}|_{\Gamma}\|^{\frac{1}{2m}}_{L^{2}(0,T;{\bf H}^{m-\frac{1}{2}}(\Gamma))}
(35) C+C𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ)).\displaystyle\leq C+C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}.

Finally, by combining all the estimates (33), (34) and (35), we arrive at the desired result (30). This completes the proof of Proposition 3.2. ∎

This proposition implies that for m=32+εm=\frac{3}{2}+\varepsilon where 0<ε<sδ10<\varepsilon<s-\delta\ll 1, we can continue estimating the right hand-side of (26) to obtain

(36) 𝝋L2(0,T;𝐇m(𝒪))+t𝝋L2(0,T;𝐇m2(𝒪))C(1+𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ))).\displaystyle\|\boldsymbol{\varphi}\|_{L^{2}(0,T;{\bf H}^{m}(\mathcal{O}))}+\|\partial_{t}\boldsymbol{\varphi}\|_{L^{2}(0,T;{\bf H}^{m-2}(\mathcal{O}))}\leq C(1+\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}).

Moreover, thanks to the Lions-Magenes theorem, (36) further gives us

(37) 𝝋C(0,T;𝐇12+ε(𝒪))=𝝋C(0,T;𝐇m1(𝒪))C(1+𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ))),\displaystyle\|\boldsymbol{\varphi}\|_{C(0,T;{\bf H}^{\frac{1}{2}+\varepsilon}(\mathcal{O}))}=\|\boldsymbol{\varphi}\|_{C(0,T;{\bf H}^{m-1}(\mathcal{O}))}\leq C(1+\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}),

which will ultimately be used in deriving bounds for the nonlinear term in the Navier-Stokes equations.

Now we turn our attention to deriving the relevant estimates for 𝐪{\bf q}. We will do so by using the relation (27) and the estimates (36) and (37).
First, since we have the embedding H12(𝒪)L3(𝒪)H^{\frac{1}{2}}(\mathcal{O})\hookrightarrow L^{3}(\mathcal{O}), estimate (36) gives us,

(38) 𝐪L2(0,T;𝐇1(𝒪))C𝐀𝜼L(0,T;𝐇2.5(𝒪))𝝋L2(0,T;𝐇32(𝒪))C𝜼L(0,T;𝐇2(Γ))(1+𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ)))C𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ)).\begin{split}\|{\bf q}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}&\leq C\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2.5}(\mathcal{O}))}\|\boldsymbol{\varphi}\|_{L^{2}(0,T;{\bf H}^{\frac{3}{2}}(\mathcal{O}))}\\ &\leq C\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}(1+\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))})\\ &\leq C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}.\end{split}

However, to deal with the nonlinear term in the Navier-Stokes equations, this estimate will not be sufficient. Thus, using (37) we arrive at the following estimate which is later used to find bounds for the terms I7I_{7} and I8I_{8}:

(39) 𝐪L(0,T;𝐇12(𝒪))C𝐀𝜼L(0,T;𝐇2.5(𝒪))𝝋L(0,T;𝐇m1(𝒪))C𝜼L(0,T;𝐇2(Γ))(1+𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ)))C𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ)).\begin{split}\|{\bf q}\|_{L^{\infty}(0,T;{\bf H}^{\frac{1}{2}}(\mathcal{O}))}&\leq C\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2.5}(\mathcal{O}))}\|\boldsymbol{\varphi}\|_{L^{\infty}(0,T;{\bf H}^{m-1}(\mathcal{O}))}\\ &\leq C\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}(1+\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))})\\ &\leq C\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}.\end{split}

Finally, we expand:

(40) t𝐪\displaystyle\partial_{t}{\bf q} =tJ𝜼𝐀𝜼1𝝋+J𝜼t𝐀𝜼1𝝋+J𝜼𝐀𝜼1t𝝋.\displaystyle=\partial_{t}J_{\boldsymbol{\eta}}\nabla{\bf A}_{\boldsymbol{\eta}}^{-1}\boldsymbol{\varphi}+J_{\boldsymbol{\eta}}\partial_{t}\nabla{\bf A}_{\boldsymbol{\eta}}^{-1}\boldsymbol{\varphi}+J_{\boldsymbol{\eta}}\nabla{\bf A}_{\boldsymbol{\eta}}^{-1}\partial_{t}\boldsymbol{\varphi}.

We know that (see e.g. [6]),

tJ𝜼=(J𝜼)2tr((cof 𝐀𝜼)Tt𝐀𝜼),\partial_{t}J_{{\boldsymbol{\eta}}}=-(J_{{\boldsymbol{\eta}}})^{-2}\text{tr}((\text{cof }{\bf A}_{\boldsymbol{\eta}})^{T}\partial_{t}\nabla{\bf A}_{\boldsymbol{\eta}}),

and thus handling of the first two terms on the right-hand side of (40) is straight-forward, as these terms remain bounded in L2(0,T;𝐋2(𝒪))L^{2}(0,T;{\bf L}^{2}(\mathcal{O})). For the third term on the right-hand side of (40) we apply Theorem 8.1 in [1]. Using (36), we observe that for

m=32+ε,ε<δ8,andanyq>32m=\frac{3}{2}+\varepsilon,\ \varepsilon<\frac{\delta}{8},\ {\rm and\ any}\ q>\frac{3}{2}

the following bound on 𝐪{\bf q} holds true:

t𝐪L2(0,T;𝐇m2(𝒪))\displaystyle\|\partial_{t}{\bf q}\|_{L^{2}(0,T;{\bf H}^{m-2}(\mathcal{O}))} CJ𝜼𝐀𝜼1L(0,T;𝐇q(𝒪))t𝝋L2(0,T;𝐇m2(𝒪))\displaystyle\leq C\|J_{\boldsymbol{\eta}}\nabla{\bf A}_{\boldsymbol{\eta}}^{-1}\|_{L^{\infty}(0,T;{\bf H}^{q}(\mathcal{O}))}\|\partial_{t}\boldsymbol{\varphi}\|_{L^{2}(0,T;{\bf H}^{m-2}(\mathcal{O}))}
C𝜼L(0,T;𝐇2+ε(Γ))2𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ))\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{2}_{L^{\infty}(0,T;{\bf H}^{2+\varepsilon}(\Gamma))}\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}
C𝜼L(0,T;𝐇2(Γ))74𝜼L(0,T;𝐇2+8ε(Γ))14𝜼L(0,T;𝐇2+δ(Γ))12𝜼L2(0,T;𝐇2+κ(Γ))\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{\frac{7}{4}}_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}\|{\boldsymbol{\eta}}\|^{\frac{1}{4}}_{L^{\infty}(0,T;{\bf H}^{2+8\varepsilon}(\Gamma))}\|{\boldsymbol{\eta}}\|^{\frac{1}{2}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}
(41) C𝜼L(0,T;𝐇2+δ(Γ))34𝜼L2(0,T;𝐇2+κ(Γ)).\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{\frac{3}{4}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}.

This completes the derivation of estimates for 𝐪{\bf q} and 𝝍\boldsymbol{\psi} that we will use to bound the integrals IjI_{j}, 1j111\leq j\leq 11 in (29).

We start discussing the bounds of integrals IjI_{j}, 1j111\leq j\leq 11 in (29) by noticing that since 𝝃\boldsymbol{\xi} is smooth, the bounds for the first 4 terms on the right hand side follow straight from the energy estimates derived in (6). That is,

|I1++I4|C,\displaystyle|I_{1}+...+I_{4}|\leq C,

where C>0C>0 depends only on the given data 𝐮0,𝐯0,𝜼0{\bf u}_{0},{\bf v}_{0},{\boldsymbol{\eta}}_{0}. Note that, this constant technically also depends on the norms t𝝃L2(0,;𝐋2(𝒪))\|\partial_{t}\boldsymbol{\xi}\|_{L^{2}(0,;{\bf L}^{2}(\mathcal{O}))} and 𝝃L(0,T;𝐂2(Γ))\|\boldsymbol{\xi}\|_{L^{\infty}(0,T;{\bf C}^{2}(\Gamma))} which, according to our choice, are equal to 1.

Next, we present the derivation of the estimates that require further explanation. We begin with I5I_{5}. To estimate I5I_{5} we will use (41). Since H0r(𝒪)=Hr(𝒪)H^{r}_{0}(\mathcal{O})=H^{r}(\mathcal{O}) for any r<12r<\frac{1}{2}, we obtain the following estimate which holds for 32<m<2\frac{3}{2}<m<2:

(42) |I5|\displaystyle|I_{5}| =|0T𝒪J𝜼𝐮t𝐪|0TJ𝜼𝐮𝐇2m(𝒪)t𝐪𝐇m2(𝒪)\displaystyle=|\int_{0}^{T}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}{\bf u}\cdot\partial_{t}{\bf q}|\leq\int_{0}^{T}\|J_{\boldsymbol{\eta}}{\bf u}\|_{{\bf H}^{2-m}(\mathcal{O})}\|\partial_{t}{\bf q}\|_{{\bf H}^{m-2}(\mathcal{O})}
C0TJ𝜼𝐇32(𝒪)𝐮𝐇1(𝒪)t𝐪𝐇m2(𝒪)\displaystyle\leq C\int_{0}^{T}\|J_{\boldsymbol{\eta}}\|_{{\bf H}^{\frac{3}{2}}(\mathcal{O})}\|{\bf u}\|_{{\bf H}^{1}(\mathcal{O})}\|\partial_{t}{\bf q}\|_{{\bf H}^{m-2}(\mathcal{O})}
C𝐀𝜼L(0,T;𝐇52(𝒪))𝐮L2(0,T;𝐇1(𝒪))t𝐪L2(0,T;𝐇m2(𝒪))\displaystyle\leq C\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{\frac{5}{2}}(\mathcal{O}))}\|{\bf u}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}\|\partial_{t}{\bf q}\|_{L^{2}(0,T;{\bf H}^{m-2}(\mathcal{O}))}
C𝜼L(0,T;𝐇2(Γ))𝐮L2(0,T;𝐇1(𝒪))𝜼L(0,T;𝐇2+δ(Γ))34𝜼L2(0,T;𝐇2+κ(Γ))\displaystyle\leq C\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}\|{\bf u}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}^{\frac{3}{4}}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}
CK18+116𝜼L(0,T;𝐇2+δ(Γ))2+18𝜼L2(0,T;𝐇2+κ(Γ))2,\displaystyle\leq CK^{8}_{1}+\frac{1}{16}\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}^{2}+\frac{1}{8}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}^{2},

where K1K_{1} is the constant from Lemma 3.1 that depends on α1\alpha_{1}, 𝐀𝜼L(0,T;𝐖1,(𝒪))\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))} and (𝐀𝜼)1L(0,T;𝐖1,(𝒪))\|({\bf A}_{\boldsymbol{\eta}})^{-1}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}.

Remark 5.

Note that the choice 32<m<2\frac{3}{2}<m<2 plays an important role here as we use the duality between HrH^{r} and HrH^{-r} for r<12r<\frac{1}{2} on the right hand-side in the first line of the estimate of I5I_{5} above.

To estimate the nonlinear terms in I6I_{6} we use (39), to obtain

|I6|\displaystyle|I_{6}| C0T(𝐮+𝐰𝐋6(𝒪))𝐮𝐇1(𝒪)𝐪𝐋3(𝒪)C0T𝐮𝐇1(𝒪)2𝐪𝐇12(𝒪)\displaystyle\leq C\int_{0}^{T}(\|{\bf u}+{\bf w}\|_{{\bf L}^{6}(\mathcal{O})})\|{\bf u}\|_{{\bf H}^{1}(\mathcal{O})}\|{\bf q}\|_{{\bf L}^{3}(\mathcal{O})}\leq C\int_{0}^{T}\|{\bf u}\|^{2}_{{\bf H}^{1}(\mathcal{O})}\|{\bf q}\|_{{\bf H}^{\frac{1}{2}}(\mathcal{O})}
C𝐮L2(0,T;𝐇1(𝒪))2𝐪L(0,T;𝐇12(𝒪))\displaystyle\leq C\|{\bf u}\|^{2}_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}\|{\bf q}\|_{L^{\infty}(0,T;{\bf H}^{\frac{1}{2}}(\mathcal{O}))}
CK18+116𝜼L(0,T;𝐇2+δ(Γ))2+18𝜼L2(0,T;𝐇2+κ(Γ))2.\displaystyle\leq CK_{1}^{8}+\frac{1}{16}\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}^{2}+\frac{1}{8}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}^{2}.

Similarly, using (39) we estimate I7I_{7}:

|I7|=|0t𝒪tJ𝜼𝐮𝐪|\displaystyle|I_{7}|=|\int_{0}^{t}\int_{\mathcal{O}}\partial_{t}J_{\boldsymbol{\eta}}{\bf u}\cdot{\bf q}| 𝐰L2(0,T;𝐇1(𝒪))𝐮L2(0,T;𝐋6(𝒪))𝐪L(0,T;𝐋3(𝒪))\displaystyle\leq\|{\bf w}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}\|{\bf u}\|_{L^{2}(0,T;{\bf L}^{6}(\mathcal{O}))}\|{\bf q}\|_{L^{\infty}(0,T;{\bf L}^{3}(\mathcal{O}))}
CK14+116𝜼L(0,T;𝐇2+δ(Γ))2+18𝜼L2(0,T;𝐇2+κ(Γ))2.\displaystyle\leq CK^{4}_{1}+\frac{1}{16}\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}^{2}+\frac{1}{8}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}^{2}.

The symmetrized gradient integral I8I_{8} is estimated using (38) to obtain

|I8|\displaystyle|I_{8}| 𝐮L2(0,T;𝐇1(𝒪))𝐪L2(0,T;𝐇1(𝒪))\displaystyle\leq\|{\bf u}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}\|{\bf q}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}
CK14+116𝜼L(0,T;𝐇2+δ(Γ))2+18𝜼L2(0,T;𝐇2+κ(Γ))2.\displaystyle\leq CK_{1}^{4}+\frac{1}{16}\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}^{2}+\frac{1}{8}\|{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}^{2}.

Hence, absorbing appropriate terms on the left hand side we obtain

(43) 14𝜼L(0,T;𝐇2+δ(Γ))2+12𝜼L2(0,T;𝐇2+κ(Γ))2K2+𝜼0𝐇2+δ(Γ)2,\displaystyle\frac{1}{4}\|{\boldsymbol{\eta}}\|^{2}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}+\frac{1}{2}\|{\boldsymbol{\eta}}\|^{2}_{L^{2}(0,T;{\bf H}^{2+\kappa}(\Gamma))}\leq K_{2}+\|{\boldsymbol{\eta}}_{0}\|^{2}_{{\bf H}^{2+\delta}(\Gamma)},

where the constant K2K_{2} depends on 𝐀𝜼L(0,T;𝐖1,(𝒪)),(𝐀𝜼)1L(0,T;𝐖1,(𝒪))\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))},\|({\bf A}_{\boldsymbol{\eta}})^{-1}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))} and α\alpha due to its dependence on K1K_{1} from Lemma 3.1. Recall here that for any 0<δ<s0<\delta<s, we chose κ=1(sδ)\kappa=1-(s-\delta).

Bootstrap argument: We will next prove the estimate (19) i.e. we will get rid of the dependence of K2K_{2}, appearing in the right-hand side of (43), on the norm 𝐀𝜼L(0,T;𝐖1,(𝒪))\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))} and on α<inf𝒪×(0,T)J𝜼\alpha<\inf_{\mathcal{O}\times(0,T)}J_{\boldsymbol{\eta}}. We do this by possibly shrinking the time length on which this desired estimate holds by using a bootstrap argument [29, Propostion 1.21]. Observe that, if for a fixed α\alpha and some C0>0C_{0}>0 we have

(44) 𝐀𝜼L(0,T;𝐖1,(𝒪))2C0andinf𝒪×(0,T)J𝜼>α,\displaystyle\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}\leq 2C_{0}\quad\text{and}\quad\inf_{\mathcal{O}\times(0,T)}J_{{\boldsymbol{\eta}}}>\alpha,

then according to (43) there exists a constant K2>0K_{2}>0 depending on C0C_{0} such that,

𝜼L(0,T;𝐇2+δ(Γ))K2.\displaystyle\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\leq K_{2}.

Furthermore, Sobolev and interpolation inequalities imply for any δ>0\delta>0 that

𝐀𝜼𝐀𝜼0L(0,T;𝐖1,(𝒪))\displaystyle\|{\bf A}_{\boldsymbol{\eta}}-{\bf A}_{{\boldsymbol{\eta}}_{0}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))} C𝐀𝜼𝐀𝜼0L(0,T;𝐇5+δ2(𝒪))C𝜼𝜼0L(0,T;𝐇2+δ2(Γ))\displaystyle\leq C\|{\bf A}_{\boldsymbol{\eta}}-{\bf A}_{{\boldsymbol{\eta}}_{0}}\|_{L^{\infty}(0,T;{\bf H}^{\frac{5+\delta}{2}}(\mathcal{O}))}\leq C\|{\boldsymbol{\eta}}-{\boldsymbol{\eta}}_{0}\|_{L^{\infty}(0,T;{\bf H}^{2+\frac{\delta}{2}}(\Gamma))}
C𝜼𝜼0L(0,T;𝐇2+δ(Γ))2+δ22+δ𝜼𝜼0L(0,T;𝐋2(Γ))δ22+δ\displaystyle\leq C\|{\boldsymbol{\eta}}-{\boldsymbol{\eta}}_{0}\|^{\frac{2+\frac{\delta}{2}}{2+\delta}}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\boldsymbol{\eta}}-{\boldsymbol{\eta}}_{0}\|^{\frac{\frac{\delta}{2}}{2+\delta}}_{L^{\infty}(0,T;{\bf L}^{2}(\Gamma))}
C(K2)2+δ22+δTδ2(2+δ)t𝜼L(0,T;𝐋2(Γ))δ2(2+δ)\displaystyle\leq C(K_{2})^{\frac{2+\frac{\delta}{2}}{2+\delta}}T^{\frac{\delta}{2(2+\delta)}}\|\partial_{t}{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf L}^{2}(\Gamma))}^{\frac{\delta}{2(2+\delta)}}
(45) C(K2)βC11βT1β,where β=2+δ22+δ,\displaystyle\leq C(K_{2})^{\beta}C_{1}^{1-\beta}T^{1-\beta},\qquad\text{where }\beta=\frac{2+\frac{\delta}{2}}{2+\delta},

where CC depends only on 𝒪\mathcal{O}, K2K_{2} depends on C0C_{0}, and the constant C1C_{1} appearing in (6) depends on the given data 𝐮0,𝐯0,𝜼0{\bf u}_{0},{\bf v}_{0},{\boldsymbol{\eta}}_{0}.

Similarly,

inf𝒪×(0,T)J𝜼(t)\displaystyle\inf_{\mathcal{O}\times(0,T)}J_{{\boldsymbol{\eta}}}(t) inf𝒪J0sup𝒪×(0,T)|J𝜼J𝜼0|\displaystyle\geq\inf_{\mathcal{O}}J_{0}-\sup_{\mathcal{O}\times(0,T)}|J_{{\boldsymbol{\eta}}}-J_{{\boldsymbol{\eta}}_{0}}|
inf𝒪J0C𝐀𝜼𝐀𝜼0L(0,T;𝐖1,(𝒪))\displaystyle\geq\inf_{\mathcal{O}}J_{0}-C\|{\bf A}_{\boldsymbol{\eta}}-{\bf A}_{{\boldsymbol{\eta}}_{0}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}
(46) inf𝒪J0C(K2)βC11βT1β.\displaystyle\geq\inf_{\mathcal{O}}J_{0}-C(K_{2})^{\beta}C_{1}^{1-\beta}T^{1-\beta}.

Hence, for small enough T0>0T_{0}>0, the hypothesis (44) then implies that

(47) 𝐀𝜼L(0,T0;𝐖1,(𝒪))C0 andinf𝒪×(0,T0)J𝜼>2α.\displaystyle\|{\bf A}_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf W}^{1,\infty}(\mathcal{O}))}\leq C_{0}\quad\text{ and}\quad\inf_{\mathcal{O}\times(0,T_{0})}J_{{\boldsymbol{\eta}}}>2\alpha.

This concludes our bootstrap argument and thus the proof of Theorem 1.

Next ,we focus on the proof of Theorem 2 to establish the estimate (20).

3.2. The temporal regularity result for fluid and structure velocities.

In this section we prove Theorem 2. Namely, the aim of this section is to show that for (𝐮,𝜼)({\bf u},{\boldsymbol{\eta}}), a pair of smooth functions that solves (LABEL:weaksol), there exists a constant C>0C>0, depending only on 𝒪\mathcal{O} and the given data, such that the following estimate holds for any h>0h>0:

(48) τh𝐮𝐮L2(h,T0;𝐋2(𝒪))+τht𝜼t𝜼L2(h,T0;𝐋2(Γ))Ch18,\displaystyle\|\tau_{h}{\bf u}-{\bf u}\|_{L^{2}(h,T_{0};{\bf L}^{2}(\mathcal{O}))}+\|\tau_{h}\partial_{t}{\boldsymbol{\eta}}-\partial_{t}{\boldsymbol{\eta}}\|_{L^{2}(h,T_{0};{\bf L}^{2}(\Gamma))}\leq Ch^{\frac{1}{8}},

where T0T_{0} is the time length appearing in Theorem 1.

To obtain the two terms on the left-hand side of the estimate above, we will construct an appropriate pair of test functions (𝐪,𝝍)({\bf q},\boldsymbol{\psi}) for the weak formulation (LABEL:weaksol) on the fixed domain 𝒪\mathcal{O}. A typical approach to obtaining results of this kind for the weak solutions of Navier-Stokes equations posed on a fixed domain, is to use the time integral tht\int_{t-h}^{t} of the solution as a test function. This approach cannot directly be employed in the case of moving boundary problems since the fluid velocity at different times is defined on different domains. Thus we face issues, due to the motion of the fluid domain, that arise due to the incompressibility condition and the kinematic coupling condition. Our plan, in the spirit of [11], is to construct the desired test function by first modifying the solution (𝐮,𝐯)({\bf u},{\bf v}) appropriately and then integrating it from tht-h to tt, for any t[0,T]t\in[0,T] and h>0h>0. This modification must preserve the divergence of the fluid velocity and its boundary behavior (i.e. the kinematic coupling condition) which is not trivial to find due to the time-varying domain. Note also that, due to the mismatch between the spatial regularity of the structure velocity t𝜼\partial_{t}{\boldsymbol{\eta}} and that of the test function 𝝍\boldsymbol{\psi} in the weak formulation Definition 2, this modification must also include a construction of a spatially regularized version of the structure velocity so that its time integral can be used as a test function in (LABEL:weaksol).

For the construction of the fluid test function, we will first extend the structure velocity 𝐯:=t𝜼{\bf v}:=\partial_{t}{\boldsymbol{\eta}} in the steady fluid domain 𝒪\mathcal{O}, subtract it from 𝐮{\bf u} and then integrate the resulting function from tht-h to hh against an appropriate kernel that possesses the desired property of preserving divergence in 𝒪\mathcal{O} and flux across 𝒪\partial\mathcal{O}. Now, to balance out this extra term in the fluid test function, i.e. the extension of 𝐯{\bf v}, and to construct the test function 𝝍\boldsymbol{\psi} that has the desired spatial H2H^{2}-regularity, we add the extension of the structure velocity 𝐯{\bf v} projected on a finite dimensional subspace of 𝐇2{\bf H}^{2} to the fluid test function. This finite-dimensional projection also enjoys nice properties that result in the second term appearing on the left-hand side of (20).

Finally, due to the addition of these two extra terms (i.e. the extension of 𝐯{\bf v} and that of its finite dimensional truncation) the transformed-divergence of the fluid test function has to be corrected. For that purpose, we construct a Bogovski-type operator on the physical moving domain with the aid of the Bogovski operator on the fixed domain 𝒪\mathcal{O}. This is crucial. Since, on the fixed domain, the test functions are required to satisfy the transformed-divergence-free condition, we correct it by multiplying it with the inverse of the cofactor matrix of 𝐀𝜼{\bf A}_{\boldsymbol{\eta}} and by using the Piola identity. We will now give precise definitions of our construction.

We fix h>0h>0. Let PMP_{M} denote the orthonormal projector in L2(Γ)L^{2}(\Gamma) onto the space span1iM{φi}\text{span}_{1\leq i\leq M}\{\varphi_{i}\}, where φi\varphi_{i} satisfies Δφi=λiφi-\Delta\varphi_{i}=\lambda_{i}\varphi_{i} and φi=0\varphi_{i}=0 on Γ\partial\Gamma. For any 𝐯𝐋2(Γ){\bf v}\in{\bf L}^{2}(\Gamma) we use the notation 𝐯M=PM𝐯{\bf v}_{M}=P_{M}{\bf v} (i.e. subscript MM) where PMP_{M} is the projection onto span{φj}1jM{}_{1\leq j\leq M}\{\varphi_{j}\}. We know that λMM2\lambda_{M}\sim M^{2} and hence we will choose

λM=ch34.\displaystyle\lambda_{M}=ch^{-\frac{3}{4}}.

We now construct a simple extension of 𝐯:=t𝜼{\bf v}:=\partial_{t}{\boldsymbol{\eta}} in the fluid domain. Let 𝐰=𝐯χ{\bf w}={\bf v}\chi where χ\chi is a cut-off function applied to 𝐯{\bf v} so that it does not have any contributions at the boundary 𝒪\partial\mathcal{O} i.e. χ\chi is a function smooth in 𝒪\mathcal{O} such that χ(x,y,1)=1\chi(x,y,1)=1 and χ(x,y,0)=0\chi(x,y,0)=0. Then, for any t[0,T0]t\in[0,T_{0}], we define our fluid test function as follows (see also [30]):

𝐪(t)\displaystyle{\bf q}(t) :=(J𝜼(t))1𝐀𝜼(t)tht(J𝜼(s)(𝐀𝜼(s))1(𝐮(s)𝐰(s)))𝑑s\displaystyle:=(J_{{\boldsymbol{\eta}}}(t))^{-1}\nabla{\bf A}_{\boldsymbol{\eta}}(t)\int_{t-h}^{t}\left(J_{\boldsymbol{\eta}}(s)(\nabla{\bf A}_{\boldsymbol{\eta}}(s))^{-1}({\bf u}(s)-{\bf w}(s))\right)ds
+tht(𝐰M(s)(bM(s,t)b0(s,t))𝝃0(s)χ)𝑑s\displaystyle+\int_{t-h}^{t}\left({\bf w}_{M}(s)-\left(\frac{b_{M}(s,t)}{b_{0}(s,t)}\right)\boldsymbol{\xi}_{0}(s)\chi\right)ds
(J𝜼(t))1𝐀𝜼(t)tht(div(J𝜼(s)(𝐀𝜼(s))1𝐰(s)J𝜼(t)(𝐀𝜼(t))1𝐰M(s))\displaystyle-(J_{{\boldsymbol{\eta}}}(t))^{-1}\nabla{\bf A}_{\boldsymbol{\eta}}(t)\int_{t-h}^{t}\mathcal{B}\Big{(}\text{div}\left(J_{\boldsymbol{\eta}}(s)(\nabla{\bf A}_{\boldsymbol{\eta}}(s))^{-1}{\bf w}(s)-J_{{\boldsymbol{\eta}}}(t)(\nabla{\bf A}_{\boldsymbol{\eta}}(t))^{-1}{\bf w}_{M}(s)\right)
+(bM(s,t)b0(s,t))div(J𝜼(t)𝐀𝜼1(t)𝝃0(s)χ))ds.\displaystyle+\left(\frac{b_{M}(s,t)}{b_{0}(s,t)}\right)\text{div}(J_{{\boldsymbol{\eta}}}(t)\nabla{\bf A}_{\boldsymbol{\eta}}^{-1}(t)\boldsymbol{\xi}_{0}(s)\chi)\Big{)}ds.

In other words, we define the fluid test function 𝐪{\bf q} as the time integral from tht-h to tt of a modification 𝐮M{\bf u}_{M} of 𝐮{\bf u}:

(49) 𝐪(t):=tht[Pt1Ps(𝐮(s)𝐰(s))+(𝐰M(s)𝖼M(s,t)𝝃0(s)χ)]𝑑sthtPt1(div(Ps𝐰(s)Pt𝐰M(s)+𝖼M(s,t)Pt𝝃0(s)χ))𝑑s:=tht𝐮M(s,t)𝑑s,\begin{split}{\bf q}(t)&:=\int_{t-h}^{t}\left[P_{t}^{-1}P_{s}({\bf u}(s)-{\bf w}(s))+\left({\bf w}_{M}(s)-\mathsf{c}_{M}(s,t)\boldsymbol{\xi}_{0}(s)\chi\right)\right]ds\\ &-\int_{t-h}^{t}P_{t}^{-1}\mathcal{B}\Big{(}\text{div}\left(P_{s}{\bf w}(s)-P_{t}{\bf w}_{M}(s)+\mathsf{c}_{M}(s,t)P_{t}\boldsymbol{\xi}_{0}(s)\chi\right)\Big{)}ds\\ &:=\int_{t-h}^{t}{\bf u}_{M}(s,t)ds,\end{split}

where:

  • 𝐰=𝐯χ{\bf w}={\bf v}\chi and 𝐰M=𝐯Mχ{\bf w}_{M}={\bf v}_{M}\chi,

  • χ\chi is a smooth function such that χ(x,y,1)=1\chi(x,y,1)=1 and χ(x,y,0)=0\chi(x,y,0)=0 for (x,y)Γ(x,y)\in\Gamma,

  • Pt(𝐟):=((J𝜼)𝐀𝜼1𝐟)(t)P_{t}({\bf f}):=((J_{{\boldsymbol{\eta}}})\nabla{\bf A}^{-1}_{{\boldsymbol{\eta}}}\cdot{\bf f})(t) is the Piola transformation composed with the ALE map 𝐀𝜼{\bf A}_{\boldsymbol{\eta}} for any 𝐟𝐋2(𝒪){\bf f}\in{\bf L}^{2}(\mathcal{O}).

  • \mathcal{B} is the Bogovski operator on the fixed domain 𝒪\mathcal{O}. Recall that (see e.g. [9, 8]) if 𝒪f=0\int_{\mathcal{O}}f=0 then

    (f)=f,and(f)𝐇01(𝒪)cfL2(𝒪).\nabla\cdot\mathcal{B}(f)=f,\quad\text{and}\quad\|\mathcal{B}(f)\|_{{\bf H}^{1}_{0}(\mathcal{O})}\leq c\|{f}\|_{L^{2}(\mathcal{O})}.
  • The correction term 𝖼M𝝃0χ\mathsf{c}_{M}\boldsymbol{\xi}_{0}\chi where 𝖼M(s,t)=b(s,s)bM(s,t)b0(s,t)L([0,T]2)\mathsf{c}_{M}(s,t)=\frac{b(s,s)-b_{M}(s,t)}{b_{0}(s,t)}\in L^{\infty}([0,T]^{2}) ensures that the above condition 𝒪f=0\int_{\mathcal{O}}f=0 is met.

  • Here, bM(s,t)=Γ(𝜼(t)×𝐯M(s))b_{M}(s,t)=\int_{\Gamma}(\nabla{\boldsymbol{\eta}}(t)\times{\bf v}_{M}(s)), b(s,t)=Γ(𝜼(t)×𝐯(s))b(s,t)=\int_{\Gamma}(\nabla{\boldsymbol{\eta}}(t)\times{\bf v}(s)) and b0(s,t)=Γ(𝜼(t)×𝝃0(s))b_{0}(s,t)=\int_{\Gamma}(\nabla{\boldsymbol{\eta}}(t)\times\boldsymbol{\xi}_{0}(s)),

  • 𝝃0𝐂0(Γ×[0,T0])\boldsymbol{\xi}_{0}\in{\bf C}^{\infty}_{0}(\Gamma\times[0,T_{0}]) is chosen such that b0(s,t)=1b_{0}(s,t)=1 for any s,t[0,T0]s,t\in[0,T_{0}].

For the structure test function we define:

(50) 𝝍(t):=tht(𝐯M(s)𝖼M(s,t)𝝃0(s))𝑑s.\displaystyle\boldsymbol{\psi}(t):=\int_{t-h}^{t}\left({\bf v}_{M}(s)-\mathsf{c}_{M}(s,t)\boldsymbol{\xi}_{0}(s)\right)ds.

We will first show that 𝐪{\bf q} has the required regularity. Thanks to the embedding

W1,(0,T;𝐋2(Γ))L(0,T;𝐇2(Γ))C0,θ(0,T;𝐇22θ(Γ)).W^{1,\infty}(0,T;{\bf L}^{2}(\Gamma))\cap L^{\infty}(0,T;{\bf H}^{2}(\Gamma))\hookrightarrow C^{0,\theta}(0,T;{\bf H}^{2-2\theta}(\Gamma)).

we have for any θ(0,1)\theta\in(0,1) and m22θm\geq 2-2\theta, that

(51) τh𝜼𝜼L(0,T0;𝐇m(Γ))𝜼C0,θ(0,T0;𝐇22θ(Γ))hθChθ.\displaystyle\|\tau_{h}{\boldsymbol{\eta}}-{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf H}^{m}(\Gamma))}\leq\|{\boldsymbol{\eta}}\|_{C^{0,\theta}(0,T_{0};{\bf H}^{2-2\theta}(\Gamma))}h^{\theta}\leq Ch^{\theta}.

Moreover, setting s=ths=t-h for any t[h,T0]t\in[h,T_{0}] we see that

|𝖼M(s,t)|\displaystyle|\mathsf{c}_{M}(s,t)| |b(s,s)bM(s,t)|=|Γ(𝜼(t)𝜼(s))×𝐯M(s)+𝜼(s)×(𝐯M𝐯)(s)|\displaystyle\leq|b(s,s)-b_{M}(s,t)|=|\int_{\Gamma}(\nabla{\boldsymbol{\eta}}(t)-\nabla{\boldsymbol{\eta}}(s))\times{\bf v}_{M}(s)+\nabla{\boldsymbol{\eta}}(s)\times({\bf v}_{M}-{\bf v})(s)|
𝜼(t)𝜼(s)𝐇1(Γ)𝐯M𝐋2(Γ)+𝜼(s)𝐇2(Γ)𝐯M(s)𝐯(s)𝐇1(Γ)\displaystyle\leq\|{\boldsymbol{\eta}}(t)-{\boldsymbol{\eta}}(s)\|_{{\bf H}^{1}(\Gamma)}\|{\bf v}_{M}\|_{{\bf L}^{2}(\Gamma)}+\|{\boldsymbol{\eta}}(s)\|_{{\bf H}^{2}(\Gamma)}\|{\bf v}_{M}(s)-{\bf v}(s)\|_{{\bf H}^{-1}(\Gamma)}
h12𝜼C0,12(0,T0;𝐇1(Γ))+λM1𝐯(s)𝐇1(Γ)\displaystyle\lesssim h^{\frac{1}{2}}\|{\boldsymbol{\eta}}\|_{C^{0,\frac{1}{2}}(0,T_{0};{\bf H}^{1}(\Gamma))}+\lambda_{M}^{-1}\|{\bf v}(s)\|_{{\bf H}^{1}(\Gamma)}
(52) Ch12.\displaystyle\leq Ch^{\frac{1}{2}}.

This estimate along with (18) and Lemma 3.1 further gives us the necessary bounds for the test function 𝐪{\bf q}:

𝐪L(0,T0;𝐇1(𝒪))\displaystyle\|{\bf q}\|_{L^{\infty}(0,T_{0};{\bf H}^{1}(\mathcal{O}))} sup0tT0(Pt𝐇2.5+δ(𝒪)2tht(𝐮+𝐰𝐇1(𝒪)+𝐰M𝐇1(𝒪)+|𝖼M(s,t)|)𝑑s)\displaystyle\leq\sup_{0\leq t\leq T_{0}}\left(\|P_{t}\|^{2}_{{\bf H}^{2.5+\delta}(\mathcal{O})}\int_{t-h}^{t}(\|{\bf u}+{\bf w}\|_{{\bf H}^{1}(\mathcal{O})}+\|{\bf w}_{M}\|_{{\bf H}^{1}(\mathcal{O})}+|\mathsf{c}_{M}(s,t)|)ds\right)
h12𝜼L(0,T0;𝐇2+δ(Γ))4(𝐮L2(0,T0;𝐇1(𝒪))+𝐰L2(0,T0;𝐇1(𝒪)))\displaystyle\leq h^{\frac{1}{2}}\|{\boldsymbol{\eta}}\|^{4}_{L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma))}(\|{\bf u}\|_{L^{2}(0,T_{0};{\bf H}^{1}(\mathcal{O}))}+\|{\bf w}\|_{L^{2}(0,T_{0};{\bf H}^{1}(\mathcal{O}))})
(53) Ch12.\displaystyle\leq Ch^{\frac{1}{2}}.

Next, we readily observe that,

𝐪|Γ=𝝍, on (0,T)×Γ.{\bf q}|_{\Gamma}=\boldsymbol{\psi},\qquad\text{ on }(0,T)\times\Gamma.

Moreover, thanks to Theorem 1.7-1 in [6], we have that

(54) div𝜼(t)𝐪M(t)=J𝜼(t)1thtdiv(Ps𝐮(s))=J𝜼(t)1thtJ𝜼(s) div𝜼(s)𝐮(s)=0.\begin{split}&\text{div}^{{\boldsymbol{\eta}}(t)}{\bf q}_{M}(t)={J_{{\boldsymbol{\eta}}}(t)}^{-1}\int_{t-h}^{t}\text{div}(P_{s}{\bf u}(s))={J_{{\boldsymbol{\eta}}}(t)}^{-1}\int_{t-h}^{t}{J_{\boldsymbol{\eta}}(s)\text{ div}^{{\boldsymbol{\eta}}(s)}{\bf u}(s)}=0.\end{split}

Hence (𝐪M,𝝍M)({\bf q}_{M},\boldsymbol{\psi}_{M}) is a valid test function for (LABEL:weaksol). For this pair of test functions we get:

(55) hT0𝒪J𝜼𝐮t(tht𝐮M)hT0Γt𝜼(ttht𝐯M)=hT0ΓΔ𝜼Δ𝝍M+γhT0ΓΛ1+st𝜼:Λ1+s𝝍MhT0𝒪J𝜼(𝐮𝜼𝐮𝐪𝐰𝜼𝜼𝐮𝐪M)+hT0𝒪tJ𝜼𝐮𝐪M2νhT0𝒪J𝜼𝐃𝜼(𝐮):𝐃𝜼(𝐪M)=I1++I5.\begin{split}&-\int_{h}^{T_{0}}\int_{\mathcal{O}}{J}_{{\boldsymbol{\eta}}}{\bf u}\cdot\partial_{t}\left(\int_{t-h}^{t}{{\bf u}_{M}}\right)-\int_{h}^{T_{0}}\int_{\Gamma}\partial_{t}{{\boldsymbol{\eta}}}\left(\partial_{t}\int_{t-h}^{t}{\bf v}_{M}\right)=\int_{h}^{T_{0}}\int_{\Gamma}\Delta{\boldsymbol{\eta}}\cdot\Delta\boldsymbol{\psi}_{M}\\ &+\gamma\int_{h}^{T_{0}}\int_{\Gamma}\Lambda^{1+s}\partial_{t}{\boldsymbol{\eta}}:\Lambda^{1+s}\boldsymbol{\psi}_{M}-\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{{\boldsymbol{\eta}}}({\bf u}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q}-{\bf w}^{{\boldsymbol{\eta}}}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q}_{M})\\ &+\int_{h}^{T_{0}}\int_{\mathcal{O}}\partial_{t}J_{\boldsymbol{\eta}}{\bf u}\cdot{\bf q}_{M}-2\nu\int_{h}^{T_{0}}\int_{\mathcal{O}}{J}_{{\boldsymbol{\eta}}}\,{\bf D}^{{{\boldsymbol{\eta}}}}({\bf u}):{\bf D}^{{{\boldsymbol{\eta}}}}({\bf q}_{M})\\ &=I_{1}+...+I_{5}.\end{split}

Before estimating each term on the right-hand side of the equation above, we observe that the first term on the left-hand side can be written as,

hT0𝒪J𝜼𝐮\displaystyle\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}{\bf u} t(tht𝐮M)=hT0𝒪J𝜼𝐮(𝐮M(t,t)𝐮M(th,t)+thtt𝐮M(s,t)ds)\displaystyle\cdot\partial_{t}\left(\int_{t-h}^{t}{{\bf u}_{M}}\right)=\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}{\bf u}\cdot\left({\bf u}_{M}(t,t)-{\bf u}_{M}(t-h,t)+\int_{t-h}^{t}\partial_{t}{\bf u}_{M}(s,t)ds\right)
=hT0𝒪J𝜼𝐮(𝐮(t)𝐮(th))\displaystyle=\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}{\bf u}\cdot({\bf u}(t)-{\bf u}(t-h))
+hT0𝒪J𝜼𝐮(𝐮M(t,t)𝐮(t)𝐮M(th,t)+𝐮(th)+thtt𝐮M(s,t)ds)\displaystyle+\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}{\bf u}\cdot\left({\bf u}_{M}(t,t)-{\bf u}(t)-{\bf u}_{M}(t-h,t)+{\bf u}(t-h)+\int_{t-h}^{t}\partial_{t}{\bf u}_{M}(s,t)ds\right)
=I01+I02+I03.\displaystyle=I_{0}^{1}+I_{0}^{2}+I_{0}^{3}.

Observe that the term I01I_{0}^{1} on the right-hand side gives us the desired term τh𝐮𝐮L2(0,T;𝐋2(𝒪))\|\tau_{h}{\bf u}-{\bf u}\|_{L^{2}(0,T;{\bf L}^{2}(\mathcal{O}))} in the left-hand side of (48) since we can write

I01\displaystyle I_{0}^{1} =hT0𝒪J𝜼(t)𝐮(t)(𝐮(t)𝐮(th))\displaystyle=\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}(t){\bf u}(t)\cdot({\bf u}(t)-{\bf u}(t-h))
=12hT0𝒪J𝜼(|𝐮(t)|2|𝐮(th)|2+|𝐮(t)𝐮(th)|2).\displaystyle=\frac{1}{2}\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}\left(|{\bf u}(t)|^{2}-|{\bf u}(t-h)|^{2}+|{\bf u}(t)-{\bf u}(t-h)|^{2}\right).

The second term on the right side in the expression for I01I_{0}^{1} above is the desired term, whereas the first one can be bounded as follows:

hT0𝒪J𝜼\displaystyle\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}} (|𝐮(t)|2|𝐮(th)|2)=hT0𝒪J𝜼(t)|𝐮(t)|2J𝜼(th)|𝐮(th)|2\displaystyle\left(|{\bf u}(t)|^{2}-|{\bf u}(t-h)|^{2}\right)=\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}(t)|{\bf u}(t)|^{2}-J_{\boldsymbol{\eta}}(t-h)|{\bf u}(t-h)|^{2}
hT0𝒪(J𝜼(t)J𝜼(th))|𝐮(th)|2\displaystyle-\int_{h}^{T_{0}}\int_{\mathcal{O}}(J_{\boldsymbol{\eta}}(t)-J_{\boldsymbol{\eta}}(t-h))|{\bf u}(t-h)|^{2}
h(sup0tT0𝒪J𝜼(t)|𝐮(t)|2+tJ𝜼L(0,T0;𝐋2(𝒪))𝐮L2(0,T0;𝐋4(𝒪))2)\displaystyle\leq h\left(\sup_{0\leq t\leq T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}(t)|{\bf u}(t)|^{2}+\|\partial_{t}J_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf L}^{2}(\mathcal{O}))}\|{\bf u}\|^{2}_{L^{2}(0,T_{0};{\bf L}^{4}(\mathcal{O}))}\right)
Ch.\displaystyle\leq Ch.

Before analyzing I02I_{0}^{2} we treat the third term I03I_{0}^{3}. We observe that for any t(0,T0)t\in(0,T_{0}) we have

t𝐮M(s,t)𝐋32(𝒪)\displaystyle\|\partial_{t}{\bf u}_{M}(s,t)\|_{{\bf L}^{\frac{3}{2}}(\mathcal{O})} tPt1Ps(𝐮(s)𝐰(s))𝐋32(𝒪)+supthstt𝖼M(s,t)𝝃0(s)χ𝐋2(𝒪)\displaystyle\leq\|\partial_{t}P_{t}^{-1}P_{s}({\bf u}(s)-{\bf w}(s))\|_{{\bf L}^{\frac{3}{2}}(\mathcal{O})}+\sup_{t-h\leq s\leq t}\partial_{t}\mathsf{c}_{M}(s,t)\|\boldsymbol{\xi}_{0}(s)\chi\|_{{\bf L}^{2}(\mathcal{O})}
C(Ps𝐋(𝒪)tPt𝐋2(𝒪)𝐮(s)𝐰(s)𝐋6(𝒪)+𝐯(t)𝐇1(Γ)𝐯M(s)𝐋2(Γ)).\displaystyle\leq C(\|P_{s}\|_{{\bf L}^{\infty}(\mathcal{O})}\|\partial_{t}P_{t}\|_{{\bf L}^{2}(\mathcal{O})}\|{\bf u}(s)-{\bf w}(s)\|_{{\bf L}^{6}(\mathcal{O})}+\|{\bf v}(t)\|_{{\bf H}^{1}(\Gamma)}\|{\bf v}_{M}(s)\|_{{\bf L}^{2}(\Gamma)}).

Hence, we arrive at the following estimate for the term I03I_{0}^{3},

|I03|\displaystyle|I_{0}^{3}| =|hT0𝒪J𝜼𝐮(thtt𝐮M(s,t)ds)𝑑𝐱𝑑t|\displaystyle=|\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}{\bf u}\left(\int_{t-h}^{t}\partial_{t}{\bf u}_{M}(s,t)ds\right)d{\bf x}dt|
h12hT0(J𝜼(t)𝐋6(𝒪)𝐮(t)𝐋3(𝒪)(thtt𝐮M(s,t)𝐋32(𝒪)2𝑑s)12)𝑑t\displaystyle\leq h^{\frac{1}{2}}\int_{h}^{T_{0}}\left(\|J_{\boldsymbol{\eta}}(t)\|_{{\bf L}^{6}(\mathcal{O})}\|{\bf u}(t)\|_{{\bf L}^{3}(\mathcal{O})}\left(\int_{t-h}^{t}\|\partial_{t}{\bf u}_{M}(s,t)\|^{2}_{{\bf L}^{\frac{3}{2}}(\mathcal{O})}ds\right)^{\frac{1}{2}}\right)dt
Ch12.\displaystyle\leq Ch^{\frac{1}{2}}.

Now we begin with our calculations for the term I02I_{0}^{2}. First note that

𝐰M𝐰𝐋2(𝒪)C𝐯M𝐯𝐇12(Γ)CλM34𝐯𝐇1(Γ).\displaystyle\|{\bf w}_{M}-{\bf w}\|_{{\bf L}^{2}(\mathcal{O})}\leq C\|{\bf v}_{M}-{\bf v}\|_{{\bf H}^{-\frac{1}{2}}(\Gamma)}\leq C\lambda_{M}^{-\frac{3}{4}}\|{\bf v}\|_{{\bf H}^{1}(\Gamma)}.

Moreover, thanks to the embedding H12(𝒪)L3(𝒪)H^{\frac{1}{2}}(\mathcal{O})\hookrightarrow L^{3}(\mathcal{O}) and the fact that the ALE maps satisfy (13), the estimate (51) gives us,

supthstPsPt𝐋3(𝒪)\displaystyle\sup_{t-h\leq s\leq t}\|P_{s}-P_{t}\|_{{\bf L}^{3}(\mathcal{O})} 𝜼L(0,T0;𝐇2+δ(Γ))supthst𝜼(s)𝜼(t)𝐇1(Γ)\displaystyle\leq\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma))}\sup_{t-h\leq s\leq t}\|{\boldsymbol{\eta}}(s)-{\boldsymbol{\eta}}(t)\|_{{\bf H}^{1}(\Gamma)}
h12𝜼C0,12(0,T0;𝐇1(Γ))\displaystyle\leq h^{\frac{1}{2}}\|{\boldsymbol{\eta}}\|_{C^{0,\frac{1}{2}}(0,T_{0};{\bf H}^{1}(\Gamma))}
Ch12.\displaystyle\leq Ch^{\frac{1}{2}}.

These two estimates put together give us,

hT0\displaystyle\int_{h}^{T_{0}} 𝐮M(th,t)𝐮(th)𝐋2(𝒪)2dt\displaystyle\|{\bf u}_{M}(t-h,t)-{\bf u}(t-h)\|^{2}_{{\bf L}^{2}(\mathcal{O})}dt
hT0(Pt1𝐋(𝒪)PthPt𝐋3(𝒪)𝐮𝐰𝐋6(𝒪)+𝐰𝐰M𝐋2(𝒪)+|𝖼M(th,t)|)2𝑑t\displaystyle\leq\int_{h}^{T_{0}}(\|P^{-1}_{t}\|_{{\bf L}^{\infty}(\mathcal{O})}\|P_{t-h}-P_{t}\|_{{\bf L}^{3}(\mathcal{O})}\|{\bf u}-{\bf w}\|_{{\bf L}^{6}(\mathcal{O})}+\|{\bf w}-{\bf w}_{M}\|_{{\bf L}^{2}(\mathcal{O})}+|\mathsf{c}_{M}(t-h,t)|)^{2}dt
Ch.\displaystyle\leq Ch.

Hence, for the term I02I_{0}^{2} we find,

|I02|\displaystyle|I_{0}^{2}| =|hT0𝒪J𝜼𝐮(𝐮M(t,t)𝐮(t)𝐮M(th,t)+𝐮(th))𝑑x𝑑t|\displaystyle=|\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{\boldsymbol{\eta}}{\bf u}\cdot({\bf u}_{M}(t,t)-{\bf u}(t)-{\bf u}_{M}(t-h,t)+{\bf u}(t-h))dxdt|
J𝜼L(0,T0;𝐋p(𝒪))𝐮L2(0,T0;𝐇1(𝒪))(hT0𝐮M(th,t)𝐮(t)𝐋2(𝒪)2𝑑t)12\displaystyle\leq\|J_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf L}^{p}(\mathcal{O}))}\|{\bf u}\|_{L^{2}(0,T_{0};{\bf H}^{1}(\mathcal{O}))}(\int_{h}^{T_{0}}\|{\bf u}_{M}(t-h,t)-{\bf u}(t)\|^{2}_{{\bf L}^{2}(\mathcal{O})}dt)^{\frac{1}{2}}
Ch12.\displaystyle\leq Ch^{\frac{1}{2}}.

Similarly, we treat the second term on the left-hand side of (55) which will produce the second term on the left-hand side of our desired inequality (48). We write,

hT0Γ𝐯(t)t(tht𝐯M(s)𝑑s)dt\displaystyle\int_{h}^{T_{0}}\int_{\Gamma}{\bf v}(t)\,\partial_{t}\left(\int_{t-h}^{t}{{\bf v}_{M}}(s)ds\right)dt =hT0Γ𝐯(t)(𝐯M(t)𝐯M(th)(𝖼M(t,t)𝖼M(th,t))𝝃0)\displaystyle=\int_{h}^{T_{0}}\int_{\Gamma}{\bf v}(t)({\bf v}_{M}(t)-{\bf v}_{M}(t-h)-(\mathsf{c}_{M}(t,t)-\mathsf{c}_{M}(t-h,t))\boldsymbol{\xi}_{0})
=J01+J02.\displaystyle=J_{0}^{1}+J_{0}^{2}.

Observe that due to orthonormality of PMP_{M} we have

J01\displaystyle J_{0}^{1} =hT0Γ𝐯(t)(𝐯M(t)𝐯M(th))𝑑x𝑑y𝑑t\displaystyle=\int_{h}^{T_{0}}\int_{\Gamma}{\bf v}(t)({\bf v}_{M}(t)-{\bf v}_{M}(t-h))dxdydt
=hT0Γ𝐯M(t)(𝐯M(t)𝐯M(th))𝑑x𝑑y𝑑t\displaystyle=\int_{h}^{T_{0}}\int_{\Gamma}{\bf v}_{M}(t)({\bf v}_{M}(t)-{\bf v}_{M}(t-h))dxdydt
=12hT0Γ(|𝐯M(t)|2|𝐯M(th)|2+|𝐯M(t)𝐯M(th)|2)𝑑x𝑑y𝑑t.\displaystyle=\frac{1}{2}\int_{h}^{T_{0}}\int_{\Gamma}(|{\bf v}_{M}(t)|^{2}-|{\bf v}_{M}(t-h)|^{2}+|{\bf v}_{M}(t)-{\bf v}_{M}(t-h)|^{2})dxdydt.

Here, as mentioned previously, the second term is another one of our desired terms in (48) whereas the first term can be bounded as follows,

hT0(𝐯M(t)𝐋2(Γ)2𝐯M(th)𝐋2(Γ)2)𝑑t=T0hT0𝐯M(t)𝐋2(Γ)2𝑑th𝐯L(0,T0;𝐋2(Γ))2.\int_{h}^{T_{0}}(\|{\bf v}_{M}(t)\|_{{\bf L}^{2}(\Gamma)}^{2}-\|{\bf v}_{M}(t-h)\|_{{\bf L}^{2}(\Gamma)}^{2})dt=\int_{T_{0}-h}^{T_{0}}\|{\bf v}_{M}(t)\|_{{\bf L}^{2}(\Gamma)}^{2}dt\leq h\|{\bf v}\|^{2}_{L^{\infty}(0,T_{0};{\bf L}^{2}(\Gamma))}.

Notice that here we also used the property of projection that states PM𝐯𝐇kγmk2𝐯𝐇m\|P_{M}{\bf v}\|_{{\bf H}^{k}}\leq\gamma^{\frac{m-k}{2}}\|{\bf v}\|_{{\bf H}^{m}}. Thanks to (52), we readily further deduce that

|J02|=|hT0Γ(𝖼M(t,t)𝖼M(th,t))𝝃0|Ch12.\displaystyle|J_{0}^{2}|=|\int_{h}^{T_{0}}\int_{\Gamma}(\mathsf{c}_{M}(t,t)-\mathsf{c}_{M}(t-h,t))\boldsymbol{\xi}_{0}|\leq Ch^{\frac{1}{2}}.

This completes the treatment of the terms on the left-hand side of (55). Hence, by combining all the estimates, we summarize that so far we have

τh𝐮𝐮L2(0,T0;𝐋2(𝒪))+τht𝜼t𝜼L2(0,T0;𝐋2(Γ))Ch12+j=18|Ij|\displaystyle\|\tau_{h}{\bf u}-{\bf u}\|_{L^{2}(0,T_{0};{\bf L}^{2}(\mathcal{O}))}+\|\tau_{h}\partial_{t}{\boldsymbol{\eta}}-\partial_{t}{\boldsymbol{\eta}}\|_{L^{2}(0,T_{0};{\bf L}^{2}(\Gamma))}\leq Ch^{\frac{1}{2}}+\sum_{j=1}^{8}|I_{j}|

Now we estimate the terms Ij,j=1,,5I_{j},j=1,...,5 that appear on the right-hand side of (55). We start with I1I_{1}, which is one of the more crucial terms. Thanks to the property of projection operators stating PM𝐯𝐇kγmk2𝐯𝐇m\|P_{M}{\bf v}\|_{{\bf H}^{k}}\leq\gamma^{\frac{m-k}{2}}\|{\bf v}\|_{{\bf H}^{m}} and the estimate (52), we obtain

|I1|\displaystyle|I_{1}| =|hT0ΓΔ𝜼Δ𝝍|=hT0ΓΔ𝜼Δ(tht(𝐯M(s)𝖼M(s,t)𝝃0(s))𝑑s)\displaystyle=|\int_{h}^{T_{0}}\int_{\Gamma}\Delta{\boldsymbol{\eta}}\cdot\Delta\boldsymbol{\psi}|=\int_{h}^{T_{0}}\int_{\Gamma}\Delta{\boldsymbol{\eta}}\cdot\Delta\left(\int_{t-h}^{t}\left({\bf v}_{M}(s)-\mathsf{c}_{M}(s,t)\boldsymbol{\xi}_{0}(s)\right)ds\right)
hT0tht𝜼(t)𝐇2(Γ)(𝐯M(s)𝐇2(Γ)+|𝖼M(s,t)|𝝃0(s)𝐇2(Γ))𝑑s𝑑t\displaystyle\leq\int_{h}^{T_{0}}\int_{t-h}^{t}\|{\boldsymbol{\eta}}(t)\|_{{\bf H}^{2}(\Gamma)}\left(\|{\bf v}_{M}(s)\|_{{\bf H}^{2}(\Gamma)}+|\mathsf{c}_{M}(s,t)|\|\boldsymbol{\xi}_{0}(s)\|_{{\bf H}^{2}(\Gamma)}\right)dsdt
hT0tht𝜼(t)𝐇2(Γ)(λM𝐯(s)𝐋2(Γ)+|𝖼M(s,t)|𝝃0(s)𝐇2(Γ))𝑑s𝑑t\displaystyle\leq\int_{h}^{T_{0}}\int_{t-h}^{t}\|{\boldsymbol{\eta}}(t)\|_{{\bf H}^{2}(\Gamma)}\left(\lambda_{M}\|{\bf v}(s)\|_{{\bf L}^{2}(\Gamma)}+|\mathsf{c}_{M}(s,t)|\|\boldsymbol{\xi}_{0}(s)\|_{{\bf H}^{2}(\Gamma)}\right)dsdt
Ch𝜼L(0,T0;𝐇2(Γ))(λM𝐯L(0,T0;𝐋2(Γ))+h12)Ch14.\displaystyle\leq Ch\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf H}^{2}(\Gamma))}(\lambda_{M}\|{\bf v}\|_{L^{\infty}(0,T_{0};{\bf L}^{2}(\Gamma))}+h^{\frac{1}{2}})\leq Ch^{\frac{1}{4}}.

Similarly, for the next term we obtain

|I2|=|γhT0ΓΛ1+st𝜼:Λ1+s𝝍|h12𝐯L2(0,T;𝐇1+s(Γ))Ch12.\displaystyle|I_{2}|=|\gamma\int_{h}^{T_{0}}\int_{\Gamma}\Lambda^{1+s}\partial_{t}{\boldsymbol{\eta}}:\Lambda^{1+s}\boldsymbol{\psi}|\leq h^{\frac{1}{2}}\|{\bf v}\|_{L^{2}(0,T;{\bf H}^{1+s}(\Gamma))}\leq Ch^{\frac{1}{2}}.

The next two terms I3,I4I_{3},I_{4} are treated using (53). For the nonlinear term we have

|I3|\displaystyle|I_{3}| =|hT0𝒪J𝜼(𝐮𝜼𝐮𝐪𝐰𝜼𝜼𝐮𝐪)|\displaystyle=|\int_{h}^{T_{0}}\int_{\mathcal{O}}J_{{\boldsymbol{\eta}}}({\bf u}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q}-{\bf w}^{{\boldsymbol{\eta}}}\cdot\nabla^{{\boldsymbol{\eta}}}{\bf u}\cdot{\bf q})|
CJ𝜼L(0,T0;𝐋6(𝒪))𝐮L2(0,T0;𝐋6(𝒪))𝐮L2(0,T0;𝐇1(𝒪))𝐪L(0,T0;𝐋6(𝒪))\displaystyle\leq C\|J_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf L}^{6}(\mathcal{O}))}\|{\bf u}\|_{L^{2}(0,T_{0};{\bf L}^{6}(\mathcal{O}))}\|{\bf u}\|_{L^{2}(0,T_{0};{\bf H}^{1}(\mathcal{O}))}\|{\bf q}\|_{L^{\infty}(0,T_{0};{\bf L}^{6}(\mathcal{O}))}
C𝜼L(0,T0;𝐇2(𝒪))𝐮L2(0,T0;𝐇1(𝒪))2𝐪L(0,T0;𝐇1(𝒪))\displaystyle\leq C\|{\boldsymbol{\eta}}\|_{L^{\infty}(0,T_{0};{\bf H}^{2}(\mathcal{O}))}\|{\bf u}\|^{2}_{L^{2}(0,T_{0};{\bf H}^{1}(\mathcal{O}))}\|{\bf q}\|_{L^{\infty}(0,T_{0};{\bf H}^{1}(\mathcal{O}))}
Ch12.\displaystyle\leq Ch^{\frac{1}{2}}.

The terms I4I_{4} and I5I_{5} are treated identically. For I4I_{4} we use the fact that 𝐰𝜼{\bf w}^{{\boldsymbol{\eta}}} is the harmonic extension of 𝐯{\bf v} in 𝒪\mathcal{O} which implies that t𝐀𝜼𝐇k+12(𝒪)C𝐯𝐇k(Γ)\|\partial_{t}{\bf A}_{\boldsymbol{\eta}}\|_{{\bf H}^{k+\frac{1}{2}}(\mathcal{O})}\leq C\|{\bf v}\|_{{\bf H}^{k}(\Gamma)}. Hence (53) leads to

|I4|=|hT0𝒪tJ𝜼𝐮𝐪|tJ𝜼L2(0,T;𝐋2(𝒪))𝐮L2(0,T;𝐇1(𝒪))𝐪L(0,T;𝐇1(𝒪))Ch12.\displaystyle|I_{4}|=|\int_{h}^{T_{0}}\int_{\mathcal{O}}\partial_{t}J_{\boldsymbol{\eta}}{\bf u}\cdot{\bf q}|\leq\|\partial_{t}J_{\boldsymbol{\eta}}\|_{L^{2}(0,T;{\bf L}^{2}(\mathcal{O}))}\|{\bf u}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}\|{\bf q}\|_{L^{\infty}(0,T;{\bf H}^{1}(\mathcal{O}))}\leq Ch^{\frac{1}{2}}.

Next, for I5I_{5} we see that,

|I5|\displaystyle|I_{5}| =|hT0𝒪J𝜼𝐃𝜼(𝐮):𝐃𝜼(𝐪)|CJ𝜼L(0,T;𝐋(Γ))𝐃𝜼(𝐮)L2(0,T;𝐋2(𝒪))𝐃𝜼(𝐪)L2(0,T;𝐋2(𝒪))\displaystyle=|\int_{h}^{T_{0}}\int_{\mathcal{O}}{J}_{{\boldsymbol{\eta}}}\,{\bf D}^{{{\boldsymbol{\eta}}}}({\bf u}):{\bf D}^{{{\boldsymbol{\eta}}}}({\bf q})|\leq C\|J_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf L}^{\infty}(\Gamma))}\|{\bf D}^{{{\boldsymbol{\eta}}}}({\bf u})\|_{L^{2}(0,T;{\bf L}^{2}(\mathcal{O}))}\|{\bf D}^{{{\boldsymbol{\eta}}}}({\bf q})\|_{L^{2}(0,T;{\bf L}^{2}(\mathcal{O}))}
CJ𝜼L(0,T;𝐋(Γ))(𝐀𝜼)1L(0,T;𝐋(Γ))2𝐮L2(0,T;𝐋2(𝒪))𝐪L2(0,T;𝐋2(𝒪))\displaystyle\leq C\|J_{\boldsymbol{\eta}}\|_{L^{\infty}(0,T;{\bf L}^{\infty}(\Gamma))}\|(\nabla{\bf A}_{\boldsymbol{\eta}})^{-1}\|^{2}_{L^{\infty}(0,T;{\bf L}^{\infty}(\Gamma))}\|\nabla{\bf u}\|_{L^{2}(0,T;{\bf L}^{2}(\mathcal{O}))}\|\nabla{\bf q}\|_{L^{2}(0,T;{\bf L}^{2}(\mathcal{O}))}
C𝜼L(0,T;𝐇2+δ(Γ))3𝐮L2(0,T;𝐇1(𝒪))𝐪L2(0,T;𝐇2(𝒪))Ch12.\displaystyle\leq C\|{\boldsymbol{\eta}}\|^{3}_{L^{\infty}(0,T;{\bf H}^{2+\delta}(\Gamma))}\|{\bf u}\|_{L^{2}(0,T;{\bf H}^{1}(\mathcal{O}))}\|{\bf q}\|_{L^{2}(0,T;{\bf H}^{2}(\mathcal{O}))}\leq Ch^{\frac{1}{2}}.

Finally, we collect all the terms and obtain that,

τh𝐮𝐮L2(h,T0;𝐋2(𝒪))+τh𝐯M𝐯ML2(h,T0;𝐋2(Γ))Ch14.\displaystyle\|\tau_{h}{\bf u}-{\bf u}\|_{L^{2}(h,T_{0};{\bf L}^{2}(\mathcal{O}))}+\|\tau_{h}{\bf v}_{M}-{\bf v}_{M}\|_{L^{2}(h,T_{0};{\bf L}^{2}(\Gamma))}\leq Ch^{\frac{1}{4}}.

Moreover, due to the following property of the projection PMP_{M}

𝐯PM𝐯L2(0,T;𝐋2(Γ))CλM12𝐯L2(0,T;𝐇1(Γ))Ch38𝐯L2(0,T;𝐇1(Γ)),\|{\bf v}-P_{M}{\bf v}\|_{L^{2}(0,T;{\bf L}^{2}(\Gamma))}\leq C\lambda_{M}^{-\frac{1}{2}}\|{\bf v}\|_{L^{2}(0,T;{\bf H}^{1}(\Gamma))}\leq Ch^{\frac{3}{8}}\|{\bf v}\|_{L^{2}(0,T;{\bf H}^{1}(\Gamma))},

we come to our desired result,

τh𝐮𝐮L2(h,T0;𝐋2(𝒪))2+τh𝐯𝐯L2(h,T0;𝐋2(Γ))2Ch14.\displaystyle\|\tau_{h}{\bf u}-{\bf u}\|^{2}_{L^{2}(h,T_{0};{\bf L}^{2}(\mathcal{O}))}+\|\tau_{h}{\bf v}-{\bf v}\|^{2}_{L^{2}(h,T_{0};{\bf L}^{2}(\Gamma))}\leq Ch^{\frac{1}{4}}.

This completes the proof of Theorem 2.

4. Existence result

In this section we will provide a brief proof of Theorem 3, namely, we prove the existence of a weak solution to our FSI problem. We will work with the problem posed on the fixed domain, and consider a weak formulation of the problem in the sense of Definition 2. The plan is to discretize the problem in time and construct approximate solutions by employing the Lie operator splitting strategy to decouple the FSI problem into a fluid and a structure subproblem. This is done in the spirit of [22] (see also [23, 34]). For the convenience of the reader we will reproduce all the important steps in this section and for complementary details we refer to [35]. We introduce the splitting scheme in the next subsection and then discuss the strategy for the rest of our proof in subsequent subsections.

4.1. The Lie operator splitting scheme

The overarching idea behind the Lie operator splitting scheme is to solve the evolution equation dϕdt+S(ϕ)=0\frac{d\phi}{dt}+S(\phi)=0 by splitting the operator SS as a nontrivial sum S=S1+S2S=S_{1}+S_{2}. The time interval (0,T)(0,T) is divided into NN sub-intervals of size Δt\Delta t and on each sub-interval (nΔt,(n+1)Δt)(n\Delta t,(n+1)\Delta t) the evolution equations dϕidt+Si(ϕi)=0\frac{d\phi_{i}}{dt}+S_{i}(\phi_{i})=0, i=1,2i=1,2, are solved. In our case, we semidiscretize the problem in time and use operator splitting to divide the coupled problem along the dynamic coupling condition into two subproblems: a fluid and a structure subproblem. The initial value for the structure sub-problem is taken to be the solution from the previous step, whereas the initial value for the fluid sub-problem is taken to be the just calculated solution found in the first sub-problem.

Our strategy is that in the first (structure) subproblem, we keep fluid velocity the same and update only the structure displacement and structure velocity, while in the second (fluid) subproblem we update the fluid and the structure velocities while keeping the structure displacement the same. The kinematic coupling condition is enforced in the second subproblem.

For any NN\in\mathbb{N}, we denote the time step by Δt=TN\Delta t=\frac{T}{N} and use the notation tn=nΔtt^{n}=n\Delta t for n=0,1,,Nn=0,1,...,N. For any NN\in\mathbb{N} we introduce the following discrete total energy and dissipation for i=0,1i=0,1 and n=0,1,..,N1n=0,1,..,N-1:

(56) ENn+i2=12(𝒪Jn|𝐮n+i2|2𝑑x+𝐯n+i2𝐋2(Γ)2+𝜼n+i2𝐇2(Γ)2+1N𝜼n+i2𝐇3(Γ)2),DNn=Δt𝒪2νJn|𝐃ηn(𝐮n+1)|2𝑑x+𝐯n+1𝐇1+s(Γ)2.\begin{split}E_{N}^{n+\frac{i}{2}}&=\frac{1}{2}\Big{(}\int_{\mathcal{O}}J^{n}|{\bf u}^{n+\frac{i}{2}}|^{2}dx+\|{\bf v}^{n+\frac{i}{2}}\|^{2}_{{\bf L}^{2}(\Gamma)}+\|{\boldsymbol{\eta}}^{n+\frac{i}{2}}\|^{2}_{{\bf H}^{2}(\Gamma)}+\frac{1}{N}\|{\boldsymbol{\eta}}^{n+\frac{i}{2}}\|^{2}_{{\bf H}^{3}(\Gamma)}\Big{)},\\ D_{N}^{n}&=\Delta t\int_{\mathcal{O}}2\nu J^{n}|{\bf D}^{\eta^{n}}({\bf u}^{n+1})|^{2}dx+\|{\bf v}^{n+1}\|^{2}_{{\bf H}^{1+s}(\Gamma)}.\end{split}

The splitting scheme consisting of the two subproblems is defined as follows. Let (𝐮0,𝜼0,𝐯0)=(𝐮0,𝜼0,𝐯0)({\bf u}^{0},{\boldsymbol{\eta}}^{0},{\bf v}^{0})=({\bf u}_{0},{\boldsymbol{\eta}}_{0},{\bf v}_{0}) be the initial data. Then at the jthj^{th} time level, we update the vector (𝐮n+j2,𝜼n+j2,𝐯n+j2)({\bf u}^{n+\frac{j}{2}},{\boldsymbol{\eta}}^{n+\frac{j}{2}},{\bf v}^{n+\frac{j}{2}}), where j=1,2j=1,2 and n=0,1,2,,N1n=0,1,2,...,N-1, according to the following scheme.

Structure sub-problem: For any nNn\leq N we look for (𝜼n+12,𝐯n+12)({\boldsymbol{\eta}}^{n+\frac{1}{2}},{\bf v}^{n+\frac{1}{2}}) such that

(57) 𝐮n+12=𝐮n,Γ(𝜼n+12𝜼n)ϕ=(Δt)Γ𝐯n+12ϕ,Γ(𝐯n+12𝐯n)𝝍+(Δt)ΓΔ𝜼n+12Δ𝝍+(Δt)NΓ3𝜼n+123𝝍=0,\begin{split}{\bf u}^{n+\frac{1}{2}}&={\bf u}^{n},\\ \int_{\Gamma}({\boldsymbol{\eta}}^{n+\frac{1}{2}}-{\boldsymbol{\eta}}^{n})\boldsymbol{\phi}&=(\Delta t)\int_{\Gamma}{\bf v}^{n+\frac{1}{2}}\boldsymbol{\phi},\\ \int_{\Gamma}\left({\bf v}^{n+\frac{1}{2}}-{\bf v}^{n}\right)\boldsymbol{\psi}&+(\Delta t)\int_{\Gamma}\Delta{\boldsymbol{\eta}}^{n+\frac{1}{2}}\cdot\Delta\boldsymbol{\psi}+\frac{(\Delta t)}{N}\int_{\Gamma}\nabla^{3}{\boldsymbol{\eta}}^{n+\frac{1}{2}}\cdot\nabla^{3}\boldsymbol{\psi}=0,\end{split}

for any ϕ𝐋2(Γ)\boldsymbol{\phi}\in{\bf L}^{2}(\Gamma) and 𝝍𝐇3(Γ)𝒱S\boldsymbol{\psi}\in{\bf H}^{3}(\Gamma)\cap\mathscr{V}_{S}.

Remark 6.

We notice that we have augmented this subproblem with a regularizing term, which is the last term on the left hand-side of the third equation (the elastodynamics equation) in (57). This term will vanish when we pass NN\to\infty. The presence of this term is attributed to the fact that splitting the FSI problem along the dynamic coupling condition causes a mismatch between the structure velocity and the trace of the fluid velocity on Γ\Gamma (see also the second subproblem (LABEL:second)). This is not ideal for an application of the regularity result of Theorem 1 at the level of approximate formulations (see Theorem 4.3 below). This mismatch is taken care of by using bounds on numerical dissipation that result from the addition of this regularization term (see the discussion following (LABEL:mismatch)).

Before commenting on the existence of a solution to the structure subproblem (57), we introduce the second subproblem that updates the fluid and structure velocities in the fixed domain formulation (LABEL:weaksol).

Fluid sub-problem: Introduce the following functional space for the fluid velocity

𝒱n:={𝐮𝐇1(𝒪):𝜼n𝐮=0 on 𝒪,𝐮=0 on Γr}.\mathscr{V}^{n}:=\{{\bf u}\in{\bf H}^{1}(\mathcal{O}):\nabla^{{\boldsymbol{\eta}}^{n}}\cdot{\bf u}=0\text{ on }\mathcal{O},{\bf u}=0\text{ on }\Gamma_{r}\}.

We look for (𝐮n+1,𝐯n+1)𝒱n×𝐇1+s(Γ)({\bf u}^{n+1},{\bf v}^{n+1})\in\mathscr{V}^{n}\times{\bf H}^{1+s}(\Gamma) such that the following equations are satisfied for any (𝐪,𝝍)𝒱n×𝐇1+s(Γ)({\bf q},\boldsymbol{\psi})\in\mathscr{V}^{n}\times{\bf H}^{1+s}(\Gamma) such that 𝐪|Γ=𝝍{\bf q}|_{\Gamma}=\boldsymbol{\psi}:

(58) 𝜼n+1:=𝜼n+12,𝒪Jn(𝐮n+1𝐮n+12)𝐪+12𝒪(Jn+1Jn)𝐮n+1𝐪+12(Δt)𝒪Jn((𝐮n𝐰n+1)𝜼n𝐮n+1𝐪(𝐮n𝐰n+1)𝜼n𝐪𝐮n+1)+2ν(Δt)𝒪Jn𝐃𝜼n(𝐮n+1)𝐃𝜼n(𝐪)+Γ(𝐯n+1𝐯n+12)𝝍+(Δt)ΓΛ1+s𝐯n+1Λ1+s𝝍=0,\begin{split}&\qquad\qquad{{\boldsymbol{\eta}}}^{n+1}:={{\boldsymbol{\eta}}}^{n+\frac{1}{2}},\\ &\int_{\mathcal{O}}J^{n}\left({\bf u}^{n+1}-{\bf u}^{n+{\frac{1}{2}}}\right){\bf q}+\frac{1}{2}\int_{\mathcal{O}}\left(J^{n+1}-J^{n}\right){\bf u}^{n+1}\cdot{\bf q}\\ &+\frac{1}{2}(\Delta t)\int_{\mathcal{O}}J^{n}(({\bf u}^{n}-{\bf w}^{{n+1}})\cdot\nabla^{{{\boldsymbol{\eta}}}^{n}}{\bf u}^{n+1}\cdot{\bf q}-({\bf u}^{n}-{\bf w}^{{n+1}})\cdot\nabla^{{{\boldsymbol{\eta}}}^{n}}{\bf q}\cdot{\bf u}^{n+1})\\ &+2\nu(\Delta t)\int_{\mathcal{O}}J^{n}{\bf D}^{{{\boldsymbol{\eta}}}^{n}}({\bf u}^{n+1})\cdot{\bf D}^{{{\boldsymbol{\eta}}}^{n}}({\bf q})+\int_{\Gamma}({\bf v}^{n+1}-{\bf v}^{n+\frac{1}{2}})\boldsymbol{\psi}\\ &+(\Delta t)\int_{\Gamma}\Lambda^{1+s}{\bf v}^{n+1}\cdot\Lambda^{1+s}\boldsymbol{\psi}=0,\end{split}

and the kinematic coupling condition is satisfied

𝐮n+1|Γ=𝐯n+1.{\bf u}^{n+1}|_{\Gamma}={\bf v}^{n+1}.

Here, we use the notation

𝐰n=1Δt(𝐀𝜼n+1𝐀𝜼n),Jn=det𝐀𝜼n,{\bf w}^{n}=\frac{1}{\Delta t}({\bf A}_{{{\boldsymbol{\eta}}}^{n+1}}-{\bf A}_{{{\boldsymbol{\eta}}}^{n}}),\quad J^{n}=\text{det}\nabla{\bf A}_{{{\boldsymbol{\eta}}}^{n}},

where 𝐀𝜼{\bf A}_{{\boldsymbol{\eta}}} denotes the solution to (10) corresponding to the boundary data 𝐢𝐝+𝜼{\bf id}+{\boldsymbol{\eta}}.

Equations (57) and (LABEL:second) define the two steps in our splitting scheme.

Next we discuss the existence of unique solutions for the two subproblems (57) and (LABEL:second).

Theorem 4.1 (Existence and uniqueness result for the subproblems).

The following statements hold true:

  1. (1)

    Given 𝜼n𝐇2(Γ){\boldsymbol{\eta}}^{n}\in{\bf H}^{2}(\Gamma) and 𝐯n𝐋2(Γ){\bf v}^{n}\in{\bf L}^{2}(\Gamma) there exist unique 𝜼n+12,𝐯n+12𝐇2(Γ){\boldsymbol{\eta}}^{n+\frac{1}{2}},{\bf v}^{n+\frac{1}{2}}\in{\bf H}^{2}(\Gamma) that solve (57), and the following semidiscrete energy inequality holds:

    (59) En+12+12𝐯n+12𝐯n𝐋2(Γ)2+12𝜼n+12𝜼n𝐇2(Γ)2+12N3𝜼n+123𝜼n𝐋2(Γ)2En.\begin{split}E^{n+\frac{1}{2}}+\frac{1}{2}\|{\bf v}^{n+\frac{1}{2}}-{\bf v}^{n}\|_{{\bf L}^{2}(\Gamma)}^{2}+\frac{1}{2}\|{\boldsymbol{\eta}}^{n+\frac{1}{2}}-{\boldsymbol{\eta}}^{n}\|_{{\bf H}^{2}(\Gamma)}^{2}+\frac{1}{2N}\|\nabla^{3}{\boldsymbol{\eta}}^{n+\frac{1}{2}}-\nabla^{3}{\boldsymbol{\eta}}^{n}\|_{{\bf L}^{2}(\Gamma)}^{2}\leq E^{n}.\end{split}
  2. (2)

    Given (𝐮n+12,𝐯n)𝒱n1×𝐇1+s(Γ)({\bf u}^{n+\frac{1}{2}},{\bf v}^{n})\in\mathscr{V}^{n-1}\times{\bf H}^{1+s}(\Gamma) and 𝐯n+12𝐇2(Γ){\bf v}^{n+\frac{1}{2}}\in{\bf H}^{2}(\Gamma), and 𝜼n𝐇2(Γ){\boldsymbol{\eta}}^{n}\in{\bf H}^{2}(\Gamma) assume that inf𝒪Jn>α\inf_{\mathcal{O}}J^{n}>\alpha for some fixed α>0\alpha>0 for every 0nN0\leq n\leq N. Then there exists a unique (𝐮n+1,𝐯n+1)𝒱n×𝐇1+s(Γ)({\bf u}^{n+1},{\bf v}^{n+1})\in\mathscr{V}^{n}\times{\bf H}^{1+s}(\Gamma) that solves (LABEL:second), and the solution satisfies the following energy estimate

    (60) En+1+Dn+12𝒪Jn(|𝐮n+1𝐮n|2)+12Γ|𝐯n+1𝐯n+12|2En+12,\begin{split}E^{n+1}+D^{n}+\frac{1}{2}\int_{\mathcal{O}}J^{n}\left(|{\bf u}^{n+1}-{\bf u}^{n}|^{2}\right)&+\frac{1}{2}\int_{\Gamma}|{\bf v}^{n+1}-{\bf v}^{n+\frac{1}{2}}|^{2}\leq E^{n+\frac{1}{2}},\end{split}

    where EnE_{n} and DnD_{n} are defined in (56).

Proof.

The proof of this theorem involves an application of the Lax-Milgram Lemma in a way similar to the proofs of Propositions 1, 2, 3 and 4 in [22]. ∎

The rest of the proof of Theorem 3 can be divided into 3 parts: Constructing approximate solutions, finding uniform estimates for the approximate solutions and then passing NN\to\infty to prove that the limiting function is the desired solution, which involves a construction of appropriate test functions to be able to pass to the limit. We start with the construction of approximate solutions.

4.2. Approximate solutions

In this subsection we will define two sequences of approximate solutions corresponding to the fluid velocity 𝐮{\bf u}, structure displacement 𝜼{\boldsymbol{\eta}} and the structure velocity 𝐯{\bf v}. First, as is common with time-discretizations, we define the following approximations that are piece-wise constant in time: For t(nΔt,(n+1)Δt]t\in(n\Delta t,(n+1)\Delta t] we let

(61) 𝐮N(t,)=𝐮n+1,𝜼N(t,)=𝜼n+1,𝐯N(t,)=𝐯n+1,𝐯N(t,)=𝐯n+12.\displaystyle{\bf u}_{N}(t,\cdot)={\bf u}^{n+1},\quad{\boldsymbol{\eta}}_{N}(t,\cdot)={\boldsymbol{\eta}}^{n+1},\quad{\bf v}_{N}(t,\cdot)={\bf v}^{n+1},\quad{\bf v}^{*}_{N}(t,\cdot)={\bf v}^{n+\frac{1}{2}}.

Furthermore, we define the corresponding piecewise linear interpolations: for t[tn,tn+1]t\in[t^{n},t^{n+1}] we let

(62) 𝐮~N(t,)=ttnΔt𝐮n+1+tn+1tΔt𝐮n,𝐯~N(t,)=ttnΔt𝐯n+1+tn+1tΔt𝐯n.𝜼~N(t,)=ttnΔt𝜼n+1+tn+1tΔt𝜼n.\begin{split}&\tilde{\bf u}_{N}(t,\cdot)=\frac{t-t^{n}}{\Delta t}{\bf u}^{n+1}+\frac{t^{n+1}-t}{\Delta t}{\bf u}^{n},\quad\tilde{\bf v}_{N}(t,\cdot)=\frac{t-t^{n}}{\Delta t}{\bf v}^{n+1}+\frac{t^{n+1}-t}{\Delta t}{\bf v}^{n}.\\ &\tilde{\boldsymbol{\eta}}_{N}(t,\cdot)=\frac{t-t^{n}}{\Delta t}{\boldsymbol{\eta}}^{n+1}+\frac{t^{n+1}-t}{\Delta t}{\boldsymbol{\eta}}^{n}.\end{split}

Observe that

t𝜼~N=𝐯N.\partial_{t}\tilde{\boldsymbol{\eta}}_{N}={\bf v}^{*}_{N}.

We now define 𝐀𝜼N{\bf A}_{{\boldsymbol{\eta}}_{N}} as the piecewise constant interpolations of the approximate ALE maps 𝐀𝜼n{\bf A}_{{\boldsymbol{\eta}}^{n}}. Observe that, by definition, 𝐀𝜼N{\bf A}_{{\boldsymbol{\eta}}_{N}} solves (10) with boundary value 𝜼N{\boldsymbol{\eta}}_{N} on Γ\Gamma. We denote its Jacobian by JN:=det𝐀𝜼NJ_{N}:=\text{det}\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}}, which by definition is the piecewise constant interpolation of the functions JnJ^{n}. We will also require piecewise linear interpolations of JnJ^{n} which we will denote by J~N\tilde{J}_{N}. Along with that we also define the approximate ALE velocity 𝐰N{\bf w}_{N} to be the piecewise constant interpolation of 𝐰n{\bf w}^{n}. Note that, by definition, 𝐰N{\bf w}_{N} solves (10) with boundary data 𝐯N{\bf v}^{*}_{N}.

Using this notation, we combine the two subproblems (57) and (LABEL:second) and then write the weak formulation satisfied by the approximate solutions in monolithic form as follows:

(63) 0T𝒪(τΔtJN)t𝐮~N𝐪N0TΓt𝐯~N𝝍N=120T𝒪tJ~N𝐮N𝐪N120T𝒪(τΔtJN)((τΔt𝐮N𝐰N)τΔt𝜼N𝐮N𝐪N(τΔt𝐮N𝐰N)τΔt𝜼N𝐪N𝐮N)2ν0T𝒪(τΔtJN)𝐃τΔt𝜼N(𝐮N)𝐃τΔt𝜼N(𝐪N)0TΓΔ𝜼NΔ𝝍N1N0TΓ3𝜼N3𝝍N+γ0TΓΛ1+s𝐯NΛ1+s𝝍N,\begin{split}&-\int_{0}^{T}\int_{\mathcal{O}}(\tau_{\Delta t}J_{N})\partial_{t}{\tilde{\bf u}}_{N}\cdot{\bf q}_{N}-\int_{0}^{T}\int_{\Gamma}\partial_{t}{\tilde{\bf v}}_{N}\cdot\boldsymbol{\psi}_{N}=\frac{1}{2}\int_{0}^{T}\int_{\mathcal{O}}\partial_{t}\tilde{J}_{N}{\bf u}_{N}\cdot{\bf q}_{N}\\ &-\frac{1}{2}\int_{0}^{T}\int_{\mathcal{O}}(\tau_{\Delta t}{J}_{N})((\tau_{\Delta t}{\bf u}_{N}-{\bf w}_{N})\cdot\nabla^{\tau_{\Delta t}{{\boldsymbol{\eta}}}_{N}}{\bf u}_{N}\cdot{\bf q}_{N}-(\tau_{\Delta t}{\bf u}_{N}-{\bf w}_{N})\cdot\nabla^{\tau_{\Delta t}{{\boldsymbol{\eta}}}_{N}}{\bf q}_{N}\cdot{\bf u}_{N})\\ &-2\nu\int_{0}^{T}\int_{\mathcal{O}}(\tau_{\Delta t}{J}_{N}){\bf D}^{\tau_{\Delta t}{{\boldsymbol{\eta}}}_{N}}({\bf u}_{N})\cdot{\bf D}^{\tau_{\Delta t}{{\boldsymbol{\eta}}}_{N}}({\bf q}_{N})-\int_{0}^{T}\int_{\Gamma}\Delta{{\boldsymbol{\eta}}}_{N}\cdot\Delta\boldsymbol{\psi}_{N}-\frac{1}{N}\int_{0}^{T}\int_{\Gamma}\nabla^{3}{{\boldsymbol{\eta}}}_{N}\cdot\nabla^{3}\boldsymbol{\psi}_{N}\\ &+\gamma\int_{0}^{T}\int_{\Gamma}\Lambda^{1+s}{\bf v}_{N}\cdot\Lambda^{1+s}\boldsymbol{\psi}_{N},\end{split}

for any (𝐪N,𝝍N)({\bf q}_{N},\boldsymbol{\psi}_{N}) where 𝐪N(t)𝒱F𝜼N(t){\bf q}_{N}(t)\in\mathscr{V}_{F}^{{\boldsymbol{\eta}}_{N}}(t) and 𝝍N(t)𝐇3(Γ)\boldsymbol{\psi}_{N}(t)\in{\bf H}^{3}(\Gamma) satisfy 𝐪N|Γ=𝝍N{\bf q}_{N}|_{\Gamma}=\boldsymbol{\psi}_{N}. Moreover, we have

τΔt𝜼N𝐮N=0,𝐮N|Γ=𝐯N\displaystyle\nabla^{\tau_{\Delta t}{\boldsymbol{\eta}}_{N}}\cdot{\bf u}_{N}=0,\qquad{\bf u}_{N}|_{\Gamma}={\bf v}_{N}
𝐮N(0,)=𝐮0,𝜼N(0,)=𝜼0,𝐯N(0,)=𝐯0.\displaystyle{\bf u}_{N}(0,\cdot)={\bf u}_{0},\,\,{\boldsymbol{\eta}}_{N}(0,\cdot)={\boldsymbol{\eta}}_{0},\,\,{\bf v}_{N}(0,\cdot)={\bf v}_{0}.

In the subsequent sections we will show that these sequences are bounded independently of NN in certain appropriate spaces which will allow us to extract subsequences converging in weak and strong topologies of appropriate subspaces of the energy space. Using these convergence results we aim to pass NN\to\infty in (LABEL:weakapprox).

4.3. Uniform estimates

In this section we will obtain the estimates, uniform in NN, for the approximate solutions defined in Section 4.2.

Theorem 4.2.

Assume, for some fixed α>0\alpha>0, that inf𝒪Jn>α\inf_{\mathcal{O}}J^{n}>\alpha for every 0nN0\leq n\leq N. Then there exists a constant C>0C>0, independent of NN and ε\varepsilon, such that

  1. (1)

    En+1C,En+12CE^{n+1}\leq C,E^{n+\frac{1}{2}}\leq C, for every n=0,1,..,Nn=0,1,..,N.

  2. (2)

    n=0N1DnC\sum_{n=0}^{N-1}D^{n}\leq C.

  3. (3)

    n=0N1(𝐯n+12𝐯n𝐋2(Γ)2+𝜼n+12𝜼n𝐇2(Γ)2+1N𝜼n+12𝜼n𝐇3(Γ)2)C\sum_{n=0}^{N-1}\left(\|{\bf v}^{n+\frac{1}{2}}-{\bf v}^{n}\|_{{\bf L}^{2}(\Gamma)}^{2}+\|{\boldsymbol{\eta}}^{n+\frac{1}{2}}-{\boldsymbol{\eta}}^{n}\|_{{\bf H}^{2}(\Gamma)}^{2}+\frac{1}{N}\|{\boldsymbol{\eta}}^{n+\frac{1}{2}}-{\boldsymbol{\eta}}^{n}\|_{{\bf H}^{3}(\Gamma)}^{2}\right)\leq C.

  4. (4)

    n=0N1(𝒪Jn(|𝐮n+1𝐮n|2)+𝐯n+1𝐯n+12𝐋2(Γ)2)C\sum_{n=0}^{N-1}\left(\int_{\mathcal{O}}J^{n}\left(|{\bf u}^{n+1}-{\bf u}^{n}|^{2}\right)+\|{\bf v}^{n+1}-{\bf v}^{n+\frac{1}{2}}\|^{2}_{{\bf L}^{2}(\Gamma)}\right)\leq C,

where the discrete energy ENnE_{N}^{n} and dissipation DNnD_{N}^{n} are defined in (56).

Proof.

For a fixed NN\in\mathbb{N}, we add the energy estimates for the two subproblems (59) and (60), sum over m1m\geq 1, summing 0nm10\leq n\leq m-1 and then take supremum over 1mN1\leq m\leq N. This gives us

(64) sup1mNENm+n=0N1DNn+n=0N1CNnE0,\begin{split}\sup_{1\leq m\leq N}E_{N}^{m}+\sum_{n=0}^{N-1}D_{N}^{n}+\sum_{n=0}^{N-1}C_{N}^{n}&\leq E^{0},\end{split}

where,

CNn:=𝐯n+12𝐯n𝐋2(Γ)2+𝜼n+12𝜼n𝐋2(Γ)2+𝒪Jn(|𝐮n+1𝐮n|2)+𝐯n+1𝐯n+12𝐋2(Γ)2.C^{n}_{N}:=\|{\bf v}^{n+\frac{1}{2}}-{\bf v}^{n}\|_{{\bf L}^{2}(\Gamma)}^{2}+\|{\boldsymbol{\eta}}^{n+\frac{1}{2}}-{\boldsymbol{\eta}}^{n}\|_{{\bf L}^{2}(\Gamma)}^{2}+\int_{\mathcal{O}}J^{n}\left(|{\bf u}^{n+1}-{\bf u}^{n}|^{2}\right)+\|{\bf v}^{n+1}-{\bf v}^{n+\frac{1}{2}}\|^{2}_{{\bf L}^{2}(\Gamma)}.

Next, we obtain uniform bounds for the approximate structure displacements and fluid velocity.

Theorem 4.3.

There exists T0>0T_{0}>0 such that for any 0<δ<s0<\delta<s,

  1. (1)

    The sequences {𝜼N},{𝜼~N}\{{{\boldsymbol{\eta}}}_{N}\},\{{\tilde{\boldsymbol{\eta}}}_{N}\} are bounded, independently of NN, in L(0,T0;𝐇2+δ(Γ))L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma)).

  2. (2)

    The sequence {𝐀𝜼N}\{{\bf A}_{{\boldsymbol{\eta}}_{N}}\} is bounded, independently of NN in L(0,T0;𝐂1,δ(𝒪¯)){L^{\infty}(0,T_{0};{\bf C}^{1,\delta}(\bar{\mathcal{O}}))} and for some α>0\alpha>0, the sequence of approximate Jacobians satisfies inf𝒪×(0,T)J𝜼N>α\inf_{\mathcal{O}\times(0,T)}J_{{\boldsymbol{\eta}}_{N}}>\alpha, for all NN.

  3. (3)

    The sequence {𝐮N}\{{\bf u}_{N}\} is bounded, independently of NN, in L2(0,T0;𝐇1(𝒪))L(0,T0;𝐋2(𝒪)).L^{2}(0,T_{0};{\bf H}^{1}(\mathcal{O}))\cap L^{\infty}(0,T_{0};{\bf L}^{2}(\mathcal{O})).

Proof.

We begin by observing that since 𝜼~N\tilde{\boldsymbol{\eta}}_{N} and 𝜼N{\boldsymbol{\eta}}_{N} belong to L(0,T;𝐇3(Γ))L^{\infty}(0,T;{\bf H}^{3}(\Gamma)), until some time, depending on NN, the condition inf𝒪×(0,T)JN>α\inf_{\mathcal{O}\times(0,T)}J_{N}>\alpha must be satisfied. We define TNmax{T}_{N}^{max} to be maximal interval on which this lower bound for the approximate Jacobian exists and hence the approximate solutions (𝐮N,𝜼~N)({\bf u}_{N},\tilde{\boldsymbol{\eta}}_{N}) are defined, i.e. TNmax{T}_{N}^{max} is a maximal time-interval such that assumptions of Theorem 4.1 hold.

First, we prove Statement (1). The proof of Statement (1) will rely heavily on the results of Theorem 1. We first find uniform bounds for the structure displacement until time TNmax:=nNmax(TN)>0T^{max}_{N}:=n^{max}_{N}(\frac{T}{N})>0, for some nNmaxn^{max}_{N}\in\mathbb{N} dependent on NN. For this purpose we reproduce the construction of the test functions and other important details from Theorem 1 for the semi-discrete case. We take

(65) 𝐪N=J𝜼N1𝐀𝜼N𝝋N,𝝍N=(Δκ𝜼NcN𝝃N),\displaystyle{\bf q}_{N}=-J_{{\boldsymbol{\eta}}_{N}}^{-1}\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}}\,\boldsymbol{\varphi}_{N},\qquad\boldsymbol{\psi}_{N}=-(\Delta^{\kappa}{\boldsymbol{\eta}}_{N}-c_{N}\boldsymbol{\xi}_{N}),

where 𝝋N\boldsymbol{\varphi}_{N} is the solution of (25) with boundary data 𝝋|Γ=J𝜼N(𝐀𝜼N)1|Γ(Δκ𝜼NcN𝝃N)\boldsymbol{\varphi}|_{\Gamma}=J_{{\boldsymbol{\eta}}_{N}}(\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}})^{-1}|_{\Gamma}\left(\Delta^{\kappa}{\boldsymbol{\eta}}_{N}-c_{N}\boldsymbol{\xi}_{N}\right), where

(66) cN=Γ𝜼N×Δκ𝜼N,\displaystyle{c}_{N}={\int_{\Gamma}\nabla{\boldsymbol{\eta}}_{N}\times\Delta^{\kappa}{\boldsymbol{\eta}}_{N}},

and 𝝃N\boldsymbol{\xi}_{N} is a smooth function satisfying Γ𝜼N×𝝃N=1{\int_{\Gamma}\nabla{\boldsymbol{\eta}}_{N}\times\boldsymbol{\xi}_{N}}=1 for every t[0,T]t\in[0,T].

We use this test pair (𝐪N,𝝍N)({\bf q}_{N},\boldsymbol{\psi}_{N}), defined in (65), in the weak formulation (LABEL:weakapprox) on the time interval (0,TNmax)(0,T^{max}_{N}) and follow, with some modifications, the steps presented in the proof of Theorem 1.

Observe that, 𝐯N{\bf v}_{N} is not equal to the structure velocity t𝜼~N\partial_{t}\tilde{\boldsymbol{\eta}}_{N}. Due to this mismatch caused by time-discretization and splitting, the term 0TNmaxΓΛ1+s𝐯NΛ1+s𝝍N\int_{0}^{T^{max}_{N}}\int_{\Gamma}\Lambda^{{1+s}}{\bf v}_{N}\cdot\Lambda^{1+s}\boldsymbol{\psi}_{N}, appearing in (LABEL:weakapprox) requires explanation. To that end, we write

(67) γ0TNmaxΓΛ1+s𝐯NΛ1+s𝝍N=γ0TNmaxΓΛ1+st𝜼~NΛ1+s+2κ𝜼~N+γ0TNmaxΓΛ1+s(𝐯N𝐯N)Λ1+s𝝍N+γ0TNmaxΓΛ1+s𝐯NΛ1+s+2κ(𝜼N𝜼~N).\begin{split}&\gamma\int_{0}^{T^{max}_{N}}\int_{\Gamma}\Lambda^{1+s}{\bf v}_{N}\cdot\Lambda^{1+s}\boldsymbol{\psi}_{N}=\gamma\int_{0}^{T^{max}_{N}}\int_{\Gamma}\Lambda^{{1+s}}\partial_{t}\tilde{\boldsymbol{\eta}}_{N}\cdot\Lambda^{{1+s}+2\kappa}\tilde{\boldsymbol{\eta}}_{N}\\ &+\gamma\int_{0}^{T^{max}_{N}}\int_{\Gamma}\Lambda^{{1+s}}({\bf v}_{N}-{\bf v}^{*}_{N})\cdot\Lambda^{{1+s}}\boldsymbol{\psi}_{N}+\gamma\int_{0}^{T^{max}_{N}}\int_{\Gamma}\Lambda^{{1+s}}{\bf v}^{*}_{N}\cdot\Lambda^{1+s+2\kappa}({\boldsymbol{\eta}}_{N}-\tilde{\boldsymbol{\eta}}_{N}).\end{split}

Observe that the first term on the right-hand side produces the desired L(0,TNmax;𝐇2+δ(Γ))L^{\infty}(0,T^{max}_{N};{\bf H}^{2+\delta}(\Gamma))-norm of 𝜼~N\tilde{\boldsymbol{\eta}}_{N} in Statement (1). We will now show that the remaining two terms are bounded. Thanks to integration-by-parts and the bounds on numerical dissipation obtained in Theorem 4.2, we obtain

|γ0TNmaxΓΛ1+s(𝐯N𝐯N)Λ1+s𝝍N|\displaystyle|\gamma\int_{0}^{T^{max}_{N}}\int_{\Gamma}\Lambda^{{1+s}}({\bf v}_{N}-{\bf v}^{*}_{N})\cdot\Lambda^{{1+s}}\boldsymbol{\psi}_{N}| =|γ0TNmaxΓ(𝐯N𝐯N)Λ2+2s𝝍N|\displaystyle=|\gamma\int_{0}^{T^{max}_{N}}\int_{\Gamma}({\bf v}_{N}-{\bf v}^{*}_{N})\cdot\Lambda^{2+2s}\boldsymbol{\psi}_{N}|
𝐯N𝐯NL2(0,T;𝐋2(Γ))𝜼NL2(0,T;𝐇3(Γ))\displaystyle\leq\|{\bf v}_{N}-{\bf v}^{*}_{N}\|_{L^{2}(0,T;{\bf L}^{2}(\Gamma))}\|{\boldsymbol{\eta}}_{N}\|_{L^{2}(0,T;{\bf H}^{3}(\Gamma))}
((Δt)n=0nNmax1𝐯n+12𝐯n𝐋2(Γ)2)12𝜼NL2(0,T;𝐇3(Γ)).\displaystyle\leq\left((\Delta t)\sum_{n=0}^{n^{max}_{N}-1}\|{\bf v}^{n+\frac{1}{2}}-{\bf v}^{n}\|_{{\bf L}^{2}(\Gamma)}^{2}\right)^{\frac{1}{2}}\|{\boldsymbol{\eta}}_{N}\|_{L^{2}(0,T;{\bf H}^{3}(\Gamma)).}

Similarly, to bound the second term on the right hand side of (LABEL:mismatch), we first note, using Theorem 4.2 (3), that

0TNmax𝜼N𝜼~N𝐇3(Γ)2𝑑t\displaystyle\int_{0}^{T^{max}_{N}}\|{\boldsymbol{\eta}}_{N}-\tilde{\boldsymbol{\eta}}_{N}\|^{2}_{{\bf H}^{3}(\Gamma)}dt =n=0nNmax1tntn+11Δt(ttn)𝜼n+1+(tn+1tΔt)𝜼n𝐇3(Γ)2𝑑t\displaystyle=\sum_{n=0}^{n^{max}_{N}-1}\int_{t^{n}}^{t^{n+1}}\frac{1}{\Delta t}\|(t-t^{n}){\boldsymbol{\eta}}^{n+1}+(t^{n+1}-t-\Delta t){\boldsymbol{\eta}}^{n}\|^{2}_{{\bf H}^{3}(\Gamma)}dt
=n=0nNmax1𝜼n+1𝜼n𝐇3(Γ)2tntn+1(ttnΔt)2𝑑t\displaystyle=\sum_{n=0}^{n^{max}_{N}-1}\|{\boldsymbol{\eta}}^{n+1}-{\boldsymbol{\eta}}^{n}\|_{{\bf H}^{3}(\Gamma)}^{2}\int_{t^{n}}^{t^{n+1}}\left(\frac{t-t^{n}}{\Delta t}\right)^{2}dt
CTNN=CT.\displaystyle\leq\frac{CT}{N}\cdot N=CT.

This gives us,

|γ0TNmaxΓΛ1+s𝐯NΛ1+s+2κ(𝜼N𝜼~N)|\displaystyle|\gamma\int_{0}^{T^{max}_{N}}\int_{\Gamma}\Lambda^{{1+s}}{\bf v}^{*}_{N}\cdot\Lambda^{1+s+2\kappa}({\boldsymbol{\eta}}_{N}-\tilde{\boldsymbol{\eta}}_{N})| 𝐯NL2(0,TNmax;𝐇1+s(Γ))𝜼N𝜼~NL2(0,TNmax;𝐇3(Γ))\displaystyle\leq\|{\bf v}^{*}_{N}\|_{L^{2}(0,T^{max}_{N};{\bf H}^{1+s}(\Gamma))}\|{\boldsymbol{\eta}}_{N}-\tilde{\boldsymbol{\eta}}_{N}\|_{L^{2}(0,T^{max}_{N};{\bf H}^{3}(\Gamma))}
C.\displaystyle\leq C.

The rest of the terms in the formulation (LABEL:weakapprox) with the test function (65) follow the bounds obtained in the proof of Theorem 1. Hence, the proof leading up to (43), gives us that

(68) 𝜼~NL(0,TNmax;𝐇2+δ(Γ))2+𝜼NL2(0,TNmax;𝐇3(sδ)(Γ))2K2+𝜼0𝐇2+δ(Γ)2,\displaystyle\|\tilde{\boldsymbol{\eta}}_{N}\|^{2}_{L^{\infty}(0,{T^{max}_{N}};{\bf H}^{2+\delta}(\Gamma))}+\|{\boldsymbol{\eta}}_{N}\|^{2}_{L^{2}(0,T^{max}_{N};{\bf H}^{3-(s-\delta)}(\Gamma))}\leq K_{2}+\|{\boldsymbol{\eta}}_{0}\|^{2}_{{\bf H}^{2+\delta}(\Gamma)},

where K2K_{2} depends on 𝐀𝜼NL(0,TNmax;𝐖1,(𝒪))\|{\bf A}_{{\boldsymbol{\eta}}_{N}}\|_{L^{\infty}(0,T^{max}_{N};{\bf W}^{1,\infty}(\mathcal{O}))} and inf𝒪×(0,TNmax)J𝜼N\inf_{\mathcal{O}\times(0,T^{max}_{N})}J_{{\boldsymbol{\eta}}_{N}}. Now we use the continuity argument to get rid of this dependence of K2K_{2} on the LHS norms by possibly reducing the length of the time interval.

Let C0>0C_{0}>0 and α>0\alpha>0 be such that 𝐀𝜼0𝐖1,(𝒪)<2C0\|{\bf A}_{{\boldsymbol{\eta}}_{0}}\|_{{\bf W}^{1,\infty}(\mathcal{O})}<2C_{0} and inf𝒪J0>α\inf_{\mathcal{O}}J_{0}>\alpha. We hypothesize that

(69) 𝐀𝜼~NL(0,T;𝐖1,(𝒪))2C0andinf𝒪×(0,T)J𝜼~N>α.\displaystyle\|{\bf A}_{\tilde{\boldsymbol{\eta}}_{N}}\|_{L^{\infty}(0,T;{\bf W}^{1,\infty}(\mathcal{O}))}\leq 2C_{0}\quad\text{and}\quad\inf_{\mathcal{O}\times(0,T)}J_{\tilde{\boldsymbol{\eta}}_{N}}>\alpha.

Now we choose T0T_{0} such that expression on the right hand side of (45) is smaller then C0C_{0}, and the expression on the RHS of (46) greater then 2α2\alpha. We define,

TN=min{T0,TNmax}.T_{N}=\min\{T_{0},{T}_{N}^{max}\}.

Then, thanks to the Sobolev embeddings used in (45), we conclude that

(70) 𝐀𝜼~NL(0,TN;𝐖1,(𝒪))C0andinf𝒪×(0,TN)J𝜼~N>2α.\displaystyle\|{\bf A}_{\tilde{\boldsymbol{\eta}}_{N}}\|_{L^{\infty}(0,{T_{N}};{\bf W}^{1,\infty}(\mathcal{O}))}\leq C_{0}\quad\text{and}\quad\inf_{\mathcal{O}\times(0,T_{N})}J_{\tilde{\boldsymbol{\eta}}_{N}}>2\alpha.

Now, owing to the fact that by construction 𝜼~N\tilde{\boldsymbol{\eta}}_{N} and thus 𝐀𝜼~N{\bf A}_{\tilde{\boldsymbol{\eta}}_{N}} are continuous in time, we can show that the conditions (a)-(d) of the Bootstrap principle [29, Propostion 1.21] are satisfied. Hence we have proven that, for any 0<δ<s0<\delta<s, the sequence {𝜼~N}\{{\tilde{\boldsymbol{\eta}}}_{N}\} is bounded, independently of NN, in L(0,TN;𝐇2+δ(Γ))L^{\infty}(0,T_{N};{\bf H}^{2+\delta}(\Gamma)).

To finish the proof of Statement (1) in Theorem 4.3, we show that there exists N0N_{0} such that

(71) TN=T0for allNN0.\displaystyle T_{N}=T_{0}\quad\text{for all}\quad N\geq N_{0}.

We prove (71) by contradiction. Assume that (71) is not true. Recall that, by (70) we have that inf𝒪J𝜼~N(t)>2α\inf_{\mathcal{O}}J_{\tilde{\boldsymbol{\eta}}_{N}}(t)>2\alpha for every t[0,TNmax]t\in[0,T_{N}^{max}] where TNmax=nNmaxΔtT_{N}^{max}=n_{N}^{max}\Delta t. However, then for small enough Δt\Delta t we can prolong the approximate solution, i.e. we can obtain JnNmax+1>αJ^{n_{N}^{max}+1}>\alpha which contradicts maximality of TNmaxT_{N}^{max}.

Hence we have now shown that, for any 0<δ<s0<\delta<s the sequence {𝜼~N}\{{\tilde{\boldsymbol{\eta}}}_{N}\} is bounded, independently of NN, in L(0,T0;𝐇2+δ(Γ))L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma)).

The rest of the statements follow thanks to (13) and Lemma 3.1. ∎

Thanks to Theorems 4.2 and 4.3, we can immediately conclude that there exist 𝜼C(0,T;𝐇2+δ(Γ))H1(0,T;𝐇1+s(Γ)){\boldsymbol{\eta}}\in C(0,T;{\bf H}^{2+\delta}(\Gamma))\cap H^{1}(0,T;{\bf H}^{1+s}(\Gamma)) for some 0<δ<s0<\delta<s, 𝐮L(0,T;𝐋2(𝒪))L2(0,T;𝐇1(𝒪)){\bf u}\in L^{\infty}(0,T;{\bf L}^{2}(\mathcal{O}))\cap L^{2}(0,T;{\bf H}^{1}(\mathcal{O})) and 𝐯L(0,T;𝐋2(Γ))L2(0,T;𝐇1+s(Γ)){\bf v}\in L^{\infty}(0,T;{\bf L}^{2}(\Gamma))\cap L^{2}(0,T;{\bf H}^{1+s}(\Gamma)) such that the following weak and weak convergence results hold, up to a subsequence, as NN\to\infty:

  1. (1)

    𝜼N𝜼{{\boldsymbol{\eta}}}_{N}\rightharpoonup{\boldsymbol{\eta}} weakly in L(0,T0;𝐇2+δ(Γ))L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma)) for any 0<δ<s0<\delta<s.

  2. (2)

    𝜼~N𝜼\tilde{\boldsymbol{\eta}}_{N}\rightharpoonup{\boldsymbol{\eta}} weakly in C(0,T0;𝐇2+δ(Γ))C(0,T_{0};{\bf H}^{2+\delta}(\Gamma)) for any 0<δ<s0<\delta<s.

  3. (3)

    𝐮N𝐮{\bf u}_{N}\rightharpoonup{\bf u} weakly in L2(0,T0;𝐇1(𝒪))L^{2}(0,T_{0};{\bf H}^{1}(\mathcal{O})) and weakly in L(0,T0;𝐋2(𝒪)).L^{\infty}(0,T_{0};{\bf L}^{2}(\mathcal{O})).

  4. (4)

    𝐯N𝐯{\bf v}_{N}\rightharpoonup{\bf v} weakly in L2(0,T0;𝐇1+s(Γ))L^{2}(0,T_{0};{\bf H}^{1+s}(\Gamma)) and weakly in L(0,T0;𝐋2(Γ))L^{\infty}(0,T_{0};{\bf L}^{2}(\Gamma)).

  5. (5)

    𝐯N𝐯{\bf v}^{*}_{N}\rightharpoonup{\bf v} weakly in L(0,T0;𝐋2(Γ))L^{\infty}(0,T_{0};{\bf L}^{2}(\Gamma)).

Furthermore,

t𝜼=𝐯a.e. in 𝒪×(0,T).\partial_{t}{\boldsymbol{\eta}}={\bf v}\quad a.e.\text{ in }\mathcal{O}\times(0,T).

We now seek to upgrade these results to strong convergence results to be able to pass to the limit in our nonlinear problem. In particular, due to the geometric nonlinearity introduced in the fluid equations via the ALE maps associated with the motion of the fluid domain and its boundary, we will require stronger convergence result for the structure displacements to be able to pass NN\to\infty in the approximate weak formulation. Thus, we start with the following result on strong convergence of structure displacements.

Proposition 4.4.

There exists a subsequence {𝛈N}\{{\boldsymbol{\eta}}_{N}\} of approximate structure displacements such that

(72) 𝜼N𝜼 strongly in L(0,T0;𝐇2+δ(Γ)) for any 0δ<s.\displaystyle{\boldsymbol{\eta}}_{N}\to{\boldsymbol{\eta}}\quad\text{ strongly in }L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma))\quad\text{ for any }0\leq\delta<s.
Proof.

We begin by recalling the Aubin-Lions compactness lemma which states that the following embedding is compact

L(0,T;H2+ε(Γ))W1,(0,T;L2(Γ))C([0,T];H2+δ(Γ)),for any 0δ<ε.L^{\infty}(0,T;H^{2+\varepsilon}(\Gamma))\cap W^{1,\infty}(0,T;L^{2}(\Gamma))\subset\subset C([0,T];H^{2+\delta}(\Gamma)),\quad\text{for any }0\leq\delta<\varepsilon.

Due to the uniform boundedness of 𝜼~N\tilde{\boldsymbol{\eta}}_{N} in L(0,T0;𝐇2+ε(Γ))L^{\infty}(0,T_{0};{\bf H}^{2+\varepsilon}(\Gamma)), for appropriately small ε>0\varepsilon>0 (see Theorem 4.2 (1)), and the uniform boundedness of 𝐯N=t𝜼~N{\bf v}^{*}_{N}=\partial_{t}\tilde{\boldsymbol{\eta}}_{N} in L(0,T0;𝐋2(Γ))L^{\infty}(0,T_{0};{\bf L}^{2}(\Gamma)), the compact embedding above implies that the sequence

(73) 𝜼~N𝜼 strongly in C([0,T0];𝐇2+δ(Γ)), for any 0δ<s.\displaystyle\tilde{\boldsymbol{\eta}}_{N}\to{\boldsymbol{\eta}}\text{ strongly in }C([0,T_{0}];{\bf H}^{2+\delta}(\Gamma)),\quad\text{ for any }0\leq\delta<s.

Then, by comparing the definitions (61) and (LABEL:approxlinear) we conclude that (72) holds for the sequences {𝜼N}\{{\boldsymbol{\eta}}_{N}\} and {τΔt𝜼N}\{\tau_{\Delta t}{\boldsymbol{\eta}}_{N}\}. ∎

The consequences of the strong convergence (72) in regards to approximate ALE maps are summarized in the following Proposition.

Proposition 4.5.

The ALE maps, defined by (10), satisfy the following strong convergence properties:

(74) 𝐀𝜼N𝐀𝜼inL(0,T0;𝐖2,31δ(𝒪))(𝒪)),\displaystyle{\bf A}_{{\boldsymbol{\eta}}_{N}}\to{\bf A}_{{\boldsymbol{\eta}}}\ {\rm in}\ {L^{\infty}(0,T_{0};{\bf W}^{2,\frac{3}{1-\delta}}(\mathcal{O}))}(\mathcal{O})),
(75) 𝐀𝜼N𝐀𝜼 in L(0,T0;𝐂(𝒪)),\displaystyle\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}}\rightarrow\nabla{\bf A}_{{\boldsymbol{\eta}}}\quad\text{ in }L^{\infty}(0,T_{0};{\bf C}(\mathcal{O})),
(76) (𝐀𝜼N)1(𝐀𝜼)1 in L(0,T0;𝐂(𝒪))\displaystyle(\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}})^{-1}\rightarrow(\nabla{\bf A}_{{\boldsymbol{\eta}}})^{-1}\quad\text{ in }L^{\infty}(0,T_{0};{\bf C}(\mathcal{O}))
(77) JNJ𝜼=det𝐀𝜼, in L(0,T0;C(𝒪)).\displaystyle J_{N}\rightarrow J_{{\boldsymbol{\eta}}}=\text{det}\nabla{\bf A}_{\boldsymbol{\eta}},\quad\text{ in }L^{\infty}(0,T_{0};C(\mathcal{O})).

Furthermore, let 𝐰𝛈=t𝐀𝛈{\bf w}_{\boldsymbol{\eta}}=\partial_{t}{\bf A}_{\boldsymbol{\eta}} be the solution of (10) with respect to the boundary data 𝐯{\bf v}. Then:

(78) 𝐰N𝐰𝜼 in L2(0,T0;𝐇32(𝒪)),\displaystyle{\bf w}_{N}\to{\bf w}_{\boldsymbol{\eta}}\quad\text{ in }L^{2}(0,T_{0};{\bf H}^{\frac{3}{2}}(\mathcal{O})),
(79) tJ~NtJ𝜼=J𝜼(𝜼𝐰𝜼) in L2(0,T;L3(𝒪)).\displaystyle\partial_{t}\tilde{J}_{N}\to\partial_{t}J_{\boldsymbol{\eta}}=J_{\boldsymbol{\eta}}(\nabla^{\boldsymbol{\eta}}\cdot{\bf w}_{\boldsymbol{\eta}})\quad\text{ in }L^{2}(0,T;L^{3}(\mathcal{O})).
Proof.

First observe that due to the linearity of (10), the bounds (14) and Proposition 4.4, we have the following estimate, which implies strong convergence (74):

(80) 𝐀𝜼N𝐀𝜼L(0,T0;𝐖2,31δ(𝒪))C𝜼N𝜼L(0,T0;𝐇2+δ(𝒪))0,\displaystyle\|{\bf A}_{{\boldsymbol{\eta}}_{N}}-{\bf A}_{{\boldsymbol{\eta}}}\|_{L^{\infty}(0,T_{0};{\bf W}^{2,\frac{3}{1-\delta}}(\mathcal{O}))}\leq C\|{{\boldsymbol{\eta}}_{N}}-{{\boldsymbol{\eta}}}\|_{L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\mathcal{O}))}\to 0,

where 𝐀𝜼{\bf A}_{\boldsymbol{\eta}} solves (10) with respect to the boundary data 𝜼{\boldsymbol{\eta}}.

Estimate (80), along with Proposition 4.4, imply the strong convergence results (75) - (77), as well as (78).

To prove (79) we recall that for two matrices AA and BB, the derivative of the determinant of AA acting on matrix BB, denoted by D(det)(A)BD(det)(A)B, is given by D(det)(A)B=det(A)tr(BA1)D(det)(A)B=det(A)tr(BA^{-1}). Hence, by applying the mean value theorem to det(𝐀𝜼n)det(𝐀𝜼n+j)det(\nabla{\bf A}_{{\boldsymbol{\eta}}^{n}})-det(\nabla{\bf A}_{{\boldsymbol{\eta}}^{n+j}}) we obtain, for some β[0,1]\beta\in[0,1], that

(81) |Jn+1JnΔt|=|det(𝐀n,β)n,β(𝐀𝜼n+1𝐀𝜼nΔt)|,\begin{split}\left|\frac{J^{n+1}-J^{n}}{\Delta t}\right|&=\left|\text{det}(\nabla{\bf A}^{n,\beta})\nabla^{n,\beta}\cdot\left(\frac{{\bf A}_{{\boldsymbol{\eta}}^{n+1}}-{\bf A}_{{\boldsymbol{\eta}}^{n}}}{\Delta t}\right)\right|,\end{split}

where n,β=𝜼n+β(𝜼n+1𝜼n)\nabla^{n,\beta}=\nabla^{{\boldsymbol{\eta}}^{n}}+\beta(\nabla^{{\boldsymbol{\eta}}^{n+1}}-\nabla^{{\boldsymbol{\eta}}^{n}}) and 𝐀n,β=𝐀𝜼n+β(𝐀𝜼n+1𝐀𝜼n)\nabla{\bf A}^{n,\beta}=\nabla{\bf A}_{{\boldsymbol{\eta}}^{n}}+\beta(\nabla{\bf A}_{{\boldsymbol{\eta}}^{n+1}}-\nabla{\bf A}_{{\boldsymbol{\eta}}^{n}}). The details of these calculations can be found in [25] (cf. (73)). Thus, (74)-(78) give us that

tJ~NtJ𝜼=J𝜼(𝜼𝐰𝜼), in L2(0,T;L3(𝒪)).\displaystyle\partial_{t}\tilde{J}_{N}\to\partial_{t}J_{\boldsymbol{\eta}}=J_{\boldsymbol{\eta}}(\nabla^{\boldsymbol{\eta}}\cdot{\bf w}_{\boldsymbol{\eta}}),\quad\text{ in }L^{2}(0,T;L^{3}(\mathcal{O})).

This completes the proof Proposition 4.5. ∎

Remark 7.

We note that since 𝛈L(0,T0;𝐇2+δ(Γ)){\boldsymbol{\eta}}\in L^{\infty}(0,T_{0};{\bf H}^{2+\delta}(\Gamma)), we have that 𝐀𝛈L(0,T0;𝐂1,δ(𝒪)){\bf A}_{\boldsymbol{\eta}}\in L^{\infty}(0,T_{0};{\bf C}^{1,\delta}(\mathcal{O})). Hence, on some time interval, still denoted by (0,T0)(0,T_{0}), 𝐀𝛈(t){\bf A}_{\boldsymbol{\eta}}(t) is a diffeomorphism from 𝒪\mathcal{O} to 𝒪𝛈(t)\mathcal{O}_{\boldsymbol{\eta}}(t) for every t(0,T0)t\in(0,T_{0}). Recall, from Remark 2.1, that this is necessary to ensure the equivalence of Definitions 1 and 2 on (0,T0)(0,T_{0}).

Next, we will prove strong convergence of the fluid and structure velocities. First, we obtain the following uniform bounds for the fluid and structure velocities in the Nikoski space N18,2(0,T0;𝐋2(𝒪)×𝐋2(Γ))N^{\frac{1}{8},2}(0,T_{0};{\bf L}^{2}(\mathcal{O})\times{\bf L}^{2}(\Gamma)). See (17) to recall the definition of Nikolski spaces.

Lemma 4.6.

The sequence of approximate solutions (𝐮N,𝐯N)({\bf u}_{N},{\bf v}_{N}) is bounded uniformly in the Nikolski space N18,2(0,T0;𝐋2(𝒪)×𝐋2(Γ))N^{\frac{1}{8},2}(0,T_{0};{\bf L}^{2}(\mathcal{O})\times{\bf L}^{2}(\Gamma)).

Proof.

The proof of this Lemma relies on the steps in the proof of Theorem 2. Namely, our aim is to prove that for any h>0h>0

(82) hT0τh𝐮N𝐮N𝐋2(𝒪)2+τh𝐯N𝐯N𝐋2(Γ)2=(Δt)n=jN𝐮n𝐮nj𝐋2(𝒪)2+𝐯n𝐯nj𝐋2(Γ)2Ch18,\int_{h}^{T_{0}}\|\tau_{h}{\bf u}_{N}-{\bf u}_{N}\|^{2}_{{\bf L}^{2}(\mathcal{O})}+\|\tau_{h}{\bf v}_{N}-{\bf v}_{N}\|^{2}_{{\bf L}^{2}(\Gamma)}=(\Delta t)\sum_{n=j}^{N}\|{\bf u}^{n}-{\bf u}^{n-j}\|^{2}_{{\bf L}^{2}(\mathcal{O})}+\|{\bf v}^{n}-{\bf v}^{n-j}\|^{2}_{{\bf L}^{2}(\Gamma)}\leq Ch^{\frac{1}{8}},

where the constant C>0C>0 is independent of hh and N=T0ΔtN=\frac{T_{0}}{\Delta t}. To prove this estimate we would like to use the monolithic approximate weak formulation (LABEL:weakapprox), and replace the test functions with the appropriate solutions. However, we have to be careful since 𝐮N,𝐯N{\bf u}_{N},{\bf v}_{N} are not the solutions of the monolithic approximate weak formulation (LABEL:weakapprox), as they satisfy the corresponding subproblems obtained using the Lie splitting strategy.

To get around this difficulty, we present here the construction of a suitable pair of test functions for the approximate sub-problems (57) and (LABEL:second), that are expressed in terms of modifications of 𝐮N,𝐯N{\bf u}_{N},{\bf v}_{N}, which can be used to derive estimate (82). Their construction will mimic the construction of their continuous-in-time counterparts (49) and (50), as in the proof of Theorem 2.

Let h=jΔt+lh=j\Delta t+l for some 0jN0\leq j\leq N and l<Δtl<\Delta t. For simplicity of our presentation we will take l=0l=0 and refer the interested reader to [26] (see (3.8)-(3.10)) for the treatment of the case l>0l>0.

To construct the appropriate test functions we fix Δt\Delta t, i.e., we fix NN, and consider nn and kk such that 0nkN0\leq n\neq k\leq N. The plan is to construct 𝐮k{\bf u}^{k} and 𝐯k{\bf v}^{k} (we are dropping the subscript NN here) in a way that they can be used as test functions for the equations for 𝐮n{\bf u}^{n} and 𝐯n{\bf v}^{n}, for some 0nkN0\leq n\neq k\leq N. Due to the fact that we are working on moving domains, this is not trivial since these two solutions are defined on different fluid domains.

We start by defining a discrete version of the function 𝐮M(s,t){\bf u}_{M}(s,t) in (49) as follows:

𝐮Mk,n\displaystyle{\bf u}^{k,n}_{M} :=(Jn)1𝐀𝜼n(Jk(𝐀𝜼k)1(𝐮k𝐰k))+𝐰Mk+(bkbMk,n)ξ0χ\displaystyle:=(J^{n})^{-1}\nabla{\bf A}_{{\boldsymbol{\eta}}^{n}}\left(J^{k}(\nabla{\bf A}_{{\boldsymbol{\eta}}^{k}})^{-1}({\bf u}^{k}-{\bf w}^{k})\right)+{\bf w}^{k}_{M}+\left({b^{k}-b^{k,n}_{M}}\right)\xi_{0}\chi
(Jn)1𝐀𝜼n(div((Jk(𝐀𝜼k)1𝐰k)Jn(𝐀𝜼n)1𝐰Mk)\displaystyle-(J^{n})^{-1}\nabla{\bf A}_{{\boldsymbol{\eta}}^{n}}\mathcal{B}\Big{(}\text{div}\left((J^{k}(\nabla{\bf A}_{{\boldsymbol{\eta}}^{k}})^{-1}{\bf w}^{k})-J^{n}(\nabla{\bf A}_{{\boldsymbol{\eta}}^{n}})^{-1}{\bf w}^{k}_{M}\right)
(bkbMk,n)div((Jn)1𝐀𝜼nξ0χ)),\displaystyle-\left({b^{k}-b^{k,n}_{M}}\right)\text{div}((J^{n})^{-1}\nabla{\bf A}_{{\boldsymbol{\eta}}^{n}}\xi_{0}\chi)\Big{)},

where 𝐰k{\bf w}^{k} is the harmonic extension of 𝐯k{\bf v}^{k} in 𝒪\mathcal{O}, such that 𝐰k=0{\bf w}^{k}=0 on 𝒪Γ\partial\mathcal{O}\setminus\Gamma. Similarly, for any 0kN0\leq k\leq N, we use the notation 𝐯Mk=PM𝐯k{\bf v}^{k}_{M}=P_{M}{\bf v}^{k} and denote by 𝐰Mk{\bf w}^{k}_{M} the harmonic extension of 𝐯Mk{\bf v}^{k}_{M} in 𝒪\mathcal{O}, such that 𝐰Mk=0{\bf w}^{k}_{M}=0 on 𝒪Γ\partial\mathcal{O}\setminus\Gamma. We recall that \mathcal{B} is the Bogovski operator on the fixed domain 𝒪\mathcal{O}. We also define the discreet correction terms, bk=Γ(𝜼k×𝐯k)b^{k}=\int_{\Gamma}(\nabla{\boldsymbol{\eta}}^{k}\times{\bf v}^{k}) and bMk,n=Γ(𝜼n×𝐯Mk)b^{k,n}_{M}=\int_{\Gamma}(\nabla{\boldsymbol{\eta}}^{n}\times{\bf v}_{M}^{k}) and choose a smooth function ξ0\xi_{0} so that it satisfies Γ(𝜼n×ξ0)=1\int_{\Gamma}(\nabla{\boldsymbol{\eta}}^{n}\times\xi_{0})=1 for any nn.

Similarly, the following function is the time discretized version of 𝐯M(s,t){\bf v}_{M}(s,t) in (50):

𝐯Mk,n:=𝐯Mk(bkbMk,n)ξ0.{\bf v}^{k,n}_{M}:={\bf v}^{k}_{M}-\left({b^{k}-b^{k,n}_{M}}\right)\xi_{0}.

Then, for any nNn\leq N we define the following pair of test functions written in terms of 𝐮Mk,n{\bf u}^{k,n}_{M} and 𝐯Mk,n{\bf v}_{M}^{k,n} (compare with (49) and (50)),

(83) (𝐪n,𝝍n)=((Δt)k=nj+1n𝐮Mk,n,(Δt)k=nj+1n𝐯Mk,n).\displaystyle({\bf q}_{n},\boldsymbol{\psi}_{n})=\left((\Delta t)\sum_{k=n-j+1}^{n}{\bf u}^{k,n}_{M},\,\,(\Delta t)\sum_{k=n-j+1}^{n}{\bf v}_{M}^{k,n}\right).

Now we use (𝐪n,𝝍n)({\bf q}_{n},\boldsymbol{\psi}_{n}) as test functions in the two subproblems (57)3 and (LABEL:second)2 respectively and then apply n=0N\sum_{n=0}^{N} which yields (cf. (55)),

n=0N(𝒪((Jn+1)𝐮n+1(Jn)𝐮n)(Δtk=nj+1n𝐮Mk,n)+Γ(𝐯n+1𝐯n)(Δtk=nj+1n𝐯Mk,n))\displaystyle-\sum_{n=0}^{N}\left(\int_{\mathcal{O}}\left((J^{n+1}){\bf u}^{n+1}-(J^{n}){\bf u}^{n}\right)\left(\Delta t\sum_{k=n-j+1}^{n}{\bf u}^{k,n}_{M}\right)+\int_{\Gamma}({\bf v}^{n+1}-{\bf v}^{n})\left(\Delta t\sum_{k=n-j+1}^{n}{\bf v}_{M}^{k,n}\right)\right)
=12n=0N𝒪(Jn+1Jn)𝐮n+1(Δtk=nj+1n𝐮Mk,n)+(Δt)n=0Nb𝜼n(𝐮n+1,𝐰n,(Δtk=nj+1n𝐮Mk,n))\displaystyle=\frac{-1}{2}\sum_{n=0}^{N}\int_{\mathcal{O}}\left(J^{n+1}-J^{n}\right){\bf u}^{n+1}\cdot\left(\Delta t\sum_{k=n-j+1}^{n}{\bf u}^{k,n}_{M}\right)+(\Delta t)\sum_{n=0}^{N}b^{{\boldsymbol{\eta}}_{n}}({\bf u}^{n+1},{\bf w}^{n},\left(\Delta t\sum_{k=n-j+1}^{n}{\bf u}^{k,n}_{M}\right))
+2ν(Δt)n=0N𝒪(Jn)𝐃ηn(𝐮n+1)𝐃ηn(Δtk=nj+1n𝐮Mk,n)\displaystyle+2\nu(\Delta t)\sum_{n=0}^{N}\int_{\mathcal{O}}(J^{n}){\bf D}^{\eta^{n}}({\bf u}^{n+1})\cdot{\bf D}^{\eta^{n}}\left(\Delta t\sum_{k=n-j+1}^{n}{\bf u}^{k,n}_{M}\right)
+(Δt)n=0NΓΔ𝜼n+12(Δtk=nj+1nΔ𝐯Mk,n)+1N(Δt)n=0NΓ3𝜼n+12(Δtk=nj+1n3𝐯Mk,n)\displaystyle+(\Delta t)\sum_{n=0}^{N}\int_{\Gamma}\Delta{\boldsymbol{\eta}}^{n+\frac{1}{2}}\cdot\left(\Delta t\sum_{k=n-j+1}^{n}\Delta{\bf v}_{M}^{k,n}\right)+\frac{1}{N}(\Delta t)\sum_{n=0}^{N}\int_{\Gamma}\nabla^{3}{\boldsymbol{\eta}}^{n+\frac{1}{2}}\cdot\left(\Delta t\sum_{k=n-j+1}^{n}\nabla^{3}{\bf v}_{M}^{k,n}\right)
+γΓΛ1+s𝐯n+12Λ1+s(Δtk=nj+1n𝐯Mk,n)\displaystyle+\gamma\int_{\Gamma}\Lambda^{1+s}{\bf v}^{n+\frac{1}{2}}\cdot\Lambda^{1+s}\left(\Delta t\sum_{k=n-j+1}^{n}{\bf v}^{k,n}_{M}\right)
=I1++I5.\displaystyle=I_{1}+...+I_{5}.

After using summation by parts formula, the two terms on the left hand side of the equation above produce the desired terms hT0τh𝐮N𝐮N𝐋2(𝒪)2+hT0τh𝐯N𝐯N𝐋2(Γ)2\int_{h}^{T_{0}}\|\tau_{h}{\bf u}_{N}-{\bf u}_{N}\|^{2}_{{\bf L}^{2}(\mathcal{O})}+\int_{h}^{T_{0}}\|\tau_{h}{\bf v}_{N}-{\bf v}_{N}\|^{2}_{{\bf L}^{2}(\Gamma)}. The rest of the terms including the six terms I1,,5I_{1,...,5} on the right-hand side of the equation above are treated identically as in the proof of Theorem 2 (see the bounds obtained for the terms I1,,6I_{1,...,6} following (55)). This completes the proof of Lemma 4.6. ∎

To utilize this result and obtain strong convergence, up to a subsequence, of the approximate solutions, we intend to use the following variant of the Aubin-Lions theorem (see [32] and [28]).

Lemma 4.7.

Assume that 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} are Banach spaces such that 𝒴0\mathcal{Y}_{0} and 𝒴\mathcal{Y} are reflexive with compact embedding of 𝒴0\mathcal{Y}_{0} in 𝒴\mathcal{Y}. Then for any m>0m>0, the embedding

L2(0,T;𝒴0)Nm,2(0,T;𝒴)L2(0,T;𝒴)L^{2}(0,T;\mathcal{Y}_{0})\cap N^{m,2}(0,T;\mathcal{Y})\hookrightarrow L^{2}(0,T;\mathcal{Y})

is compact.

Hence, combining Lemmas 4.6 and 4.7 with 𝒴0=𝐇1(𝒪)×𝐇1+s(Γ)\mathcal{Y}_{0}={\bf H}^{1}(\mathcal{O})\times{\bf H}^{1+s}(\Gamma) and 𝒴=𝐋2(𝒪)×𝐋2(Γ)\mathcal{Y}={\bf L}^{2}(\mathcal{O})\times{\bf L}^{2}(\Gamma), we see that the sequence {(𝐮N,𝐯N)}\{({\bf u}_{N},{\bf v}_{N})\} is relatively compact in L2(0,T0;𝐋2(𝒪)×𝐋2(Γ))L^{2}(0,T_{0};{\bf L}^{2}(\mathcal{O})\times{\bf L}^{2}(\Gamma)). Therefore, we obtained the following strong convergence result for fluid and structure velocities:

Proposition 4.8.

The sequence

{(𝐮N,𝐯N)}(𝐮,𝐯) strongly in L2(0,T0;𝐋2(𝒪)×𝐋2(Γ)).\displaystyle\{({\bf u}_{N},{\bf v}_{N})\}\to({\bf u},{\bf v})\quad\text{ strongly in }L^{2}(0,T_{0};{\bf L}^{2}(\mathcal{O})\times{\bf L}^{2}(\Gamma)).

This completes the convergence results for the approximate solutions that are necessary to pass to the limit as NN\to\infty in the monolithic weak formulation of approximate solutions and show that the limits satisfy the weak formulation of the continuous problem. However, as with all FSI problems defined on moving domains, for which the continuous solution is approximated by a sequence of time-discretized approximate solutions, before we can pass to the limit we need to take care of the corresponding test functions. Namely, the test functions in the approximate weak formulations given in terms of the fixed domain 𝒪\mathcal{O}, depend on NN because they satisfy the transformed divergence-free condition, where the gradient operator depends on NN via the approximate interface displacement 𝜼N{\boldsymbol{\eta}}_{N}. Therefore, before we can pass to the limit as NN\to\infty (or equivalently Δt0\Delta t\to 0) in (LABEL:weakapprox) we need to construct appropriate test functions that satisfy certain strong convergence properties, and are dense in the space of approximate and continuous test functions. This is the subject of the next section.

4.4. Construction of test functions

In this section, we will construct a pair of test functions for the approximate weak formulation (LABEL:weakapprox) and the corresponding limiting equations that have certain desired properties to pass to the limit as NN\to\infty. We begin by considering a test pair (𝐪,𝝍)({\bf q},\boldsymbol{\psi}) for some 𝐪L(0,T;𝐇1(𝒪))H1(0,T;𝐋2(𝒪)){\bf q}\in L^{\infty}(0,T;{\bf H}^{1}(\mathcal{O}))\cap H^{1}(0,T;{\bf L}^{2}(\mathcal{O})) such that 𝜼𝐪=0\nabla^{{\boldsymbol{\eta}}}\cdot{\bf q}=0, and 𝝍L(0,T;𝐇2(Γ))H1(0,T;𝐇1+s(Γ))\boldsymbol{\psi}\in L^{\infty}(0,T;{\bf H}^{2}(\Gamma))\cap H^{1}(0,T;{\bf H}^{1+s}(\Gamma)) that satisfy the kinematic coupling condition i.e. 𝐪|Γ=𝝍{\bf q}|_{\Gamma}=\boldsymbol{\psi}. Now, we will build a pair of functions (𝐪N,𝝍N)({\bf q}_{N},\boldsymbol{\psi}_{N}) that approximates (𝐪,𝝍)({\bf q},\boldsymbol{\psi}) in an appropriate sense and is also a valid test function for the approximate system (LABEL:weakapprox).

We define the approximate fluid test function 𝐪N{\bf q}_{N}, with the aid of the Piola transformation as done previously in the proof of Theorem 2 in Section 3.2:

(84) 𝐪N=J𝜼N1𝐀𝜼NJ𝜼𝐀𝜼1(𝐪𝝍χ)+𝝍χλ𝜼Nλ𝜼λ0N(ξ0χ)\displaystyle{\bf q}_{N}={J^{-1}_{{\boldsymbol{\eta}}_{N}}}\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}}{J_{{\boldsymbol{\eta}}}}\nabla{\bf A}^{-1}_{{\boldsymbol{\eta}}}\left({\bf q}-\boldsymbol{\psi}\chi\right)+\boldsymbol{\psi}\chi-\frac{\lambda^{{{\boldsymbol{\eta}}}_{N}}-\lambda^{{\boldsymbol{\eta}}}}{\lambda^{N}_{0}}(\xi_{0}\chi)
+J𝜼N1(div((J𝜼(𝐀𝜼)1J𝜼N(𝐀𝜼N)1)𝝍χλ𝜼Nλ𝜼λ0NJ𝜼N(𝐀𝜼N)1ξ0χ)),\displaystyle+{J^{-1}_{{\boldsymbol{\eta}}_{N}}}\mathcal{B}\left(\text{div}\left((J_{{\boldsymbol{\eta}}}(\nabla{\bf A}_{{\boldsymbol{\eta}}})^{-1}-J_{{\boldsymbol{\eta}}_{N}}(\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}})^{-1})\boldsymbol{\psi}\chi-\frac{\lambda^{{{\boldsymbol{\eta}}}_{N}}-\lambda^{{\boldsymbol{\eta}}}}{\lambda^{N}_{0}}J_{{\boldsymbol{\eta}}_{N}}(\nabla{\bf A}_{{\boldsymbol{\eta}}_{N}})^{-1}\xi_{0}\chi\right)\right),

and for the structure test function we let,

(85) 𝝍N=𝝍λ𝜼Nλ𝜼λ0Nξ0.\displaystyle\boldsymbol{\psi}_{N}=\boldsymbol{\psi}-\frac{\lambda^{{{\boldsymbol{\eta}}}_{N}}-\lambda^{{{\boldsymbol{\eta}}}}}{\lambda^{N}_{0}}\xi_{0}.

For the correction terms, we pick an appropriate ξ0𝐂0(Γ)\xi_{0}\in{\bf C}_{0}^{\infty}(\Gamma) such that λ0N\lambda_{0}^{N} defined below is not 0 for any NN\in\mathbb{N},

λ0N(t)=Γ(𝐢𝐝+𝜼N(t))×ξ0.\lambda_{0}^{N}(t)=-\int_{\Gamma}({\bf id}+{{\boldsymbol{\eta}}}_{N}(t))\times\nabla\xi_{0}.\quad\text{}

We also define the following corrector functions that only depend on time,

λ𝜼N(t)=Γ(𝐢𝐝+𝜼N(t))×𝝍(t),λ𝜼(t)=Γ(𝐢𝐝+𝜼(t))×𝝍(t).\lambda^{{{\boldsymbol{\eta}}}_{N}}(t)=-\int_{\Gamma}({\bf id}+{{\boldsymbol{\eta}}}_{N}(t))\times\nabla\boldsymbol{\psi}(t),\quad\lambda^{{{\boldsymbol{\eta}}}}(t)=-\int_{\Gamma}({\bf id}+{{\boldsymbol{\eta}}}(t))\times\nabla\boldsymbol{\psi}(t).

As earlier, χ(r)\chi(r) is a smooth function on 𝒪\mathcal{O} such that χ(1)=1\chi(1)=1 and χ(0)=0\chi(0)=0. Observe that the properties of the Piola transformation (see e.g. Theorem 1.7 in [6]), imply that

𝜼N𝐪N=J𝜼J𝜼N1𝜼𝐪=0,and𝐪N|Γ𝐧N=𝝍N𝐧N.\nabla^{{\boldsymbol{\eta}}_{N}}\cdot{\bf q}_{N}=J_{{\boldsymbol{\eta}}}J^{-1}_{{\boldsymbol{\eta}}_{N}}\nabla^{{\boldsymbol{\eta}}}\cdot{\bf q}=0,\qquad\text{and}\qquad{\bf q}_{N}|_{\Gamma}\cdot{\bf n}_{N}=\boldsymbol{\psi}_{N}\cdot{\bf n}_{N}.

Additionally, thanks to (14) and (74) we obtain that

𝐪N𝐪L(0,T;𝐇1(𝒪))\displaystyle\|{\bf q}_{N}-{\bf q}\|_{L^{\infty}(0,T;{\bf H}^{1}(\mathcal{O}))}
𝐀𝜼N𝐀𝜼L(0,T;𝐖2,31δ(𝒪))(𝐪L(0,T;𝐇1(𝒪))+𝝍L(0,T;𝐇2(Γ)))\displaystyle\leq\|{\bf A}_{{\boldsymbol{\eta}}_{N}}-{\bf A}_{{\boldsymbol{\eta}}}\|_{L^{\infty}(0,T;{\bf W}^{2,\frac{3}{1-\delta}}(\mathcal{O}))}\left(\|{\bf q}\|_{L^{\infty}(0,T;{\bf H}^{1}(\mathcal{O}))}+\|\boldsymbol{\psi}\|_{L^{\infty}(0,T;{\bf H}^{2}(\Gamma))}\right)
(86) +λ𝜼Nλ𝜼L(0,T)0.\displaystyle\qquad+\|\lambda^{{{\boldsymbol{\eta}}}_{N}}-\lambda^{{{\boldsymbol{\eta}}}}\|_{L^{\infty}(0,T)}\to 0.

Similarly, since λ𝜼\lambda^{\boldsymbol{\eta}} and λ𝜼N\lambda^{{\boldsymbol{\eta}}_{N}} are constant in space, we readily obtain that

(87) 𝝍N𝝍 in L(0,T;𝐇2(Γ¯)).\begin{split}\boldsymbol{\psi}_{N}\rightarrow\boldsymbol{\psi}\quad\text{ in }L^{\infty}(0,T;{\bf H}^{2}(\bar{\Gamma})).\end{split}

Thus, we have shown the following result:

Proposition 4.9.

The approximate fluid velocity test functions 𝐪N{\bf q}_{N} constructed in (84), and the approximate structure velocity test functions 𝛙N\boldsymbol{\psi}_{N} constructed in (85), satisfy the following properties:

  • 𝜼N𝐪N=0\nabla^{{\boldsymbol{\eta}}_{N}}\cdot{\bf q}_{N}=0

  • 𝐪N|Γ𝐧N=𝝍N𝐧N{\bf q}_{N}|_{\Gamma}\cdot{\bf n}_{N}=\boldsymbol{\psi}_{N}\cdot{\bf n}_{N}.

Furthermore, the following strong convergence results hold:

𝐪N𝐪inL(0,T;𝐇1(𝒪))and𝝍N𝝍 in L(0,T;𝐇2(Γ¯)),{\bf q}_{N}\rightarrow{\bf q}\ \ {\rm in}\ \ {L^{\infty}(0,T;{\bf H}^{1}(\mathcal{O}))}\ \ {\rm and}\ \ \boldsymbol{\psi}_{N}\rightarrow\boldsymbol{\psi}\quad\text{ in }L^{\infty}(0,T;{\bf H}^{2}(\bar{\Gamma})),

where (𝐪,𝛙)({\bf q},\boldsymbol{\psi}) are the test functions associated with the continuous problem.

4.5. Passing to the limit

We are now in a position to pass to the limit in the semi-discrete formulation (LABEL:weakapprox), as NN\to\infty. We use the test functions constructed above in Proposition 4.9 as the test functions in the semi-discrete formulation (LABEL:weakapprox) (these test functions are dense in the space of all test functions for the approximate problems). Then, use the weak and strong convergence results discussed above for approximate solutions, and pass to the limit as NN\to\infty in (LABEL:weakapprox) to show that the limits of approximate subsequences satisfy the weak formulation of the continuous problem stated in Definition 2. Due to Proposition 2.1, Definition 2 and Definition 1 are equivalent, which completes the proof of Theorem 3.

Acknowledgment

SC acknowledges support from the National Science Foundation grants DMS-240892, DMS-2247000, DMS-2011319. BM acknowledges partial support from the Croatian Science Foundation, project number IP-2022-10-2962 and from the Croatia-USA bilateral grant “The mathematical framework for the diffuse interface method applied to coupled problems in fluid dynamics”. KT acknowledges support from the National Science Foundation grant DMS-2407197.

References

  • [1] A. Behzadan and M. Holst. Multiplication in Sobolev spaces, revisited. Ark. Mat., 59(2):275–306, 2021.
  • [2] T. Bodnár, G. Galdi, and Š. Nečasová. Fluid-structure interaction and biomedical applications. Springer, 2014.
  • [3] S. Čanić. Moving boundary problems. Bull. Amer. Math. Soc. (N.S.), 58(1):79–106, 2021.
  • [4] A. Chambolle, B. Desjardins, M. J. Esteban, and C. Grandmont. Existence of weak solutions for the unsteady interaction of a viscous fluid with an elastic plate. J. Math. Fluid Mech., 7(3):368–404, 2005.
  • [5] S. Chen and R. Triggiani. Gevrey class semigroups arising from elastic systems with gentle dissipation: the case 0<α<120<\alpha<\frac{1}{2}. Proc. Amer. Math. Soc., 110(2):401–415, 1990.
  • [6] P. Ciarlet. Mathematical elasticity, Vol. I. volume 20 of Studies in Mathematics and its Applications. North-Holland Publishing Co., Amsterdam, 1988.
  • [7] A. Fursikov, M. Gunzburger, and L. Hou. Trace theorems for three-dimensional, time-dependent solenoidal vector fields and their applications. Trans. Amer. Math. Soc., 354(3):1079–1116, 2002.
  • [8] G. P. Galdi. An introduction to the mathematical theory of the Navier-Stokes equations. Springer Monographs in Mathematics. Springer, New York, second edition, 2011. Steady-state problems.
  • [9] M. Geiß ert, H. Heck, and M. Hieber. On the equation divu=g{\rm div}\,u=g and Bogovskiĭ’s operator in Sobolev spaces of negative order. In Partial differential equations and functional analysis, volume 168 of Oper. Theory Adv. Appl., pages 113–121. Birkhäuser, Basel, 2006.
  • [10] D. Gilbarg and N. Trudinger. Elliptic partial differential equations of second order, volume 224 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, second edition, 1983.
  • [11] C. Grandmont. Existence of weak solutions for the unsteady interaction of a viscous fluid with an elastic plate. SIAM J. Math. Anal., 40(2):716–737, 2008.
  • [12] C. Grandmont and M. Hillairet. Existence of global strong solutions to a beam-fluid interaction system. Arch. Ration. Mech. Anal., 220(3):1283–1333, 2016.
  • [13] C. Grandmont, M. Hillairet, and J. Lequeurre. Existence of local strong solutions to fluid-beam and fluid-rod interaction systems. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 36(4):1105–1149, 2019.
  • [14] C. Grandmont and L. Sabbagh. Existence and uniqueness of strong solutions to a bi-dimensional fluid-structure interaction system. 2024.
  • [15] P. Grisvard. Elliptic problems in nonsmooth domains, volume 69 of Classics in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2011. Reprint of the 1985 original [MR0775683], With a foreword by Susanne C. Brenner.
  • [16] B. Kaltenbacher, I. Kukavica, I. Lasiecka, R. Triggiani, A. Tuffaha, and J. Webster. Mathematical theory of evolutionary fluid-flow structure interactions. Springer, 2018.
  • [17] M. Kampschulte, S. Schwarzacher, and G. Sperone. Unrestricted deformations of thin elastic structures interacting with fluids. J. Math. Pures Appl. (9), 173:96–148, 2023.
  • [18] I. Kukavica and A. Tuffaha. Solutions to a fluid-structure interaction free boundary problem. Discrete Contin. Dyn. Syst., 32(4):1355–1389, 2012.
  • [19] V. Mácha, B. Muha, Š. Nečasová, A. Roy, and S. Trifunović. Existence of a weak solution to a nonlinear fluid-structure interaction problem with heat exchange. Comm. Partial Differential Equations, 47(8):1591–1635, 2022.
  • [20] D. Maity and T. Takahashi. LpL^{p} theory for the interaction between the incompressible Navier-Stokes system and a damped plate. J. Math. Fluid Mech., 23(4):Paper No. 103, 23, 2021.
  • [21] B. Muha and S. Schwarzacher. Existence and regularity of weak solutions for a fluid interacting with a non-linear shell in three dimensions. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 39(6):1369–1412, 2022.
  • [22] B. Muha and S. Čanić. Existence of a weak solution to a nonlinear fluid-structure interaction problem modeling the flow of an incompressible, viscous fluid in a cylinder with deformable walls. Arch. Ration. Mech. Anal., 207(3):919–968, 2013.
  • [23] B. Muha and S. Čanić. Existence of a solution to a fluid-multi-layered-structure interaction problem. J. Differential Equations, 256(2):658–706, 2014.
  • [24] B. Muha and S. Čanić. Existence of a weak solution to a fluid-elastic structure interaction problem with the Navier slip boundary condition. J. Differential Equations, 260(12):8550–8589, 2016.
  • [25] B. Muha and S. Čanić. Existence of a weak solution to a fluid-elastic structure interaction problem with the Navier slip boundary condition. J. Differential Equations, 260(12):8550–8589, 2016.
  • [26] B. Muha and S. Čanić. A generalization of the Aubin-Lions-Simon compactness lemma for problems on moving domains. J. Differential Equations, 266(12):8370–8418, 2019.
  • [27] S. Schwarzacher and M. Sroczinski. Weak-strong uniqueness for an elastic plate interacting with the Navier-Stokes equation. SIAM J. Math. Anal., 54(4):4104–4138, 2022.
  • [28] J. Simon. Compact sets in the space Lp(0,T;B)L^{p}(0,T;B). Ann. Mat. Pura Appl. (4), 146:65–96, 1987.
  • [29] T. Tao. Nonlinear dispersive equations, volume 106 of CBMS Regional Conference Series in Mathematics. Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2006. Local and global analysis.
  • [30] K. Tawri. A stochastic fluid-structure interaction problem with navier slip boundary condition. SIAM Journal on Mathematical Analysis, to appear, 2024. arXiv:2402.13303.
  • [31] L. Tebou. Regularity and stability for a plate model involving fractional rotational forces and damping. Zeitschrift für angewandte Mathematik und Physik, 72(4):158, 2021.
  • [32] R. Temam. Navier-Stokes equations and nonlinear functional analysis, volume 66 of CBMS-NSF Regional Conference Series in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, second edition, 1995.
  • [33] S. Trifunović and Y.-G. Wang. Existence of a weak solution to the fluid-structure interaction problem in 3D. J. Differential Equations, 268(4):1495–1531, 2020.
  • [34] S. Čanić, M. Galić, and B. Muha. Analysis of a 3D nonlinear, moving boundary problem describing fluid-mesh-shell interaction. Trans. Amer. Math. Soc., 373(9):6621–6681, 2020.
  • [35] S. Čanić, J. Kuan, B. Muha, and K. Tawri. Deterministic and Stochastic Fluid-Structure Interaction. Advances in Mathematical Fluid Mechanics. Birkhäuser/Springer, 2024, conditionally accepted.
  • [36] I. Velčić. Nonlinear weakly curved rod by Γ\Gamma-convergence. J. Elasticity, 108(2):125–150, 2012.