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

Fluid-poroviscoelastic structure interaction problem with nonlinear geometric coupling

Jeffrey Kuan, Sunčica Čanić, Boris Muha
Abstract

We investigate weak solutions to a fluid-structure interaction (FSI) problem between the flow of an incompressible, viscous fluid modeled by the Navier-Stokes equations, and a poroviscoelastic medium modeled by the Biot equations. These systems are coupled nonlinearly across an interface with mass and elastic energy, modeled by a reticular plate equation, which is transparent to fluid flow. We provide a constructive proof of the existence of a weak solution to a regularized problem. Next, a weak-classical consistency result is obtained, showing that the weak solution to the regularized problem converges, as the regularization parameter approaches zero, to a classical solution to the original problem, when such a classical solution exists. While the assumptions in the first step only require the Biot medium to be poroelastic, the second step requires additional regularity, namely, that the Biot medium is poroviscoelastic. This is the first weak solution existence result for an FSI problem with nonlinear coupling involving a Biot model for poro(visco)elastic media.

1 Introduction and motivation

In this paper we study a time-dependent nonlinearly coupled fluid-structure interaction problem between the flow of an incompressible, viscous fluid, modeled by the Navier-Stokes equations, and bulk poroviscoelasticity modeled by the Biot equations. Bulk poroviscoelasticity means that the dimensions of the “free fluid flow” domain and the poroviscoelastic medium domain are the same. In particular, in this manuscript we consider a 2D fluid-poroelastic structure interaction (FPSI) problem, which captures the main mathematical difficulties of such coupling, see Fig. 1. The free fluid flow and the Biot poro(visco)elastic medium are coupled across the current location of the interface, which is modeled by a reticular plate that has inertia and elastic energy. A reticular plate is a lattice-type structure characterized by two properties: periodicity and small thickness, where periodicity refers to periodic cells (holes) distributed in all directions [26]. The reticular plate interface is transparent to fluid flow. We are interested in the existence of finite energy weak solutions (of the Leray-Hopf type).

Refer to caption
Figure 1: A sketch of the fluid-poroelastic structure interaction domain.

The problem we study here arises in many applications. In particular, we mention encapsulation of bioartificial organs [70] and blood flow in arteries which are modeled as poro(visco)elastic media to study drug transport through the vascular walls [18, 17, 3]. The reticular plate can be used to capture the elastodynamics behavior of the intima/elastic laminae layer of arterial walls which is in direct contact with the blood flow on one side, and a poroelastic medium consisting of the arterial media/adventitia complex on the other side.

From the mathematical point of view the primary difficulties in studying Navier-Stokes equations nonlinearly coupled to bulk poro(visco)elasticity arise from the fact that the finite energy solutions do not posses sufficient regularity to (1) define the moving domain and the corresponding traces, and (2) guarantee that all the integrals in the weak formulation of the problem are well-defined. The first issue is related to the difficulties associated with fluid-structure coupling, where the fluid and structure domains are of the same dimension. The second issue is a consequence of the geometric nonlinearities associated with moving domain problems. These are the main reasons why to this day there have been no works on the existence of weak solutions for the Biot-Navier-Stokes coupled problems in which the coupling is assumed over a moving interface.

To get around these difficulties, we take the following approaches. First, the reticular plate at the interface associates mass and elastic energy to the interface, and regularizes the boundary of the fluid domain. In classical fluid-structure interaction problems involving elastic structures, this usually takes care of the issues related to the regularity of traces in moving boundary problems [58]. In the case when the structure is poroelastic, and it satisfies the Biot equations on a moving domain, this is, however, not sufficient since the energy estimates do not provide sufficient regularity of the poroelastic matrix displacement for certain integrals over the moving Biot domain in the weak formulation to be well defined. This is why we take the following two-step approach:

  1. 1.

    We introduce a “consistent regularized weak formulation” of the coupled problem by defining a suitable convolution in spatial variables and regularizing only the problematic terms in the weak formulation of the coupled problem. We prove the existence of a weak solution to this regularized problem.

  2. 2.

    We show that as the regularization parameter tends to zero, the solution to this regularized problem converges to the solution of the original nonregularized problem in the case when the original problem has a classical solution and the Biot poroelastic matrix is viscoelastic. Here, a classical solution is a solution that is smooth both temporally and spatially and it hence satisfies the system of PDEs for the original fluid-poroelastic structure problem pointwise.

The existence of a weak solution to the regularized problem was announced by the authors in [44], where only the main steps of the proof were outlined. In particular, the proofs of the existence of weak solutions to the fluid and structure subproblems used in constructing the coupled solution were omitted in [44], and only the main steps of the uniform estimates were presented. Most importantly, the proof of the main compactness result needed to address the main difficulty, the geometric nonlinearity in the regularized problem, is only outlined in [44]. Furthermore, details of the construction of appropriate test functions that are defined on moving domains are also omitted in [44]. Here we present details of all proofs, and show the weak-classical consistency result outlined in step 2 above.

The weak-classical consistency result outlined in step 2 is obtained by using a Gronwall-type estimate, which shows that the energy of the difference between the weak solution of the regularized problem and the classical (temporally and spatially smooth) solution to the original, nonregularized problem with viscoelastic Biot poroelastic matrix, converges to zero as the regularization parameter tends to zero. While the main idea is simple, the estimates are quite nontrivial due to the fact that we need to work with the integrals over regularized Biot domains and compare them with the integrals over the nonregularized moving domains. Details are presented in Section 10.

We conclude this section by noting that the main steps of the constructive proof presented in this manuscript can be used to design a numerical scheme to capture the solutions to the original (non-regularized) FPSI problem, see [62]. The main constructive steps of the proof can be summarized as follows. We semidiscretize the regularized FPSI problem in time by subdividing the time interval into NN subintervals of width Δt\Delta t. At each time step we split the reticular plate subproblem from the regularized fluid-Biot subproblem using a Lie operator splitting strategy [35]. To deal with the moving domains we use the Lagrangian map for the Biot domain, and an Arbitrary Lagrangian-Eulerian mapping for the fluid domain, which maps a fixed, reference domain onto the current, physical domain. We switch between the reference domain formulation and moving domain formulation in the proof as needed. For each Δt\Delta t, approximate solutions are constructed by “solving” the sequence of semidiscretized (linearized) problems defined on the current (approximate) moving domain for each tn=nΔt,n=1,,Nt_{n}=n\Delta t,n=1,\dots,N.

We then show uniform boundedness of the approximate solutions by deriving energy estimates that are uniform in the time discretization parameter Δt\Delta t. This will allow us to deduce the existence of weakly and weakly* convergent subsequences. Since the problem is highly nonlinear, just having weakly and weakly* subsequences is not sufficient to pass to the limit in the weak formulations of the approximate problems. Hence, we must obtain strong convergence of approximate sequences by using several compactness results: the classical Aubin-Lions compactness lemma [57] for the Biot displacement, Arzela-Ascoli for the plate displacement, Dreher and Jüngel’s compactness result [31] for the Biot and plate velocity and pore pressure, and a recent generalized Aubin-Lions-Simon compactness result by Muha and Čanić [52], to deal with the most involved part, which is the free fluid velocity defined on different time-dependent fluid domains.

Once strongly convergent subsequences are obtained from the compactness results, one would like to pass to the limit in the weak formulation to show that the limits of the subsequences are weak solutions to the regularized fluid-poroelastic structure interaction problem. However, this cannot be done yet, since the velocity test functions are also defined on moving domains and we need to construct “appropriate” test functions which can be compared for different domains, and for which we can show convergence to a test function of the limiting, continuous problem. Luckily, in contrast with the classical fluid-elastic structure interaction problems, in our case the fluid test functions decouple from the structure problem, and so it is a bit easier to construct appropriate test functions for which one can show uniform pointwise convergence to a test function for the continuous problem. With this final step, we can pass to the limit in the weak formulations of approximate problems and show that the limits of approximate subsequences satisfy the continuous weak formulation of the regularized problem.

This existence result is local in time because we can guarantee the nondegeneracy of the fluid domains both for the free fluid flow and the filtrating flow through the poroelastic medium only locally in time. However, using the approach presented in [21, Section 5] the time of existence can be extended to the maximal time until one of the following three events occurs: (1) the moving fluid domain or Biot domain degenerates (e.g., the interface touches the bottom of the fluid domain or the top of the Biot domain), (2) the pores in the poroelastic matrix denegerate in the sense that the Lagrangian mapping stops being injective, or (3) T=T=\infty.

2 Literature review

There is extensive past work on fluid-structure interaction (FSI) studying fully coupled systems involving incompressible, Newtonian fluids interacting with deformable structures.

Most of the FSI literature considers models involving purely elastic structures. The models first considered were linearly coupled FSI models [4, 5, 46], which pose the fluid equations on a fixed reference fluid domain, as a linearization that approximates real-life dynamics well when structure displacements and deformations are small.

In cases when displacements and deformations of the structure are large, they can significantly affect the fluid dynamics in which case time-dependent moving fluid domains that depend on the displacement itself must be taken into account. Such nonlinearly coupled FSI models have been extensively studied in [8, 21, 23, 24, 28, 29, 36, 37, 38, 39, 40, 45, 47, 48, 52, 53, 54, 55, 56, 61]. In such models the time-dependent and a priori unknown fluid domain evolves according to the displacement of the structure, giving rise to a fully coupled problem with two-way coupling between the fluid and structure that has significant geometric nonlinearities arising from the moving boundary. There are two broad classes of nonlinearly coupled FSI models: (1) models in which the elastic structure has a lower spatial dimension than the fluid so that the structure is for example an elastic plate or shell, and (2) models in which the fluid and structure domains have the same spatial dimension. In the first case (involving elastic structures of lower spatial dimension), the works showing existence of strong solutions include [8, 48, 37, 38], and the works showing existence of weak solutions include [21, 36, 52, 47, 38]. In the second case (involving coupled elastic structures and fluids of the same spatial dimension), well-posedness results have been studied in [28, 29, 23, 24, 45, 39, 40, 61].

Closest to the work presented in this manuscript is the work of [52] showing existence of weak solutions to a nonlinearly coupled problem between an elastic Koiter shell and an incompressible viscous fluid modeled by the Navier-Stokes equations. In [52] a splitting scheme was introduced to prove the existence of a weak solution to the nonlinearly coupled problem by semidiscretizing the fully coupled problem in time and splitting the coupled problem into fluid and structure subproblems. This scheme has proven to be a robust way for analyzing a variety of complex nonlinearly coupled (moving boundary) FSI problems involving elastic or viscoelastic structures, see [52, 53, 54, 55, 56]. In the present manuscript we adapt the splitting scheme approach to the nonlinearly coupled fluid-poroelastic structure interaction problem.

In terms of literature related to poroelastic media modeled by the Biot equations, we mention the studies by Biot, modeling soil consolidation [9, 10], the studies of fractures in porous and poroelastic materials [34, 49] and more recently, applications to biomedical science, including the study of the ocular tissue related to the onset of glaucoma [19], and the modeling of intestinal walls as poroelastic media [71]. The mathematical well-posedness of the Biot equations discussed in these models has been the focus of a number of works, including [6, 7, 60, 65, 67, 69, 11, 15, 12, 13].

In terms of fluid-poroelastic structure interaction problems, the analysis of well-posedness for linearly coupled problems were discussed in [2, 20, 66]. Recent progress in the design of bioartificial organs, see e.g., [70], sparked the need to study FPSI problems in which the fluid-structure interface itself has mass and elastic or poroelastic energy. The well-posedness for a linearly coupled FPSI problem in which the structure consists of two layers: a thin poroelastic plate located at the interface between the free fluid flow and a thick poroelastic medium modeled by the Biot equations, was obtained in [14] for both the linear and nonlinear Biot equations, where the nonlinearity refers to the dependence of the permeability tensor in the Biot equations on the fluid content. In [14] the fluid-structure interface with mass serves as a regularizing mechanism and provides sufficient information about the regularity of the interface and the free fluid domain to allow, for the first time, the proof of the existence of a finite energy weak solution.

None of the works that address weak solutions to fluid-structure interaction problems between the flow of an incompressible, viscous fluid and a poroelastic solid have taken into account nonlinear coupling over the moving interface. The goal of the current manuscript is to develop a well-posedness theory for a nonlinearly coupled (moving boundary) fluid-poroelastic structure interaction problem by constructing new tools for dealing with the equations of poroelasticity defined on a priori unknown and time-dependent domains.

3 Description of the main problem

We study fluid-poroelastic structure interaction between the flow of an incompressible, viscous fluid and a multilayered poro(visco)elastic structure consisting of two layers: a thick poro(visco)elastic layer modeled by the Biot equations, and a thin elastic layer modeled by the reticular plate equation. The problem is set on a two dimensional domain, which embodies all the main mathematical difficulties associated with the analysis of this problem. The entire two dimensional domain Ω^\hat{\Omega} is a union of the reference domain for the fluid subproblem Ω^f\hat{\Omega}_{f}, the reference domain for the Biot poroviscoelastic material Ω^b\hat{\Omega}_{b}, and the reference domain Γ^\hat{\Gamma} of the elastic reticular plate which serves as the interface separating the free fluid flow and the Biot medium:

Ω^=Ω^bΩ^fΓ^,whereΩ^b=(0,L)×(0,R),Γ^=(0,L)×{0},Ω^f=(0,L)×(R,0).\hat{\Omega}=\hat{\Omega}_{b}\cup\hat{\Omega}_{f}\cup\hat{\Gamma},\ {\rm where}\ \hat{\Omega}_{b}=(0,L)\times(0,R),\ \hat{\Gamma}=(0,L)\times\{0\},\ \hat{\Omega}_{f}=(0,L)\times(-R,0).

These domains will evolve in time, giving rise to the time-dependent Ω(t)=Ωb(t)Ωf(t)Γ(t)\Omega(t)=\Omega_{b}(t)\cup\Omega_{f}(t)\cup\Gamma(t). We will be using the hat notation to denote objects associated with the reference domain. On each subdomain we will consider the following mathematical models.

3.1 The Biot equations on a moving domain

The Biot system consists of the elastodynamics equation, which in this work will be defined on the Lagrangian domain Ω^b\hat{\Omega}_{b}, and the fluid equation, which in this work will be defined on the Eulerian domain Ωb(t){\Omega}_{b}(t). Let 𝜼^:[0,T]×Ω^b2\hat{\boldsymbol{\eta}}:[0,T]\times\hat{\Omega}_{b}\to\mathbb{R}^{2} denote the displacement of the Biot poroviscoelastic matrix from its reference configuration, and let p^:Ω^b\hat{p}:\hat{\Omega}_{b}\to\mathbb{R} denote the fluid pore pressure. To specify the fluid equation given in terms of the fluid pore pressure in Eulerian formulation, we introduce the Lagrangian map by

𝚽^bη(t,)=Id+𝜼^(t,):Ω^bΩb(t),\hat{\boldsymbol{\Phi}}_{b}^{\eta}(t,\cdot)=\text{Id}+\hat{\boldsymbol{\eta}}(t,\cdot):\hat{\Omega}_{b}\to\Omega_{b}(t), (1)

with (𝚽bη)1(t,):Ωb(t)Ω^b(\boldsymbol{\Phi}_{b}^{\eta})^{-1}(t,\cdot):\Omega_{b}(t)\to\hat{\Omega}_{b} denoting its inverse. The Biot equations are then given by:

ρbtt𝜼^\displaystyle\rho_{b}\partial_{tt}\hat{\boldsymbol{\eta}} =^S^b(^𝜼^,p^)\displaystyle=\hat{\nabla}\cdot\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p}) in Ω^b,\displaystyle\text{ in }\hat{\Omega}_{b}, (2)
c0[det(^𝚽^bη)](𝚽bη)1DDtp+αDDt𝜼\displaystyle\frac{c_{0}}{[\det(\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})]\circ(\boldsymbol{\Phi}^{\eta}_{b})^{-1}}\frac{D}{Dt}p+\alpha\nabla\cdot\frac{D}{Dt}\boldsymbol{\eta} =(κp)\displaystyle=\nabla\cdot(\kappa\nabla p) in Ωb(t),\displaystyle\text{ in }\Omega_{b}(t), (3)

where DDt=ddt+((t𝜼(t,)(𝚽bη)1(t,)))\frac{D}{Dt}=\frac{d}{dt}+\left(\left(\partial_{t}\boldsymbol{\eta}(t,\cdot)\circ(\boldsymbol{\Phi}_{b}^{\eta})^{-1}(t,\cdot)\right)\cdot\nabla\right) is the material derivative. The first equation describes the elastodynamics of the poroelastic solid matrix, while the second equation models the conservation of mass principle of the filtrating fluid, see, e.g. [64, 72] for more details about Biot equations defined on moving domains. To recover the filtration fluid velocity 𝒒\boldsymbol{q}, Darcy’s law is used:

𝒒=κp on Ωb(t),\boldsymbol{q}=-\kappa\nabla p\qquad\text{ on }\Omega_{b}(t), (4)

where κ\kappa is a positive permeability constant.

In this work, we will consider both the viscoelastic and the purely elastic consitutive models for the Biot poroelastic matrix with the Piola-Kirchhoff stress tensor for the viscoelastic case given by

S^b(𝜼,p)=2μe𝑫^(𝜼^)+λe(^𝜼^)𝑰+2μv𝑫^(𝜼^t)+λv(^𝜼^t)𝑰αdet(^𝚽^bη)p^(^𝚽^bη)t,\hat{S}_{b}(\nabla\boldsymbol{\eta},p)=2\mu_{e}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}})+\lambda_{e}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}})\boldsymbol{I}+2\mu_{v}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}}_{t})+\lambda_{v}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}}_{t})\boldsymbol{I}-\alpha\det(\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})\hat{p}(\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})^{-t}, (5)

where superscript tt denotes matrix transposition and At=(A1)tA^{-t}=(A^{-1})^{t}. The purely elastic case has the coefficients λv\lambda_{v} and μv\mu_{v} equal to zero. Here, 𝑫\boldsymbol{D} denotes the symmetrized gradient, μe\mu_{e} and λe\lambda_{e} are the Lamé parameters related to the elastic stress, μv\mu_{v} and λv\lambda_{v} are the corresponding parameters related to the viscoelastic stress, and 𝚽^bη\hat{\boldsymbol{\Phi}}^{\eta}_{b} is the Lagrangian map defined above. From the definition of the stress tensor (5), one can see that the elastodynamics of the Biot medium in (2) is described by linear elasticity with an additional term involving pore pressure. This pressure term embodies additional geometric nonlinearities arising from transforming the pressure between the Eulerian and Lagrangian frameworks.

In equation (3) the Biot material displacement 𝜼\boldsymbol{\eta} and the pore pressure pp are defined on the physical domain Ωb(t)\Omega_{b}(t) as

𝜼(t,)=𝜼^(t,(𝚽bη)1(t,)),p(t,)=p^(t,(𝚽bη)1(t,)),whereΩb(t)=𝚽^bη(t,Ω^b).\boldsymbol{\eta}(t,\cdot)=\hat{\boldsymbol{\eta}}(t,(\boldsymbol{\Phi}^{\eta}_{b})^{-1}(t,\cdot)),\qquad p(t,\cdot)=\hat{p}(t,(\boldsymbol{\Phi}_{b}^{\eta})^{-1}(t,\cdot)),\ {\rm where}\ \Omega_{b}(t)=\hat{\boldsymbol{\Phi}}_{b}^{\eta}(t,\hat{\Omega}_{b}).

We remark that in the last term of the Piola-Kirchhoff stress tensor (5), we have used the Piola transform (e.g. [25, Section 1.7.]), which is a transformation that maps tensors in Lagrangian coordinates to corresponding tensors in Eulerian coordinates in such a way that divergence-free tensors in Lagrangian coordinates remain divergence free in Eulerian coordinates [25].

We note that a priori the notion of Ωb(t)\Omega_{b}(t) is not entirely clear, unless 𝜼^\hat{\boldsymbol{\eta}} is sufficiently regular, and furthermore, the formulation of this problem makes sense only if the map 𝚽^bη=Id+𝜼^\hat{\boldsymbol{\Phi}}^{\eta}_{b}=\text{Id}+\hat{\boldsymbol{\eta}} is an injective map from Ω^b\hat{\Omega}_{b} to Ωb(t)\Omega_{b}(t). We address these important issues later.

3.2 The reticular plate equation

A reticular plate is a lattice-type structure characterized by two properties: periodicity and small thickness, where periodicity refers to periodic cells (holes) distributed in all directions [26]. Reticular plates, shells or membranes are models for reticular tissue, which is a connective tissue made up of a network of supportive fibers that provide a framework for soft organs. The elastodynamics of reticular plates, studied in [26] using homogenization, is governed by a plate-type equation, defined on the equilibrium middle surface Γ^\hat{\Gamma} of the homogenized plate or shell. The homogenized equation is given in terms of transverse displacement 𝝎^=ω^𝒆y\hat{\boldsymbol{\omega}}=\hat{\omega}\boldsymbol{e}_{y} from the reference configuration:

ρpttω^+Δ^2ω^=F^p, on Γ^,{{\rho_{p}}}\partial_{tt}\hat{\omega}+\hat{\Delta}^{2}\hat{\omega}=\hat{F}_{p},\qquad\text{ on }\hat{\Gamma}, (6)

where ρp\rho_{p} is the plate density coefficient and F^p\hat{F}_{p} is the external forcing on the plate in yy direction, to be specified later in the coupling conditions. The constant ρp\rho_{p} is the “average” plate density, which depends on the periodic structure. The in-plane bi-Laplacian Δ^2\hat{\Delta}^{2} (Laplace-Beltrami operator for curved Γ^\hat{\Gamma}’s) is associated with the elastic energy of the plate. Typically, there is a coefficient D~\tilde{D} in front of the bi-Laplacian, which contains information about the periodicity of the structure and its stiffness properties [26]. In the present work, without loss of generality, we will assume that it is equal to 11. The source term F^p\hat{F}_{p} corresponds to the loading of the poroelastic plate, which will come from the jump in the normal stress (traction) between the free fluid on one side and the thick Biot poroelastic structure on the other, see (7) below.

In our problem, the reticular plate separates the regions of free fluid flow and the Biot poroviscoelastic medium, and is transparent to the flow between the two. This means, in particular, that there is no resistance to the fluid flow passing through the reticular place. However, due to the inertia and elastic energy of the plate, the analysis of the problem will be simplified due to the regularizing effects of the plate inertia and elastic energy, as we shall see below (see e.g., Remark 5.1).

The time-dependent configuration of the plate

Γ(t)={(x,y):0<x<L,y=ω^(t,x)},\Gamma(t)=\{(x,y):0<x<L,\;y=\hat{\omega}(t,x)\},

forms the bottom boundary of the moving Biot domain Ωb(t)\Omega_{b}(t), and the remaining left, top, and right boundaries of the moving Biot domain Ωb(t)\Omega_{b}(t) are fixed in time. Hence, we impose 𝜼=0\boldsymbol{\eta}=0 on the left, top, and right boundaries of Ωb(t)\Omega_{b}(t). See Fig. 1. Hence, we can describe the moving domain Ωb(t)\Omega_{b}(t) as

Ωb(t)={(x,y):0<x<L,ω^(t,x)<y<R}.\Omega_{b}(t)=\{(x,y):0<x<L,\;\hat{\omega}(t,x)<y<R\}.

3.3 The Navier-Stokes equations on a moving domain

The free flow of an incompressible, viscous fluid will be modeled by the Navier-Stokes equations

t𝒖+(𝒖)𝒖=𝝈f(𝒖,π)𝒖=0} in Ωf(t),\left.\begin{array}[]{rcl}\displaystyle{\partial_{t}\boldsymbol{u}+(\boldsymbol{u}\cdot\nabla)\boldsymbol{u}}&=&\nabla\cdot\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\\ \displaystyle{\nabla\cdot\boldsymbol{u}}&=&0\end{array}\right\}\quad\text{ in }\Omega_{f}(t), (7)

where 𝒖\boldsymbol{u} is the fluid velocity and π\pi is the fluid pressure. The Cauchy stress tensor is given by

𝝈f(𝒖,π)=2ν𝑫(𝒖)π𝑰,\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)=2\nu\boldsymbol{D}(\boldsymbol{u})-\pi\boldsymbol{I},

where π\pi is the fluid pressure and ν\nu is kinematic viscosity coefficient. Notice that the fluid problem is defined on a moving domain, which is not known a priori. The moving fluid domain Ωf(t)\Omega_{f}(t) is a function of time and it is determined by the plate displacement ω^\hat{\omega}, as follows:

Ωf(t)={(x,y):0<x<L,R<y<ω^(t,x)}.\Omega_{f}(t)=\{(x,y):0<x<L,-R<y<\hat{\omega}(t,x)\}.

The fact that the free fluid domain depends on one of the unknowns in the problem presents a geometric nonlinearity that is difficult to deal with. We will be using the following Arbitrary Lagrangian Eulerian (ALE) mapping 𝚽^fω:Ω^fΩf(t)\hat{\boldsymbol{\Phi}}^{\omega}_{f}:\hat{\Omega}_{f}\to\Omega_{f}(t) to map the fixed reference domain Ω^f\hat{\Omega}_{f} onto the current, physical domain Ωf(t)\Omega_{f}(t):

𝚽^fω(x^,y^)=(x^,y^+(1+y^R)ω^),(x^,y^)Ω^f.\hat{\boldsymbol{\Phi}}^{\omega}_{f}(\hat{x},\hat{y})=\left(\hat{x},\hat{y}+\left(1+\frac{\hat{y}}{R}\right)\hat{\omega}\right),\qquad(\hat{x},\hat{y})\in\hat{\Omega}_{f}. (8)

In our analysis, we will use this ALE mapping to will switch between the fixed and moving boundary formulations of the coupled problem as needed.

Remark 3.1.

In numerical computations, it is typical to employ harmonic extension to construct the Arbitrary Lagrangian-Eulerian (ALE) mapping. However, given the simplicity of our geometry, we chose to utilize the explicit formula for extension to simplify the calculations related to the change of variables. Since our methodology is not contingent on the particular selection of the ALE map, in scenarios involving more complex geometries where an explicit formula is not viable, alternatives such as harmonic extension can also be utilized.

3.4 The coupling conditions

The Navier-Stokes equations (7), the Biot equations (2), (3), and the reticular plate equation (6) are coupled across the moving reticular plate interface Γ(t)\Gamma(t) via two sets of coupling conditions: the kinematic and dynamic coupling conditions. To state these conditions, we introduce the following notation:

  • The Biot Cauchy stress tensor defined on the physical domain Sb(𝜼,p){S}_{b}(\nabla\boldsymbol{\eta},p) is obtained by applying the Piola transform to the Biot Cauchy stress tensor S^b(𝜼,p)\hat{S}_{b}(\nabla\boldsymbol{\eta},p) defined on the reference domain, to obtain:

    Sb(𝜼,p)=[det(^𝚽^bη)1S^b(^𝜼^,p^)(^𝚽^bη)t](𝚽bη)1\displaystyle S_{b}(\nabla\boldsymbol{\eta},p)=[\det(\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})^{-1}\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p})(\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})^{t}]\circ(\boldsymbol{\Phi}^{\eta}_{b})^{-1}
    =(1det(^𝚽^bη)[2μe𝑫^(𝜼^)+λe(^𝜼^)+2μv𝑫^(𝜼^t)+λv(^𝜼^t)](^𝚽^bη)t)(𝚽bη)1αp𝑰.\displaystyle=\left(\frac{1}{\det(\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})}\left[2\mu_{e}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}})+\lambda_{e}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}})+2\mu_{v}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}}_{t})+\lambda_{v}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}}_{t})\right](\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})^{t}\right)\circ(\boldsymbol{\Phi}^{\eta}_{b})^{-1}-\alpha p\boldsymbol{I}. (9)
  • The Eulerian structure velocity of the Biot poroviscoelastic matrix is given at each point of the physical domain Ωb(t)\Omega_{b}(t) by

    𝝃(t,)=t𝜼^(t,(𝚽bη)1(t,)).\boldsymbol{\xi}(t,\cdot)=\partial_{t}\hat{\boldsymbol{\eta}}\left(t,(\boldsymbol{\Phi}^{\eta}_{b})^{-1}(t,\cdot)\right). (10)
  • The normal unit vector to the moving interface Γ(t)\Gamma(t) will be denoted by 𝒏(t)\boldsymbol{n}(t), and the normal unit vector to the reference configuration of the interface Γ^\hat{\Gamma} will be denoted by 𝒏^\hat{\boldsymbol{n}}. Note that 𝒏^=𝒆y\hat{\boldsymbol{n}}=\boldsymbol{e}_{y}. The vectors 𝒏(t)\boldsymbol{n}(t) and 𝒏^\hat{\boldsymbol{n}} point outward from Ωf(t)\Omega_{f}(t) and Ωf\Omega_{f}, and inward towards Ωb(t)\Omega_{b}(t) and Ωb\Omega_{b}.

The following two sets of coupling conditions give rise to a well-defined bounded energy of the coupled problem: (I) Kinematic coupling conditions:

  • Continuity of normal components of velocity (conservation of mass of the fluid):

    𝒖𝒏(t)=(𝒒+𝝃)𝒏(t), on (0,T)×Γ(t).\boldsymbol{u}\cdot\boldsymbol{n}(t)=(\boldsymbol{q}+\boldsymbol{\xi})\cdot\boldsymbol{n}(t),\qquad\text{ on }(0,T)\times\Gamma(t). (11)
  • Slip in the tangential component of free fluid velocity, known as the Beavers-Joseph-Saffman condition [42, 43]:

    β(𝝃𝒖)𝝉(t)=𝝈f𝒏(t)𝝉(t), on (0,T)×Γ(t),\beta(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}(t)=\boldsymbol{\sigma}_{f}\boldsymbol{n}(t)\cdot\boldsymbol{\tau}(t),\qquad\text{ on }(0,T)\times\Gamma(t), (12)

    where β0\beta\geq 0 is a constant and 𝝉(t)\boldsymbol{\tau}(t) is the rightward pointing unit tangent vector to Γ(t)\Gamma(t).

  • Continuity of displacements:

    𝜼^=ω^𝒆y, on (0,T)×Γ^.\hat{\boldsymbol{\eta}}=\hat{\omega}\boldsymbol{e}_{y},\qquad\text{ on }(0,T)\times\hat{\Gamma}. (13)

(II) Dynamic coupling conditions:

  • Balance of forces describing the body forcing on the plate as the difference between the normal components of normal stress coming from the Biot medium on one side, and free fluid flow on the other:

    F^p=det(𝚽^fω)[𝝈f(𝒖,π)𝚽^fω](𝚽^fω)t𝒏^𝒏^+S^b(^𝜼^,p^)𝒏^𝒏^|Γ^, on (0,T)×Γ^,\hat{F}_{p}=-\det(\nabla\hat{\boldsymbol{\Phi}}^{\omega}_{f})[\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\circ\hat{\boldsymbol{\Phi}}^{\omega}_{f}](\nabla\hat{\boldsymbol{\Phi}}^{\omega}_{f})^{-t}\hat{\boldsymbol{n}}\cdot\hat{\boldsymbol{n}}+\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p})\hat{\boldsymbol{n}}\cdot\hat{\boldsymbol{n}}|_{\hat{\Gamma}},\qquad\text{ on }{{(0,T)\times\hat{\Gamma}}}, (14)

    where 𝚽^fω\hat{\boldsymbol{\Phi}}^{\omega}_{f} is the Arbitrary Lagrangian-Eulerian (ALE) mapping defined in (16).

  • Balance of pressure at the interface:

    𝝈f(𝒖,π)𝒏(t)𝒏(t)+12|𝒖|2=p, on (0,T)×Γ(t).-\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}(t)\cdot\boldsymbol{n}(t)+\frac{1}{2}|\boldsymbol{u}|^{2}=p,\qquad\text{ on }(0,T)\times\Gamma(t). (15)

    See [52, 68, 30, 27] for the use of the dynamic (total) pressure π+12|𝒖|2\pi+\frac{1}{2}|\boldsymbol{u}|^{2} on the left-hand side of (15).

3.5 The initial and boundary conditions

For the fluid, we will assume rigid walls on Ωf(t)Γ(t)\partial\Omega_{f}(t)\setminus\Gamma(t) and impose a no-slip condition

𝒖=0, on Ωf(t)Γ(t).\boldsymbol{u}=0,\qquad\text{ on }\partial\Omega_{f}(t)\setminus\Gamma(t).

Similarly, we will assume that the boundaries of the Biot poroviscoelastic medium, excluding the interface Γ(t)\Gamma(t), are rigid and impose

𝜼^=0 and p^=0, on Ω^bΓ^.\hat{\boldsymbol{\eta}}=0\ \ \text{ and }\ \ \hat{p}=0,\qquad\text{ on }\partial\hat{\Omega}_{b}\setminus\hat{\Gamma}.

Finally, we prescribe the following initial conditions:

𝒖(0)=𝒖0inΩf(0),\displaystyle\boldsymbol{u}(0)=\boldsymbol{u}_{0}\quad{\rm in}\;\Omega_{f}(0),
𝜼^(0)=𝜼^0,t𝜼^(0)=𝝃^0inΩ^b,\displaystyle\hat{\boldsymbol{\eta}}(0)=\hat{\boldsymbol{\eta}}_{0},\;\partial_{t}\hat{\boldsymbol{\eta}}(0)=\hat{\boldsymbol{\xi}}_{0}\quad{\rm in}\;\hat{\Omega}_{b},
ω^(0)=ω^0,tω^(0)=ζ^0inΓ^,\displaystyle\hat{\omega}(0)=\hat{\omega}_{0},\;\partial_{t}\hat{\omega}(0)=\hat{\zeta}_{0}\quad{\rm in}\;\hat{\Gamma},
p^(0)=p^0inΩ^b.\displaystyle\hat{p}(0)=\hat{p}_{0}\quad{\rm in}\;\hat{\Omega}_{b}.

3.6 Preview of the main results

Our first main result is the existence of a weak solution to a regularized FPSI problem, where there is a regularization parameter δ>0\delta>0. The regularization will involve spatially regularizing the Biot displacement by extending the displacement 𝜼^\hat{\boldsymbol{\eta}} on Ω^b\hat{\Omega}_{b} to a larger domain and using spatial convolution by a smooth compactly supported kernel, scaled by δ\delta. This regularized FPSI problem will be introduced in Sec. 5. The existence result for the regularized FPSI problem holds for both elastic and viscoelastic Biot material. Here we state the theorem informally and refer the reader to Theorem 5.1 for the precise statement.

Theorem 3.1.

[Existence of a weak solution to the regularized problem] Let ρb,μe,λe,α,ρp,ν>0\rho_{b},\mu_{e},\lambda_{e},\alpha,\rho_{p},\nu>0 and μv,λv0\mu_{v},\lambda_{v}\geq 0. Moreover, assume that initial data are in the finite energy class and that initially, the interface does not touch the bottom boundary of the fluid domain and the top boundary of the Biot domain, and assume that certain compatibility conditions are satisfied. Then for every regularization parameter δ>0\delta>0, there exists T>0T>0 (potentially depending on δ>0\delta>0) such that there is a weak solution on [0,T][0,T] to the regularized problem with regularization parameter δ\delta. Furthermore, the weak solution to the regularized problem exists on a maximal time interval [0,T][0,T], where either (1) T=T=\infty or (2) TT is finite and is the time at which either:

  • the fluid or Biot domain degenerates so that the moving interface collides with the bottom boundary of Ω^f\hat{\Omega}_{f} or the top boundary of Ω^b\hat{\Omega}_{b} or

  • the (regularized) Lagrangian mapping 𝚽^bηδ\hat{\boldsymbol{\Phi}}^{\eta^{\delta}}_{b} for the Biot domain is no longer injective.

Our second main result is a weak-classical consistency result. Namely, in order to justify our regularization procedure and the corresponding definition of weak solutions to the regularized problem, we prove that weak solutions to the regularized problem indeed converge to the solution of the original (non-regularized) FPSI problem. More precisely, we prove the following result, made precise in Theorem 10.1.

Theorem 3.2.

[Weak-classical consistency] Assume that a classical (smooth) solution to the FPSI problem with a Biot poroviscoelastic medium exists on time-interval [0,T] for the case for which the viscoelasticity parameters μv,λv>0\mu_{v},\lambda_{v}>0. Then every sequence of weak solutions to the regularized problem with regularization parameter δ>0\delta>0 converges to the classical solution on [0,T][0,T] as the regularization parameter δ\delta converges to 0. In particular, the time interval of existence for the weak solutions to the regularized problem is uniform in regularization parameter and solutions to the regularized problem exist on the same time interval where the classical solution exists.

Remark 3.2.

An alternative formulation for Theorems 3.1 and 3.2 is that there exists a weak solution to an approximate problem of the original FPSI problem. Specifically, Theorem 3.2 asserts that the strong solution can be approximated by solutions to the regularized problem, the existence of which is guaranteed by Theorem 3.1.

The heart of the proof of this theorem is a bootstrap argument presented in Section 10.4. Namely, the main issue is that geometric quantities, such as the determinant of the displacement, cannot be estimated by the energy, and thus are not uniformly bounded in the regularization parameter δ\delta. We derive appropriate bounds by using a bootstrap argument in combination with optimal convergence rate estimates for the convolution regularization. The main technical issue in comparing the classical solution with weak solutions to the regularized problem is the fact that they are defined on different domains. Therefore, we use a change of variables that transfers fluid velocities as vector fields and preserves the divergence-free condition. This transformation was introduced by [41] and was used in proving weak-strong type of results in the context of FSI in [22, 59, 63]. The corresponding estimates are carried out in Section 10.3.

4 Definition of a weak solution

Because the problem under consideration is nonlinearly coupled, the fluid domain Ωf(t)\Omega_{f}(t) and the Biot poroviscoelastic domain Ωb(t)\Omega_{b}(t) in physical space are time-dependent and not known apriori. To handle the moving domains, it is useful to introduce the mappings that map the reference domains Ω^b\hat{\Omega}_{b}, Γ^\hat{\Gamma}, and Ω^f\hat{\Omega}_{f} onto the moving domains that depend on time and on the solution itself.

4.1 Mappings between reference and physical domains

Let

𝚽^bη(t,):Ω^bΩb(t),𝚽^Γω(t,):Γ^Γ(t),𝚽^fω(t,):Ω^fΩf(t),\hat{\boldsymbol{\Phi}}^{\eta}_{b}(t,\cdot):\hat{\Omega}_{b}\to\Omega_{b}(t),\qquad\hat{\boldsymbol{\Phi}}^{\omega}_{\Gamma}(t,\cdot):\hat{\Gamma}\to\Gamma(t),\qquad\hat{\boldsymbol{\Phi}}^{\omega}_{f}(t,\cdot):\hat{\Omega}_{f}\to\Omega_{f}(t),

be such that

𝚽^bη=Id+𝜼^(x^,y^),(x^,y^)Ω^b𝚽^Γω(x^,0)=(x^,ω^(x^)),x^Γ^𝚽^fω(x^,y^)=(x^,y^+(1+y^R)ω^(x^)),(x^,y^)Ω^f,\begin{array}[]{lll}&\hat{\boldsymbol{\Phi}}_{b}^{\eta}=\text{Id}+\hat{\boldsymbol{\eta}}(\hat{x},\hat{y}),&(\hat{x},\hat{y})\in\hat{\Omega}_{b}\\ &\hat{\boldsymbol{\Phi}}^{\omega}_{\Gamma}(\hat{x},0)=(\hat{x},\hat{\omega}(\hat{x})),&{{\hat{x}\in\hat{\Gamma}}}\\ &\hat{\boldsymbol{\Phi}}^{\omega}_{f}(\hat{x},\hat{y})=\left(\hat{x},\hat{y}+\left(1+\frac{\hat{y}}{R}\right)\hat{\omega}(\hat{x})\right),&(\hat{x},\hat{y})\in\hat{\Omega}_{f},\end{array} (16)

with the inverse

(𝚽fω)1(x,y)=(x,R+RR+ω^(R+y)).(\boldsymbol{\Phi}^{\omega}_{f})^{-1}(x,y)=\left(x,-R+\frac{R}{R+\hat{\omega}}(R+y)\right). (17)

We are using (x^,y^)(\hat{x},\hat{y}) to denote the coordinates on the reference domain and (x,y)(x,y) the coordinates on the physical domain. Note that these mapings are time-dependent, even though in the rest of this manuscript we will not explicitly notate this time dependence for ease of notation.

The Jacobians of the transformations are given by:

𝒥^fω=1+ω^R,𝒥^bη=det(𝑰+^𝜼^),𝒥^Γω=1+|x^ω^|2,\begin{array}[]{lll}\hat{\mathcal{J}}^{\omega}_{f}=1+\frac{\hat{\omega}}{R},\qquad\hat{\mathcal{J}}^{\eta}_{b}=\det(\boldsymbol{I}+\hat{\nabla}\hat{\boldsymbol{\eta}}),\qquad\hat{\mathcal{J}}^{\omega}_{\Gamma}=\sqrt{1+|\partial_{\hat{x}}\hat{\omega}|^{2}},\end{array} (18)

where 𝒥^Γω\hat{\mathcal{J}}^{\omega}_{\Gamma} measures the arc length difference of between the reference and deformed configuration of the plate. Notice that in the Jacobian 𝒥^fω\hat{\mathcal{J}}^{\omega}_{f} we dropped the absolute value sign since our results will hold up until the time of domain degeneracy when |ω^|R|\hat{\omega}|\geq R.

Under these mappings the functions are transformed as follows.

Tranformations under 𝚽fω\boldsymbol{\Phi}^{\omega}_{f}. The fluid velocity 𝒖\boldsymbol{u} defined on Ωf(t)\Omega_{f}(t) is transferred to the fixed reference domain Ω^f\hat{\Omega}_{f} by

𝒖^(t,x^,y^)=𝒖𝚽^f, for (x^,y^)Ω^f.\hat{\boldsymbol{u}}(t,\hat{x},\hat{y})=\boldsymbol{u}\circ\hat{\boldsymbol{\Phi}}_{f},\qquad\text{ for }(\hat{x},\hat{y})\in\hat{\Omega}_{f}.

Recall that on the moving domain Ωf(t)\Omega_{f}(t), the fluid velocity 𝒖\boldsymbol{u} is divergence free, i.e., 𝒖=0\nabla\cdot\boldsymbol{u}=0. However, when we pull the fluid velocity back to the reference domain, 𝒖^\hat{\boldsymbol{u}} is not necessarily divergence free on Ω^f\hat{\Omega}_{f}. Hence, we want to reformulate the divergence free condition on the fixed reference domain.

The divergence free condition. Let gg be a function defined on Ωf(t)\Omega_{f}(t), then

g=(g^(𝚽fω)1)=(^fωg^)(𝚽fω)1,\nabla g=\nabla\left(\hat{g}\circ(\boldsymbol{\Phi}^{\omega}_{f})^{-1}\right)=(\hat{\nabla}^{\omega}_{f}\hat{g})\circ(\boldsymbol{\Phi}^{\omega}_{f})^{-1},

where ^fω\hat{\nabla}^{\omega}_{f} is the transformed gradient operator:

^fω=(x^(R+y)x^ω^R(R+ω^)2y^RR+ω^y^)wherey=y^+(1+y^R)ω^.\hat{\nabla}^{\omega}_{f}=\begin{pmatrix}\partial_{\hat{x}}-(R+y)\partial_{\hat{x}}\hat{\omega}\frac{R}{(R+\hat{\omega})^{2}}\partial_{\hat{y}}\\ \frac{R}{R+\hat{\omega}}\partial_{\hat{y}}\end{pmatrix}\quad{\rm{where}}\quad y=\hat{y}+\left(1+\frac{\hat{y}}{R}\right)\hat{\omega}. (19)

Therefore, the divergence free condition and the symmetrized gradient on the fixed reference domain Ω^f\hat{\Omega}_{f} are:

^fω𝒖^=0,𝑫^fω(𝒖^)=12(^fω𝒖^+(^fω𝒖^)t).\hat{\nabla}^{\omega}_{f}\cdot\hat{\boldsymbol{u}}=0,\qquad\displaystyle\hat{\boldsymbol{D}}^{\omega}_{f}(\hat{\boldsymbol{u}})=\frac{1}{2}\left(\hat{\nabla}^{\omega}_{f}\hat{\boldsymbol{u}}+(\hat{\nabla}^{\omega}_{f}\hat{\boldsymbol{u}})^{t}\right).

Time derivatives. The time derivative transforms under the map 𝚽^fω\hat{\boldsymbol{\Phi}}^{\omega}_{f} as follows:

t𝒖=t𝒖^(𝒘^^fω)𝒖^where𝒘^=R+y^Rtω^𝒆y.\partial_{t}\boldsymbol{u}=\partial_{t}\hat{\boldsymbol{u}}-(\hat{\boldsymbol{w}}\cdot\hat{\nabla}^{\omega}_{f})\hat{\boldsymbol{u}}\quad{\rm{where}}\quad\hat{\boldsymbol{w}}=\frac{R+\hat{y}}{R}\partial_{t}\hat{\omega}\boldsymbol{e}_{y}. (20)

Tranformations under 𝚽bω\boldsymbol{\Phi}^{\omega}_{b}. Given a scalar function gg defined on Ωb(t)\Omega_{b}(t) the pull back of gg to the reference domain Ω^b\hat{\Omega}_{b} is given by

g^=g𝚽^bη.\hat{g}=g\circ\hat{\boldsymbol{\Phi}}^{\eta}_{b}.

We claim that for some differential operator ^bη\hat{\nabla}^{\eta}_{b}, which we will determine below,

g=(g^(𝚽bη)1)=(^bηg^)(𝚽bη)1,\nabla g=\nabla\left(\hat{g}\circ(\boldsymbol{\Phi}^{\eta}_{b})^{-1}\right)=(\hat{\nabla}_{b}^{\eta}\hat{g})\circ(\boldsymbol{\Phi}^{\eta}_{b})^{-1},

where \nabla is a gradient on the physical domain, ^\hat{\nabla} is a gradient on the reference domain, and ^bη\hat{\nabla}^{\eta}_{b} is a differential operator (different from ^\hat{\nabla}) on the reference domain. For any function gg defined on the physical domain, we have that

^(g𝚽^bη)=[(g)𝚽^bη](𝑰+^𝜼^).\hat{\nabla}\left(g\circ\hat{\boldsymbol{\Phi}}^{\eta}_{b}\right)=[(\nabla g)\circ\hat{\boldsymbol{\Phi}}^{\eta}_{b}]\cdot(\boldsymbol{I}+\hat{\nabla}\hat{\boldsymbol{\eta}}).

Hence, for

^bηg^=(g)𝚽^bη,\hat{\nabla}^{\eta}_{b}\hat{g}=(\nabla g)\circ\hat{\boldsymbol{\Phi}}^{\eta}_{b},

we get the following explicit formula for the transformed gradient operator ^bη\hat{\nabla}^{\eta}_{b} on Ω^b\hat{\Omega}_{b}:

^bηg^=(g^x^,g^y^)(𝑰+^𝜼^)1.\hat{\nabla}^{\eta}_{b}\hat{g}=\left(\frac{\partial\hat{g}}{\partial\hat{x}},\frac{\partial\hat{g}}{\partial\hat{y}}\right)\cdot(\boldsymbol{I}+\hat{\nabla}\hat{\boldsymbol{\eta}})^{-1}. (21)

Notice that the invertibility of the matrix 𝑰+^𝜼^\boldsymbol{I}+\hat{\nabla}\hat{\boldsymbol{\eta}} will be related to whether the map (x^,y^)(x^,y^)+𝜼^(x^,y^)(\hat{x},\hat{y})\to(\hat{x},\hat{y})+\hat{\boldsymbol{\eta}}(\hat{x},\hat{y}) is a bijection between Ω^b\hat{\Omega}_{b} and Ωb(t)\Omega_{b}(t).

4.2 Weak solution

We now derive the definition of a weak solution to the given FPSI problem, by means of the following formal calculation. We start with the fluid equations and multiply by a test function 𝒗\boldsymbol{v}. Recall the definition of the Eulerian structure velocity 𝝃\boldsymbol{\xi} from (10). For the inertia term of the Navier-Stokes equations, using the Reynold’s transport theorem and integration by parts, we obtain:

Ωf(t)(t𝒖+(𝒖)𝒖))𝒗=ddtΩf(t)𝒖𝒗Ωf(t)𝒖t𝒗Γ(t)(𝝃𝒏)𝒖𝒗+12Ωf(t)[((𝒖)𝒖)𝒗(𝒖)𝒗)𝒖]+12Γ(t)(𝒖𝒏)𝒖𝒗=ddtΩf(t)𝒖𝒗Ωf(t)𝒖t𝒗+12Ωf(t)[((𝒖)𝒖)𝒗((𝒖)𝒗)𝒖]+12Γ(t)(𝒖𝒏2𝝃𝒏)𝒖𝒗.\int_{\Omega_{f}(t)}(\partial_{t}\boldsymbol{u}+(\boldsymbol{u}\cdot\nabla)\boldsymbol{u}))\cdot\boldsymbol{v}=\frac{d}{dt}\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\boldsymbol{v}-\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\partial_{t}\boldsymbol{v}-\int_{\Gamma(t)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\boldsymbol{v}\\ +\frac{1}{2}\int_{\Omega_{f}(t)}[((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot\boldsymbol{v}-(\boldsymbol{u}\cdot\nabla)\boldsymbol{v})\cdot\boldsymbol{u}]+\frac{1}{2}\int_{\Gamma(t)}(\boldsymbol{u}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\boldsymbol{v}\\ =\frac{d}{dt}\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\boldsymbol{v}-\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\partial_{t}\boldsymbol{v}+\frac{1}{2}\int_{\Omega_{f}(t)}[((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot\boldsymbol{v}-((\boldsymbol{u}\cdot\nabla)\boldsymbol{v})\cdot\boldsymbol{u}]+\frac{1}{2}\int_{\Gamma(t)}(\boldsymbol{u}\cdot\boldsymbol{n}-2\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\boldsymbol{v}.

For the diffusive term of the Navier Stokes equations, we integrate by parts to obtain

Ωf(t)(𝝈f(𝒖,π))𝒗=2νΩf(t)𝑫(𝒖):𝑫(𝒗)Γ(t)𝝈f(𝒖,π)𝒏𝒗,-\int_{\Omega_{f}(t)}(\nabla\cdot\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi))\cdot\boldsymbol{v}=2\nu\int_{\Omega_{f}(t)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{v})-\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot\boldsymbol{v},

where we used the fact that the test function 𝒗\boldsymbol{v} is divergence free to eliminate the pressure, and we use that the test function satisfies 𝒗=0\boldsymbol{v}=0 on Ωf(t)Γ(t)\partial\Omega_{f}(t)\setminus\Gamma(t) due to the boundary conditions for 𝒖\boldsymbol{u}.

Next, we multiply the structure equation (2) by a test function 𝝍^\hat{\boldsymbol{\psi}} to obtain

Ω^b(ρbtt𝜼^^S^b(^𝜼^,p^))𝝍^=ρb(ddtΩ^bt𝜼^𝝍^Ωbt𝜼^t𝝍^)\displaystyle\int_{\hat{\Omega}_{b}}(\rho_{b}\partial_{tt}\hat{\boldsymbol{\eta}}-\hat{\nabla}\cdot\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p}))\cdot\hat{\boldsymbol{\psi}}=\rho_{b}\left(\frac{d}{dt}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\hat{\boldsymbol{\psi}}-\int_{\Omega_{b}}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\partial_{t}\hat{\boldsymbol{\psi}}\right)
+Ω^bS^b(^𝜼^,p^):^𝝍^+Γ^S^b(^𝜼^,p^)𝒆y𝝍^=ρb(ddtΩ^bt𝜼^𝝍^Ω^bt𝜼^t𝝍^)\displaystyle+\int_{\hat{\Omega}_{b}}\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p}):\hat{\nabla}\hat{\boldsymbol{\psi}}+\int_{\hat{\Gamma}}\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p})\boldsymbol{e}_{y}\cdot\hat{\boldsymbol{\psi}}=\rho_{b}\left(\frac{d}{dt}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\hat{\boldsymbol{\psi}}-\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\partial_{t}\hat{\boldsymbol{\psi}}\right)
+Ω^b(2μe𝑫^(𝜼^):𝑫^(𝝍^)+λe(^𝜼^)(^𝝍^)+2μv𝑫^(t𝜼^):𝑫^(𝝍^)+λv(^t𝜼^)(^𝝍^))\displaystyle+\int_{\hat{\Omega}_{b}}(2\mu_{e}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})+\lambda_{e}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}})+2\mu_{v}\hat{\boldsymbol{D}}(\partial_{t}\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})+\lambda_{v}(\hat{\nabla}\cdot\partial_{t}\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}}))
αΩb(t)p(𝝍)+Γ^S^b(𝜼^,p^)𝒆r𝝍^.\displaystyle-\alpha\int_{\Omega_{b}(t)}p(\nabla\cdot\boldsymbol{\psi})+\int_{\hat{\Gamma}}\hat{S}_{b}(\nabla\hat{\boldsymbol{\eta}},\hat{p})\boldsymbol{e}_{r}\cdot\hat{\boldsymbol{\psi}}.

Except on Γ^\hat{\Gamma}, there are no boundary terms, because 𝜼^=0\hat{\boldsymbol{\eta}}=0 on the left, top, and right boundaries of Ω^b\hat{\Omega}_{b}, and hence the same condition holds for the corresponding test function 𝝍^\hat{\boldsymbol{\psi}}. Note that in the integral over Ωb(t)\Omega_{b}(t), 𝝍:=𝝍^(𝚽bη)1\boldsymbol{\psi}:=\hat{\boldsymbol{\psi}}\circ(\boldsymbol{\Phi}^{\eta}_{b})^{-1}.

Finally, we test the second equation (3) corresponding to the evolution of the pore pressure for the Biot poroviscoelastic medium with a test function rr, and recall the definition of the Darcy velocity 𝒒\boldsymbol{q} from (4), keeping in mind that 𝒏\boldsymbol{n} is the inward normal vector to Ωb(t)\Omega_{b}(t):

Ωb(t)(c0[det(^𝚽^bη)](𝚽bη)1DDtp+αDDt𝜼(κp))r\displaystyle\int_{\Omega_{b}(t)}\left(\frac{c_{0}}{[\det(\hat{\nabla}\hat{\boldsymbol{\Phi}}^{\eta}_{b})]\circ(\boldsymbol{\Phi}_{b}^{\eta})^{-1}}\frac{D}{Dt}p+\alpha\nabla\cdot\frac{D}{Dt}\boldsymbol{\eta}-\nabla\cdot(\kappa\nabla p)\right)r
=Ω^bc0tp^r^+Ωb(t)α(DDt𝜼)r+Ωb(t)κprΓ(t)(𝒒𝒏)r\displaystyle=\int_{\hat{\Omega}_{b}}c_{0}\partial_{t}\hat{p}\cdot\hat{r}+\int_{\Omega_{b}(t)}\alpha\left(\nabla\cdot\frac{D}{Dt}\boldsymbol{\eta}\right)r+\int_{\Omega_{b}(t)}\kappa\nabla p\cdot\nabla r-\int_{\Gamma(t)}(\boldsymbol{q}\cdot\boldsymbol{n})r
=ddtΩ^bc0p^r^Ω^bc0p^tr^Ωb(t)αDDt𝜼rαΓ(t)(𝝃𝒏)r+Ωb(t)κprΓ(t)(𝒒𝒏)r.\displaystyle=\frac{d}{dt}\int_{\hat{\Omega}_{b}}c_{0}\hat{p}\cdot\hat{r}-\int_{\hat{\Omega}_{b}}c_{0}\hat{p}\cdot\partial_{t}\hat{r}-\int_{\Omega_{b}(t)}\alpha\frac{D}{Dt}\boldsymbol{\eta}\cdot\nabla r-\alpha\int_{\Gamma(t)}(\boldsymbol{\xi}\cdot\boldsymbol{n})r+\int_{\Omega_{b}(t)}\kappa\nabla p\cdot\nabla r-\int_{\Gamma(t)}(\boldsymbol{q}\cdot\boldsymbol{n})r.

There are no boundary terms except on Γ(t)\Gamma(t) from the integration by parts in the integral involving α\alpha and in the integral involving κ\kappa because of the Dirichlet boundary condition r=0r=0 (since p=0p=0) on the left, top, and right boundaries of Ω^b\hat{\Omega}_{b}.

After adding the two stress terms, and recalling the definition of 𝚽^Γω\hat{\boldsymbol{\Phi}}^{\omega}_{\Gamma} in (16) and 𝒥^Γω\hat{\mathcal{J}}^{\omega}_{\Gamma} in (18) we obtain:

Γ(t)𝝈f(𝒖,π)𝒏𝒗+Γ^S^b(^𝜼^,p^)𝒆y𝝍^=Γ(t)𝝈f(𝒖,π)𝒏(𝝍𝒗)+Γ^(S^b(^𝜼^,p^)𝒆y𝒥^Γω(𝝈f(𝒖,π)𝒏|Γ(t)𝚽Γω)𝝍^.-\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot\boldsymbol{v}+\int_{\hat{\Gamma}}\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p})\boldsymbol{e}_{y}\cdot\hat{\boldsymbol{\psi}}\\ =\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot(\boldsymbol{\psi}-\boldsymbol{v})+\int_{\hat{\Gamma}}(\hat{S}_{b}(\hat{\nabla}\hat{\boldsymbol{\eta}},\hat{p})\boldsymbol{e}_{y}-\hat{\mathcal{J}}^{\omega}_{\Gamma}\cdot(\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}|_{\Gamma(t)}\circ\boldsymbol{\Phi}^{\omega}_{\Gamma})\cdot\hat{\boldsymbol{\psi}}.

Since the displacement of the plate is only in the yy direction so that 𝜼^=ω^𝒆y\hat{\boldsymbol{\eta}}=\hat{\omega}\boldsymbol{e}_{y} on Γ^\hat{\Gamma}, the test function 𝝍^\hat{\boldsymbol{\psi}} points in the yy direction on Γ^\hat{\Gamma} as well. We will denote by φ^\hat{\varphi} the magnitude of 𝝍^|Γ^\hat{\boldsymbol{\psi}}|_{\hat{\Gamma}} so that 𝝍^=φ^𝒆y\hat{\boldsymbol{\psi}}=\hat{\varphi}\boldsymbol{e}_{y} on Γ^\hat{\Gamma}. By the dynamic coupling condition (14), we have that the previous expression is equal to

=Γ(t)𝝈f(𝒖,π)𝒏(𝝍𝒗)+Γ^F^pφ^=Γ(t)𝝈f(𝒖,π)𝒏(𝝍𝒗)+Γ^(ρpttω^+Δ^2ω^)φ^\displaystyle=\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot(\boldsymbol{\psi}-\boldsymbol{v})+\int_{\hat{\Gamma}}\hat{F}_{p}\cdot\hat{\varphi}=\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot(\boldsymbol{\psi}-\boldsymbol{v})+\int_{\hat{\Gamma}}(\rho_{p}\partial_{tt}\hat{\omega}+\hat{\Delta}^{2}\hat{\omega})\hat{\varphi}
=Γ(t)𝝈f(𝒖,π)𝒏𝒏(ψnvn)+Γ(t)𝝈f(𝒖,π)𝒏𝝉(ψτvτ)+Γ^(ρpttω^+Δ^2ω^)φ^\displaystyle=\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot\boldsymbol{n}(\psi_{n}-v_{n})+\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot\boldsymbol{\tau}(\psi_{\tau}-v_{\tau})+\int_{\hat{\Gamma}}(\rho_{p}\partial_{tt}\hat{\omega}+\hat{\Delta}^{2}\hat{\omega})\hat{\varphi}
=Γ(t)𝝈f(𝒖,π)𝒏𝒏(ψnvn)+Γ(t)β(𝝃𝒖)𝝉(ψτvτ)+Γ^(ρpttω^+Δ^2ω^)φ^\displaystyle=\int_{\Gamma(t)}\boldsymbol{\sigma}_{f}(\nabla\boldsymbol{u},\pi)\boldsymbol{n}\cdot\boldsymbol{n}(\psi_{n}-v_{n})+\int_{\Gamma(t)}\beta(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}(\psi_{\tau}-v_{\tau})+\int_{\hat{\Gamma}}(\rho_{p}\partial_{tt}\hat{\omega}+\hat{\Delta}^{2}\hat{\omega})\hat{\varphi}
=Γ(t)(12|𝒖|2p)(ψnvn)+Γ(t)β(𝝃𝒖)𝝉(ψτvτ)\displaystyle=\int_{\Gamma(t)}\left(\frac{1}{2}|\boldsymbol{u}|^{2}-p\right)(\psi_{n}-v_{n})+\int_{\Gamma(t)}\beta(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}(\psi_{\tau}-v_{\tau})
+ddt(Γ^ρptω^φ^)Γ^ρptω^tφ^+Γ^Δ^ω^Δ^φ^,\displaystyle+\frac{d}{dt}\left(\int_{\hat{\Gamma}}\rho_{p}\partial_{t}\hat{\omega}\cdot\hat{\varphi}\right)-\int_{\hat{\Gamma}}\rho_{p}\partial_{t}\hat{\omega}\cdot\partial_{t}\hat{\varphi}+\int_{\hat{\Gamma}}\hat{\Delta}\hat{\omega}\cdot\hat{\Delta}\hat{\varphi},

where we used the coupling conditions (12) and (15) in the last step. For clarity, we note that in the preceding calculation and in the remainder of the manuscript, 𝒏\boldsymbol{n} denotes the unit normal vector along Γ(t)\Gamma(t) that points upward towards Ωb(t)\Omega_{b}(t), and 𝝉\boldsymbol{\tau} denotes the unit tangential vector along Γ(t)\Gamma(t) that points to the right.

The weak formulation then follows by summing everything together. Before we state the definition of a weak solution to our FPSI problem, we introduce the following notation. Let ζ^\hat{\zeta} denote the transverse velocity of the plate so that

tω^=ζ^,\partial_{t}\hat{\omega}=\hat{\zeta}, (22)

and let ζ=ζ^(𝚽Γω)1\zeta=\hat{\zeta}\circ(\boldsymbol{\Phi}^{\omega}_{\Gamma})^{-1}.

Definition 4.1.

The ordered four-tuple (𝒖,ω^,𝜼^,p)(\boldsymbol{u},\hat{\omega},\hat{\boldsymbol{\eta}},p) satisfies the weak formulation to the nonlinearly coupled FPSI problem if for every test function (𝒗,φ^,𝝍^,r)(\boldsymbol{v},\hat{\varphi},\hat{\boldsymbol{\psi}},r) that is Cc1C^{1}_{c} in time on [0,T)[0,T) taking values in the test space, satisfying 𝝍^=φ^𝒆y\hat{\boldsymbol{\psi}}=\hat{\varphi}\boldsymbol{e}_{y} on Γ^\hat{\Gamma}, we have that

0TΩf(t)𝒖t𝒗+120TΩf(t)[((𝒖)𝒖)𝒗((𝒖)𝒗)𝒖]+120TΓ(t)(𝒖𝒏2ζ𝒆y𝒏)𝒖𝒗+2ν0TΩf(t)𝑫(𝒖):𝑫(𝒗)+0TΓ(t)(12|𝒖|2p)(ψnvn)+β0TΓ(t)(ζ𝒆y𝒖)𝝉(𝝍𝒗)𝝉ρp0TΓ^tω^tφ^+0TΓ^Δ^ω^Δ^φ^ρb0TΩ^bt𝜼^t𝝍^+2μe0TΩ^b𝑫^(𝜼^):𝑫^(𝝍^)+λe0TΩ^b(^𝜼^)(^𝝍^)+2μv0TΩ^b𝑫^(t𝜼^):𝑫^(𝝍^)+λv0TΩ^b(^t𝜼^)(^𝝍^)α0TΩb(t)p𝝍c00TΩ^bp^tr^α0TΩb(t)DDt𝜼rα0TΓ(t)(ζ𝒆y𝒏)r+κ0TΩb(t)pr0TΓ(t)((𝒖ζ𝒆y)𝒏)r=Ωf(0)𝒖(0)𝒗(0)+ρpΓ^tω^(0)φ^(0)+ρbΩ^bt𝜼^(0)𝝍^(0)+c0Ω^bp^(0)r^(0).-\int_{0}^{T}\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\partial_{t}\boldsymbol{v}+\frac{1}{2}\int_{0}^{T}\int_{\Omega_{f}(t)}[((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot\boldsymbol{v}-((\boldsymbol{u}\cdot\nabla)\boldsymbol{v})\cdot\boldsymbol{u}]+\frac{1}{2}\int_{0}^{T}\int_{\Gamma(t)}(\boldsymbol{u}\cdot\boldsymbol{n}-2\zeta\boldsymbol{e}_{y}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\boldsymbol{v}\\ +2\nu\int_{0}^{T}\int_{\Omega_{f}(t)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{v})+\int_{0}^{T}\int_{\Gamma(t)}\left(\frac{1}{2}|\boldsymbol{u}|^{2}-p\right)(\psi_{n}-v_{n})+\beta\int_{0}^{T}\int_{\Gamma(t)}(\zeta\boldsymbol{e}_{y}-\boldsymbol{u})\cdot\boldsymbol{\tau}(\boldsymbol{\psi}-\boldsymbol{v})\cdot\boldsymbol{\tau}\\ -\rho_{p}\int_{0}^{T}\int_{\hat{\Gamma}}\partial_{t}\hat{\omega}\cdot\partial_{t}\hat{\varphi}+\int_{0}^{T}\int_{\hat{\Gamma}}\hat{\Delta}\hat{\omega}\cdot\hat{\Delta}\hat{\varphi}-\rho_{b}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\partial_{t}\hat{\boldsymbol{\psi}}+2\mu_{e}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})\\ +\lambda_{e}\int_{0}^{T}\int_{\hat{\Omega}_{b}}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}})+2\mu_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\boldsymbol{D}}(\partial_{t}\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})+\lambda_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}(\hat{\nabla}\cdot\partial_{t}\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}})\\ -\alpha\int_{0}^{T}\int_{\Omega_{b}(t)}p\nabla\cdot\boldsymbol{\psi}-c_{0}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{p}\cdot\partial_{t}\hat{r}-\alpha\int_{0}^{T}\int_{\Omega_{b}(t)}\frac{D}{Dt}\boldsymbol{\eta}\cdot\nabla r-\alpha\int_{0}^{T}\int_{\Gamma(t)}(\zeta\boldsymbol{e}_{y}\cdot\boldsymbol{n})r\\ +\kappa\int_{0}^{T}\int_{\Omega_{b}(t)}\nabla p\cdot\nabla r-\int_{0}^{T}\int_{\Gamma(t)}((\boldsymbol{u}-\zeta\boldsymbol{e}_{y})\cdot\boldsymbol{n})r\\ =\int_{\Omega_{f}(0)}\boldsymbol{u}(0)\cdot\boldsymbol{v}(0)+\rho_{p}\int_{\hat{\Gamma}}\partial_{t}\hat{\omega}(0)\cdot\hat{\varphi}(0)+\rho_{b}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}(0)\cdot\hat{\boldsymbol{\psi}}(0)+c_{0}\int_{\hat{\Omega}_{b}}\hat{p}(0)\cdot\hat{r}(0). (23)
Remark 4.1.

It is immediate to see that a classical (temporally and spatially smooth) solution to the FPSI problem satisfies the weak formulation stated above. However, when considering less regular solutions (in particular, weak solutions in the class of finite-energy solutions), the above weak formulation is inadequate for the regularity of finite-energy solutions for the following reason. By the energy estimates (see Section 5.2), the regularity of the structure displacement 𝜼^\hat{\boldsymbol{\eta}} on Ω^b\hat{\Omega}_{b} is L(0,T,H1(Ω^b))L^{\infty}(0,T,H^{1}(\hat{\Omega}_{b})), which is not enough regularity to interpret the term

αΩb(t)p𝝍,\alpha\int_{\Omega_{b}(t)}p\nabla\cdot\boldsymbol{\psi},

since the test function has regularity 𝝍^H1(Ω^b)\hat{\boldsymbol{\psi}}\in H^{1}(\hat{\Omega}_{b}) on the fixed reference domain, due to the corresponding finite energy regularity of 𝜼^\hat{\boldsymbol{\eta}}. Hence, after changing variables, which adds an extra factor of det(𝑰+^𝜼^)\det(\boldsymbol{I}+\hat{\nabla}\hat{\boldsymbol{\eta}}) arising from the Jacobian, which is only in L(0,T;L1(Ω^b))L^{\infty}(0,T;L^{1}(\hat{\Omega}_{b})) in two dimensions, there is not enough regularity to guarantee that this integral is finite. Therefore, we cannot interpret the above notion of weak solution properly in the space of finite energy solutions, as the finite energy space does not have enough regularity to make sense of certain integrals in the weak formulation, involving the deformed domain Ωb(t)\Omega_{b}(t).

This is why we introduce a regularized problem, which is consistent with the original problem in the sense that weak solutions to the regularized problem converge, as the regularization parameter tends to zero, to a smooth solution of the original, nonregularized problem, when a smooth solution exists. This weak-classical consistency result will be shown in Sec. 10.

5 Regularized weak solution and statement of existence result

Since all the mathematical challenges related to the inability to properly interpret all of the terms in the weak solution arise fundamentally from the lack of regularity of 𝜼^\hat{\boldsymbol{\eta}} on Ω^b\hat{\Omega}_{b}, we will regularize 𝜼^\hat{\boldsymbol{\eta}} via a convolution with a smooth, compactly supported kernel, and introduce an appropriate regularized weak formulation of the original FPSI problem. Because we are working on a bounded domain Ω^b\hat{\Omega}_{b}, we must be careful to introduce the convolution in a way that preserves the Dirichlet condition on the left, top, and right boundaries of Ω^b=(0,L)×(0,R)\hat{\Omega}_{b}=(0,L)\times(0,R).

This is why we define an extended domain Ω~b\tilde{\Omega}_{b}:

Ω~b=[L,2L]×[R,2R],\tilde{\Omega}_{b}=[-L,2L]\times[-R,2R],

so that for δ<min(L,R)\delta<\min(L,R) the convolution of a function on Ω~b\tilde{\Omega}_{b} with a smooth function of compact support in the closed ball of radius δ\delta gives a function defined on Ω^b\hat{\Omega}_{b}. We then introduce an odd extension along the lines x^=0\hat{x}=0, x^=L\hat{x}=L, y^=0\hat{y}=0 and y^=R\hat{y}=R as follows.

Definition 5.1.

Given 𝜼^\hat{\boldsymbol{\eta}} defined on Ω^b\hat{\Omega}_{b} satisfying 𝜼^=0\hat{\boldsymbol{\eta}}=0 on x^=0\hat{x}=0, x^=L\hat{x}=L, and y^=R\hat{y}=R and 𝜼^=ω^𝒆y\hat{\boldsymbol{\eta}}=\hat{\omega}\boldsymbol{e}_{y} on y^=0\hat{y}=0, define the odd extension of η^\hat{\boldsymbol{\eta}} to Ω~b\tilde{\Omega}_{b} by keeping 𝜼^\hat{\boldsymbol{\eta}} the same on Ω^b=[0,L]×[0,R]\hat{\Omega}_{b}=[0,L]\times[0,R] and defining 𝜼^\hat{\boldsymbol{\eta}} outside of the closure of Ω^b\hat{\Omega}_{b} as follows:

  1. 1.

    On [0,L]×[R,0][0,L]\times[-R,0], set 𝜼^(x^,y^)=ω^(x^)𝒆y+(ω^(x^)𝒆y𝜼^(x^,y^))\hat{\boldsymbol{\eta}}(\hat{x},\hat{y})=\hat{\omega}(\hat{x})\boldsymbol{e}_{y}+(\hat{\omega}(\hat{x})\boldsymbol{e}_{y}-\hat{\boldsymbol{\eta}}(\hat{x},-\hat{y})).

  2. 2.

    On [0,L]×[R,2R][0,L]\times[R,2R], set 𝜼^(x^,y^)=𝜼^(x^,2Ry^)\hat{\boldsymbol{\eta}}(\hat{x},\hat{y})=-\hat{\boldsymbol{\eta}}(\hat{x},2R-\hat{y}).

  3. 3.

    On [L,0]×[R,2R][-L,0]\times[-R,2R], set 𝜼^(x^,y^)=𝜼^(x^,y^)\hat{\boldsymbol{\eta}}(\hat{x},\hat{y})=-\hat{\boldsymbol{\eta}}(-\hat{x},\hat{y}).

  4. 4.

    On [L,2L]×[R,2R][L,2L]\times[-R,2R], set 𝜼^(x^,y^)=𝜼^(2Lx^,y^)\hat{\boldsymbol{\eta}}(\hat{x},\hat{y})=-\hat{\boldsymbol{\eta}}(2L-\hat{x},\hat{y}).

Let σ\sigma be a radially symmetric function on 2\mathbb{R}^{2} with compact support in the closed ball of radius one such that 2σ=1\displaystyle\int_{\mathbb{R}^{2}}\sigma=1, and define

σδ=δ2σ(δ1𝒙), on 2.\sigma_{\delta}=\delta^{-2}\sigma(\delta^{-1}\boldsymbol{x}),\qquad\text{ on }\mathbb{R}^{2}.
Definition 5.2.

We define the following regularized functions which are spatially smooth on Ω^b\hat{\Omega}_{b}:

  • The regularized Biot displacement, which is obtained by extending 𝜼^\hat{\boldsymbol{\eta}} to Ω~b\tilde{\Omega}_{b} by odd extension and defining:

    𝜼^δ(t,x^,y^)=(𝜼^σδ)(t,x^,y^):=2d𝜼^(t,x^z,y^w)σδ(z,w)𝑑z𝑑w, on Ω^b,\hat{\boldsymbol{\eta}}^{\delta}(t,\hat{x},\hat{y})=(\hat{\boldsymbol{\eta}}*\sigma_{\delta})(t,\hat{x},\hat{y}):=\int_{{\mathbb{R}}^{2d}}\hat{\boldsymbol{\eta}}(t,\hat{x}-z,\hat{y}-w)\sigma_{\delta}(z,w)dzdw,\qquad\text{ on }\hat{\Omega}_{b}, (24)
  • The regularized Lagrangian mapping:

    𝚽^bηδ(t,)=Id+𝜼^δ(t,),\hat{\boldsymbol{\Phi}}_{b}^{\eta^{\delta}}(t,\cdot)=\text{Id}+\hat{\boldsymbol{\eta}}^{\delta}(t,\cdot), (25)
  • The regularized moving Biot domain:

    Ωbδ(t)=𝚽^bηδ(t,Ω^b).\Omega^{\delta}_{b}(t)=\hat{\boldsymbol{\Phi}}_{b}^{\eta^{\delta}}(t,\hat{\Omega}_{b}). (26)

    Note that even though the kinematic coupling condition holds for 𝜼^\hat{\boldsymbol{\eta}} in the sense that 𝜼^|Γ^=ω^𝒆y\hat{\boldsymbol{\eta}}|_{\hat{\Gamma}}=\hat{\omega}\boldsymbol{e}_{y}, it is not necessarily true that 𝜼^δ|Γ^=ω^𝒆y\hat{\boldsymbol{\eta}}^{\delta}|_{\hat{\Gamma}}=\hat{\omega}\boldsymbol{e}_{y}. Therefore, we will also define:

  • The regularized moving interface:

    Γδ(t)=𝚽^bηδ(t,Γ^).\Gamma^{\delta}(t)=\hat{\boldsymbol{\Phi}}^{\eta^{\delta}}_{b}(t,\hat{\Gamma}).

    Alternatively, Γ^δ\hat{\Gamma}^{\delta} is the plate interface if it were displaced from the reference configuration Γ^\hat{\Gamma} in the direction 𝜼^δ|Γ^\hat{\boldsymbol{\eta}}^{\delta}|_{\hat{\Gamma}}, which is a purely transverse yy displacement, as one can verify.

Note that by the way we extended 𝜼^\hat{\boldsymbol{\eta}} to the larger domain Ω~b\tilde{\Omega}_{b} we have that

𝜼^δ=0onΩ^bΓ^.\hat{\boldsymbol{\eta}}^{\delta}=0\quad{\rm on}\quad\partial\hat{\Omega}_{b}\setminus\hat{\Gamma}.

With these regularized versions of the Biot structure displacement and velocity, we can now define the notion of a weak solution to the regularized weak FPSI problem with the regularization parameter δ\delta. We start by defining the solution and test space, which are motivated by the energy estimates in Section 5.2, and then we state the regularized weak formulation in the moving domain framework and in the fixed reference domain framework.

Remark 5.1.

We have regularized the physical Biot domain using the regularized Biot displacement, which results in the regularized moving Biot domain Ωbδ(t)\Omega_{b}^{\delta}(t) as stated in (26). We emphasize that the main reason for this regularization is because the structure displacement 𝜼\boldsymbol{\eta}, which is in the finite energy space H1(Ωb)H^{1}(\Omega_{b}) (without regularization), does not posses sufficient regularity to make sense of the definition of Ωbδ(t)\Omega_{b}^{\delta}(t). However, while the moving Biot domain requires regularization, we emphasize that there is no need to regularize the fluid domain Ωf(t)\Omega_{f}(t) because the fluid domain Ωf(t)\Omega_{f}(t) can be defined without explicit reference to the Biot displacement 𝜼\boldsymbol{\eta}, and hence it is not affected by the regularity issues associated with 𝜼\boldsymbol{\eta}. In particular, Ωf(t)\Omega_{f}(t) can already be well-defined by using just the plate displacement ω\omega. Even though ω\omega is the trace of 𝜼\boldsymbol{\eta} along Γ\Gamma, the plate displacement ω\omega has additional regularity in the finite energy space, i.e., ωH02(Γ)\omega\in H_{0}^{2}(\Gamma) due to the fact that ω\omega itself satisfies the plate equation. This makes ω\omega a continuous function on Γ\Gamma which allows us to define Ωf(t)\Omega_{f}(t) unambiguously.

5.1 Functional spaces and definition of weak solutions

Definition 5.3.

(Solution and test spaces for the regularized problem)

  • Fluid function space (moving domain/Eulerian formulation).

    Vf(t)={𝒖=(ux,uy)H1(Ωf(t)):𝒖=0, and 𝒖=0 when x=0,x=L,y=R},V_{f}(t)=\{\boldsymbol{u}=(u_{x},u_{y})\in H^{1}(\Omega_{f}(t)):\nabla\cdot\boldsymbol{u}=0,\text{ and }\boldsymbol{u}=0\text{ when }x=0,x=L,y=-R\}, (27)
    𝒱f=L(0,T;L2(Ωf(t)))L2(0,T;Vf(t)).\mathcal{V}_{f}=L^{\infty}(0,T;L^{2}(\Omega_{f}(t)))\cap L^{2}(0,T;V_{f}(t)). (28)
  • Fluid function space (fixed domain/Lagrangian formulation).

    Vfω={𝒖^=(u^x,u^y)H1(Ω^f):^fω𝒖^=0, and 𝒖^=0 when x^=0,x^=L,y^=R},V^{\omega}_{f}=\{\hat{\boldsymbol{u}}=(\hat{u}_{x},\hat{u}_{y})\in H^{1}(\hat{\Omega}_{f}):\hat{\nabla}^{\omega}_{f}\cdot\hat{\boldsymbol{u}}=0,\text{ and }\hat{\boldsymbol{u}}=0\text{ when }\hat{x}=0,\hat{x}=L,\hat{y}=-R\}, (29)
    𝒱fω=L(0,T;L2(Ω^f))L2(0,T;Vfω).\mathcal{V}^{\omega}_{f}=L^{\infty}(0,T;L^{2}(\hat{\Omega}_{f}))\cap L^{2}(0,T;V^{\omega}_{f}). (30)
  • Plate function space.

    𝒱ω=W1,(0,T;L2(Γ^))L(0,T;H02(Γ^)).\mathcal{V}_{\omega}=W^{1,\infty}(0,T;L^{2}(\hat{\Gamma}))\cap L^{\infty}(0,T;H_{0}^{2}(\hat{\Gamma})). (31)
  • Biot displacement function space.

    Vd={𝜼^=(η^x,η^y)H1(Ω^b):𝜼^=0 for x^=0,x^=L,y^=R, and η^x=0 on Γ^},V_{d}=\{\hat{\boldsymbol{\eta}}=(\hat{\eta}_{x},\hat{\eta}_{y})\in H^{1}(\hat{\Omega}_{b}):\hat{\boldsymbol{\eta}}=0\text{ for }\hat{x}=0,\hat{x}=L,\hat{y}=R,\text{ and }\hat{\eta}_{x}=0\text{ on }\hat{\Gamma}\}, (32)
    𝒱b=W1,(0,T;L2(Ω^b))L(0,T;Vd)H1(0,T;Vd).\mathcal{V}_{b}=W^{1,\infty}(0,T;L^{2}(\hat{\Omega}_{b}))\cap L^{\infty}(0,T;V_{d})\cap H^{1}(0,T;V_{d}). (33)
  • Biot pore pressure function space.

    Vp={p^H1(Ω^b):p^=0 for x^=0,x^=L,y^=R},V_{p}=\{\hat{p}\in H^{1}(\hat{\Omega}_{b}):\hat{p}=0\text{ for }\hat{x}=0,\hat{x}=L,\hat{y}=R\}, (34)
    𝒬b=L(0,T;L2(Ω^b))L2(0,T;Vp).\mathcal{Q}_{b}=L^{\infty}(0,T;L^{2}(\hat{\Omega}_{b}))\cap L^{2}(0,T;V_{p}). (35)
  • Weak solution space (moving domain).

    𝒱sol={(𝒖,ω^,𝜼^,p^)𝒱f×𝒱ω×𝒱b×𝒬b:𝜼^=ω^𝒆y on Γ^}.\mathcal{V}_{\text{sol}}=\{(\boldsymbol{u},\hat{\omega},\hat{\boldsymbol{\eta}},\hat{p})\in\mathcal{V}_{f}\times\mathcal{V}_{\omega}\times\mathcal{V}_{b}\times\mathcal{Q}_{b}:\hat{\boldsymbol{\eta}}=\hat{\omega}\boldsymbol{e}_{y}\text{ on }\hat{\Gamma}\}. (36)
  • Weak solution space (fixed domain).

    𝒱solω={(𝒖^,ω^,𝜼^,p^)𝒱fω×𝒱ω×𝒱b×𝒬b:𝜼^=ω^𝒆y on Γ^}.\mathcal{V}^{\omega}_{\text{sol}}=\{(\hat{\boldsymbol{u}},\hat{\omega},\hat{\boldsymbol{\eta}},\hat{p})\in\mathcal{V}^{\omega}_{f}\times\mathcal{V}_{\omega}\times\mathcal{V}_{b}\times\mathcal{Q}_{b}:\hat{\boldsymbol{\eta}}=\hat{\omega}\boldsymbol{e}_{y}\text{ on }\hat{\Gamma}\}. (37)
  • Test space (moving domain).

    𝒱test={(𝒗,φ^,𝝍^,r^)Cc1([0,T);Vf(t)×H02(Γ^)×Vd×Vp):𝝍^=φ^𝒆y on Γ^}.\mathcal{V}_{\text{test}}=\{(\boldsymbol{v},\hat{\varphi},\hat{\boldsymbol{\psi}},\hat{r})\in C_{c}^{1}([0,T);V_{f}(t)\times H_{0}^{2}(\hat{\Gamma})\times V_{d}\times V_{p}):\hat{\boldsymbol{\psi}}=\hat{\varphi}\boldsymbol{e}_{y}\text{ on }\hat{\Gamma}\}. (38)
  • Test space (fixed domain).

    𝒱testω={(𝒗^,φ^,𝝍^,r^)Cc1([0,T);Vfω×H02(Γ^)×Vd×Vp):𝝍^=φ^𝒆y on Γ^}.\mathcal{V}^{\omega}_{\text{test}}=\{(\hat{\boldsymbol{v}},\hat{\varphi},\hat{\boldsymbol{\psi}},\hat{r})\in C_{c}^{1}([0,T);V^{\omega}_{f}\times H_{0}^{2}(\hat{\Gamma})\times V_{d}\times V_{p}):\hat{\boldsymbol{\psi}}=\hat{\varphi}\boldsymbol{e}_{y}\text{ on }\hat{\Gamma}\}. (39)
Remark 5.2.

Because Γ^\hat{\Gamma} is one dimensional, for plate displacements ω^𝒱ω\hat{\omega}\in\mathcal{V}_{\omega}, we have that ω^C(0,T;C1(Γ^))\hat{\omega}\in C(0,T;C^{1}(\hat{\Gamma})) and hence, there is a one-to-one correspondence between functions in 𝒱sol\mathcal{V}_{\text{sol}} and 𝒱solω\mathcal{V}^{\omega}_{\text{sol}} and functions in 𝒱test\mathcal{V}_{\text{test}} and 𝒱testω\mathcal{V}^{\omega}_{\text{test}}, given by composition with the ALE mapping (16).

Next, we state the weak formulation to the regularized problem as follows.

Definition 5.4.

(Weak solution to the regularized problem, moving fluid domain formulation) An ordered four-tuple (𝒖,ω^,𝜼^,p)𝒱sol(\boldsymbol{u},\hat{\omega},\hat{\boldsymbol{\eta}},p)\in\mathcal{V}_{\text{sol}} is a weak solution to the regularized nonlinearly coupled FPSI problem with regularization parameter δ\delta if for every test function (𝒗,φ^,𝝍^,r^)𝒱test(\boldsymbol{v},\hat{\varphi},\hat{\boldsymbol{\psi}},\hat{r})\in\mathcal{V}_{\text{test}},

0TΩf(t)𝒖t𝒗+120TΩf(t)[((𝒖)𝒖)𝒗((𝒖)𝒗)𝒖]+120TΓ(t)(𝒖𝒏2ζ𝒆y𝒏)𝒖𝒗+2ν0TΩf(t)𝑫(𝒖):𝑫(𝒗)+0TΓ(t)(12|𝒖|2p)(ψnvn)+β0TΓ(t)(ζ𝒆y𝒖)𝝉(𝝍𝒗)𝝉ρp0TΓ^tω^tφ^+0TΓ^Δ^ω^Δ^φ^ρb0TΩ^bt𝜼^t𝝍^+2μe0TΩ^b𝑫^(𝜼^):𝑫^(𝝍^)+λe0TΩ^b(^𝜼^)(^𝝍^)+2μv0TΩ^b𝑫^(t𝜼^):𝑫^(𝝍^)+λv0TΩ^b(^t𝜼^)(^𝝍^)α0TΩbδ(t)p𝝍c00TΩ^bp^tr^α0TΩbδ(t)DδDt𝜼rα0TΓδ(t)(ζ𝒆y𝒏δ)r+κ0TΩbδ(t)pr0TΓ(t)((𝒖ζ𝒆y)𝒏)r=Ωf(0)𝒖(0)𝒗(0)+ρpΓ^tω^(0)φ^(0)+ρbΩ^bt𝜼^(0)𝝍^(0)+c0Ω^bp^(0)r^(0),-\int_{0}^{T}\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\partial_{t}\boldsymbol{v}+\frac{1}{2}\int_{0}^{T}\int_{\Omega_{f}(t)}[((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot\boldsymbol{v}-((\boldsymbol{u}\cdot\nabla)\boldsymbol{v})\cdot\boldsymbol{u}]+\frac{1}{2}\int_{0}^{T}\int_{\Gamma(t)}(\boldsymbol{u}\cdot\boldsymbol{n}-2\zeta\boldsymbol{e}_{y}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\boldsymbol{v}\\ +2\nu\int_{0}^{T}\int_{\Omega_{f}(t)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{v})+\int_{0}^{T}\int_{\Gamma(t)}\left(\frac{1}{2}|\boldsymbol{u}|^{2}-p\right)(\psi_{n}-v_{n})+\beta\int_{0}^{T}\int_{\Gamma(t)}(\zeta\boldsymbol{e}_{y}-\boldsymbol{u})\cdot\boldsymbol{\tau}(\boldsymbol{\psi}-\boldsymbol{v})\cdot\boldsymbol{\tau}\\ -\rho_{p}\int_{0}^{T}\int_{\hat{\Gamma}}\partial_{t}\hat{\omega}\cdot\partial_{t}\hat{\varphi}+\int_{0}^{T}\int_{\hat{\Gamma}}\hat{\Delta}\hat{\omega}\cdot\hat{\Delta}\hat{\varphi}-\rho_{b}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\partial_{t}\hat{\boldsymbol{\psi}}+2\mu_{e}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})\\ +\lambda_{e}\int_{0}^{T}\int_{\hat{\Omega}_{b}}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}})+2\mu_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\boldsymbol{D}}(\partial_{t}\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})+\lambda_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}(\hat{\nabla}\cdot\partial_{t}\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}})\\ -\alpha\int_{0}^{T}\int_{\Omega_{b}^{\delta}(t)}p\nabla\cdot\boldsymbol{\psi}-c_{0}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{p}\cdot\partial_{t}\hat{r}-\alpha\int_{0}^{T}\int_{\Omega_{b}^{\delta}(t)}\frac{D^{\delta}}{Dt}\boldsymbol{\eta}\cdot\nabla r-\alpha\int_{0}^{T}\int_{\Gamma^{\delta}(t)}(\zeta\boldsymbol{e}_{y}\cdot\boldsymbol{n}^{\delta})r\\ +\kappa\int_{0}^{T}\int_{\Omega^{\delta}_{b}(t)}\nabla p\cdot\nabla r-\int_{0}^{T}\int_{\Gamma(t)}((\boldsymbol{u}-\zeta\boldsymbol{e}_{y})\cdot\boldsymbol{n})r\\ =\int_{\Omega_{f}(0)}\boldsymbol{u}(0)\cdot\boldsymbol{v}(0)+\rho_{p}\int_{\hat{\Gamma}}\partial_{t}\hat{\omega}(0)\cdot\hat{\varphi}(0)+\rho_{b}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}(0)\cdot\hat{\boldsymbol{\psi}}(0)+c_{0}\int_{\hat{\Omega}_{b}}\hat{p}(0)\cdot\hat{r}(0), (40)

where DδDt=ddt+(𝝃δ)\frac{D^{\delta}}{Dt}=\frac{d}{dt}+(\boldsymbol{\xi}^{\delta}\cdot\nabla) with 𝝃δ(t,)=t𝜼^δ(t,(𝚽bηδ)1(t,))\boldsymbol{\xi}^{\delta}(t,\cdot)=\partial_{t}\hat{\boldsymbol{\eta}}^{\delta}(t,(\boldsymbol{\Phi}^{\eta^{\delta}}_{b})^{-1}(t,\cdot)) is the material derivative with respect to the regularized displacement, 𝒏\boldsymbol{n} denotes the upward pointing normal vector to Γ(t)\Gamma(t), and 𝒏δ\boldsymbol{n}^{\delta} denotes the upward pointing normal vector to Γδ(t)\Gamma^{\delta}(t).

Notice that only four terms contain regularization via convolution with parameter δ\delta. While there are many different ways to write the regularized weak formulation, the regularization presented above is a regularization that deviates from the original, nonregularized problem, in the smallest possible number of terms, and is still consistent with the original, nonregularized problem, as we show later.

Remark 5.3.

While the solution to the regularized problem above depends on the regularization parameter δ\delta implicitly, to simplify notation we will drop the δ\delta notation whenever it is clear from the context that we are working with the solution to the regularized problem.

Remark 5.4.

We simplify notation by omitting the explicit compositions with the maps 𝚽^fω\hat{\boldsymbol{\Phi}}^{\omega}_{f}, 𝚽^Γω\hat{\boldsymbol{\Phi}}^{\omega}_{\Gamma}, 𝚽^bη\hat{\boldsymbol{\Phi}}^{\eta}_{b}, and 𝚽^bηδ\hat{\boldsymbol{\Phi}}^{\eta^{\delta}}_{b}, and their inverses. The necessary compositions with such mappings will be clear from the context. For example,

α0TΩbδ(t)p𝝍meansα0TΩbδ(t)(p^(𝚽bηδ)1)(𝝍^(𝚽bηδ)1),-\alpha\int_{0}^{T}\int_{\Omega_{b}^{\delta}(t)}p\nabla\cdot\boldsymbol{\psi}\quad{\rm means}\quad-\alpha\int_{0}^{T}\int_{\Omega_{b}^{\delta}(t)}\left(\hat{p}\circ(\boldsymbol{\Phi}^{\eta^{\delta}}_{b})^{-1}\right)\nabla\cdot\left(\hat{\boldsymbol{\psi}}\circ(\boldsymbol{\Phi}^{\eta^{\delta}}_{b})^{-1}\right),

and

0TΓ(t)((𝒖ζ𝒆y)𝒏)rmeans0TΓ(t)((𝒖(ζ(𝚽Γω)1)𝒆y)𝒏)(r^(𝚽bη)1).-\int_{0}^{T}\int_{\Gamma(t)}((\boldsymbol{u}-\zeta\boldsymbol{e}_{y})\cdot\boldsymbol{n})r\quad{\rm means}\quad-\int_{0}^{T}\int_{\Gamma(t)}\left(\left(\boldsymbol{u}-(\zeta\circ(\boldsymbol{\Phi}^{\omega}_{\Gamma})^{-1})\boldsymbol{e}_{y}\right)\cdot\boldsymbol{n}\right)\left(\hat{r}\circ(\boldsymbol{\Phi}^{\eta}_{b})^{-1}\right).

Notice that here we tacitly assume that 𝚽bηδ\boldsymbol{\Phi}^{\eta^{\delta}}_{b} is invertible. We will later justify this assumption by proving that it holds on some time interval [0,Tδ][0,T_{\delta}], where the time TδT_{\delta} may depend on the regularization parameter δ\delta. Next, we reformulate the definition of a regularized weak solution on the fixed reference domain. Recall that the Jacobians 𝒥^fω\hat{\mathcal{J}}^{\omega}_{f}, 𝒥^bη\hat{\mathcal{J}}^{\eta}_{b}, and 𝒥^Γω\hat{\mathcal{J}}^{\omega}_{\Gamma} in (18) will appear upon using a change of variables to map the problem onto the reference domain. To transform the first term in the weak formulation (40) above, we use (20) to transform the time derivatives and assume that |ω^|<R|\hat{\omega}|<R so that there is no domain degeneracy. After using (19) and (20), we get

Ωf(t)𝒖t𝒗=Ω^f(1+ω^R)𝒖^t𝒗^Ω^f(1+ω^R)𝒖^[(𝒘^^fω)𝒗^]\displaystyle\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\partial_{t}\boldsymbol{v}=\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)\hat{\boldsymbol{u}}\cdot\partial_{t}\hat{\boldsymbol{v}}-\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)\hat{\boldsymbol{u}}\cdot[(\hat{\boldsymbol{w}}\cdot\hat{\nabla}^{\omega}_{f})\hat{\boldsymbol{v}}]
=Ω^f(1+ω^R)𝒖^t𝒗^1RΩ^f𝒖^[(R+y^)tω^y^𝒗^]\displaystyle=\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)\hat{\boldsymbol{u}}\cdot\partial_{t}\hat{\boldsymbol{v}}-\frac{1}{R}\int_{\hat{\Omega}_{f}}\hat{\boldsymbol{u}}\cdot[(R+\hat{y})\partial_{t}\hat{\omega}\partial_{\hat{y}}\hat{\boldsymbol{v}}]
=Ω^f(1+ω^R)𝒖^t𝒗^12RΩ^f𝒖^[(R+y^)tω^y^𝒗^]+12RΩ^f(tω^)𝒖^𝒗^\displaystyle=\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)\hat{\boldsymbol{u}}\cdot\partial_{t}\hat{\boldsymbol{v}}-\frac{1}{2R}\int_{\hat{\Omega}_{f}}\hat{\boldsymbol{u}}\cdot[(R+\hat{y})\partial_{t}\hat{\omega}\partial_{\hat{y}}\hat{\boldsymbol{v}}]+\frac{1}{2R}\int_{\hat{\Omega}_{f}}(\partial_{t}\hat{\omega})\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{v}}
+12RΩ^f[(R+y^)tω^y^𝒖^]𝒗^12Γ^(𝒖^𝒗^)tω^\displaystyle+\frac{1}{2R}\int_{\hat{\Omega}_{f}}[(R+\hat{y})\partial_{t}\hat{\omega}\partial_{\hat{y}}\hat{\boldsymbol{u}}]\cdot\hat{\boldsymbol{v}}-\frac{1}{2}\int_{\hat{\Gamma}}(\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{v}})\partial_{t}\hat{\omega}
=Ω^f(1+ω^R)𝒖^t𝒗^12Ω^f(1+ω^R)[((𝒘^^fω)𝒗^)𝒖^((𝒘^^fω)𝒖^)𝒗^]\displaystyle=\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)\hat{\boldsymbol{u}}\cdot\partial_{t}\hat{\boldsymbol{v}}-\frac{1}{2}\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)[((\hat{\boldsymbol{w}}\cdot\hat{\nabla}^{\omega}_{f})\hat{\boldsymbol{v}})\cdot\hat{\boldsymbol{u}}-((\hat{\boldsymbol{w}}\cdot\hat{\nabla}^{\omega}_{f})\hat{\boldsymbol{u}})\cdot\hat{\boldsymbol{v}}]
+12RΩ^f(tω^)𝒖^𝒗^12Γ^(𝒖^𝒗^)tω^,\displaystyle+\frac{1}{2R}\int_{\hat{\Omega}_{f}}(\partial_{t}\hat{\omega})\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{v}}-\frac{1}{2}\int_{\hat{\Gamma}}(\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{v}})\partial_{t}\hat{\omega}, (41)

where we integrated by parts in the y^\hat{y} direction. Note that the final term in (5.1) will combine with the following term in (40):

0TΓ(t)(ζ𝒆y𝒏)𝒖𝒗=0TΓ^(𝒖^𝒗^)tω^,\int_{0}^{T}\int_{\Gamma(t)}(\zeta\boldsymbol{e}_{y}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\boldsymbol{v}=\int_{0}^{T}\int_{\hat{\Gamma}}(\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{v}})\partial_{t}\hat{\omega}, (42)

where we used 𝒏=(x^ω^,1)/𝒥^Γω\displaystyle\boldsymbol{n}=(-\partial_{\hat{x}}\hat{\omega},1)/{\hat{\mathcal{J}}^{\omega}_{\Gamma}} for the normal vector to the interface and ζ𝒆y|Γ(t)=tω^𝒆y\zeta\boldsymbol{e}_{y}|_{\Gamma(t)}=\partial_{t}\hat{\omega}\boldsymbol{e}_{y}. Because the transformation from Γ(t)\Gamma(t) to Γ^\hat{\Gamma} cancels out the factor of 𝒥^Γω\hat{\mathcal{J}}^{\omega}_{\Gamma} in the unit normal vector, it is useful to define the following renormalized normal and tangent vectors:

𝒏^ω=(x^ω^,1),𝝉^ω=(1,x^ω^).\hat{\boldsymbol{n}}^{\omega}=(-\partial_{\hat{x}}\hat{\omega},1),\qquad\hat{\boldsymbol{\tau}}^{\omega}=(1,\partial_{\hat{x}}\hat{\omega}). (43)

We similarly define

𝒏^ωδ=(x^(𝜼^δ|Γ^),1).\hat{\boldsymbol{n}}^{\omega^{\delta}}=(-\partial_{\hat{x}}(\hat{\boldsymbol{\eta}}^{\delta}|_{\hat{\Gamma}}),1). (44)

We are now ready to state the definition of a weak solution to the regularized problem on the fixed reference domain.

Definition 5.5.

(Weak solution to the regularized problem, fixed fluid domain formulation) An ordered four-tuple (𝒖^,ω^,𝜼^,p^)𝒱solω(\hat{\boldsymbol{u}},\hat{\omega},\hat{\boldsymbol{\eta}},\hat{p})\in\mathcal{V}^{\omega}_{\text{sol}} is a weak solution to the regularized nonlinearly coupled FPSI problem with regularization parameter δ\delta if for all test functions (𝒗^,φ^,𝝍^,r^)𝒱testω(\hat{\boldsymbol{v}},\hat{\varphi},\hat{\boldsymbol{\psi}},\hat{r})\in\mathcal{V}^{\omega}_{\text{test}}, the following equality holds:

0TΩ^f(1+ω^R)𝒖^t𝒗^+120TΩ^f(1+ω^R)[((𝒖^𝒘^)^fω𝒖^)𝒗^((𝒖^𝒘^)^fω𝒗^)𝒖^]12R0TΩ^f(tω^)𝒖^𝒗^+120TΓ^(𝒖^𝒏^ωζ^𝒆y𝒏^ω)𝒖^𝒗^+2ν0TΩ^f(1+ω^R)𝑫^(𝒖^):𝑫^(𝒗^)+0TΓ^(12|𝒖^|2p^)(𝝍^𝒗^)𝒏^ω+β𝒥^Γω0TΓ^(ζ^𝒆y𝒖^)𝝉^ω(𝝍^𝒗^)𝝉^ωρp0TΓ^tω^tφ^+0TΓ^Δ^ω^Δ^φ^ρb0TΩ^bt𝜼^t𝝍^+2μe0TΩ^b𝑫^(𝜼^):𝑫^(𝝍^)+λe0TΩ^b(^𝜼^)(^𝝍^)+2μv0TΩ^b𝑫^(t𝜼^):𝑫^(𝝍^)+λv0TΩ^b(^t𝜼^)(^𝝍^)α0TΩ^b𝒥^bηδp^^bηδ𝝍^c00TΩ^bp^tr^α0TΩ^b𝒥^bηδt𝜼^^bηδr^α0TΓ^(ζ^𝒆y𝒏^ωδ)r^+κ0TΩ^b𝒥^bηδ^bηδp^^bηδr^0TΓ^((𝒖^ζ^𝒆y)𝒏^ω)r^=Ωf(0)𝒖(0)𝒗(0)+ρpΓ^tω^(0)φ^(0)+ρbΩ^bt𝜼^(0)𝝍^(0)+c0Ω^bp^(0)r^(0),-\int_{0}^{T}\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)\hat{\boldsymbol{u}}\cdot\partial_{t}\hat{\boldsymbol{v}}+\frac{1}{2}\int_{0}^{T}\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)[((\hat{\boldsymbol{u}}-\hat{\boldsymbol{w}})\cdot\hat{\nabla}^{\omega}_{f}\hat{\boldsymbol{u}})\cdot\hat{\boldsymbol{v}}-((\hat{\boldsymbol{u}}-\hat{\boldsymbol{w}})\cdot\hat{\nabla}^{\omega}_{f}\hat{\boldsymbol{v}})\cdot\hat{\boldsymbol{u}}]\\ -\frac{1}{2R}\int_{0}^{T}\int_{\hat{\Omega}_{f}}(\partial_{t}\hat{\omega})\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{v}}+\frac{1}{2}\int_{0}^{T}\int_{\hat{\Gamma}}(\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{n}}^{\omega}-\hat{\zeta}\boldsymbol{e}_{y}\cdot\hat{\boldsymbol{n}}^{\omega})\hat{\boldsymbol{u}}\cdot\hat{\boldsymbol{v}}+2\nu\int_{0}^{T}\int_{\hat{\Omega}_{f}}\left(1+\frac{\hat{\omega}}{R}\right)\hat{\boldsymbol{D}}(\hat{\boldsymbol{u}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{v}})\\ +\int_{0}^{T}\int_{\hat{\Gamma}}\left(\frac{1}{2}|\hat{\boldsymbol{u}}|^{2}-\hat{p}\right)(\hat{\boldsymbol{\psi}}-\hat{\boldsymbol{v}})\cdot\hat{\boldsymbol{n}}^{\omega}+\frac{\beta}{\hat{\mathcal{J}}^{\omega}_{\Gamma}}\int_{0}^{T}\int_{\hat{\Gamma}}(\hat{\zeta}\boldsymbol{e}_{y}-\hat{\boldsymbol{u}})\cdot\hat{\boldsymbol{\tau}}^{\omega}(\hat{\boldsymbol{\psi}}-\hat{\boldsymbol{v}})\cdot\hat{\boldsymbol{\tau}}^{\omega}\\ -\rho_{p}\int_{0}^{T}\int_{\hat{\Gamma}}\partial_{t}\hat{\omega}\cdot\partial_{t}\hat{\varphi}+\int_{0}^{T}\int_{\hat{\Gamma}}\hat{\Delta}\hat{\omega}\cdot\hat{\Delta}\hat{\varphi}-\rho_{b}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\partial_{t}\hat{\boldsymbol{\psi}}+2\mu_{e}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})\\ +\lambda_{e}\int_{0}^{T}\int_{\hat{\Omega}_{b}}(\hat{\nabla}\cdot\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}})+2\mu_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\boldsymbol{D}}(\partial_{t}\hat{\boldsymbol{\eta}}):\hat{\boldsymbol{D}}(\hat{\boldsymbol{\psi}})+\lambda_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}(\hat{\nabla}\cdot\partial_{t}\hat{\boldsymbol{\eta}})(\hat{\nabla}\cdot\hat{\boldsymbol{\psi}})\\ -\alpha\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\mathcal{J}}^{\eta^{\delta}}_{b}\hat{p}\hat{\nabla}^{\eta^{\delta}}_{b}\cdot\hat{\boldsymbol{\psi}}-c_{0}\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{p}\cdot\partial_{t}\hat{r}-\alpha\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\mathcal{J}}^{\eta^{\delta}}_{b}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\hat{\nabla}^{\eta^{\delta}}_{b}\hat{r}\\ -\alpha\int_{0}^{T}\int_{\hat{\Gamma}}(\hat{\zeta}\boldsymbol{e}_{y}\cdot\hat{\boldsymbol{n}}^{\omega^{\delta}})\hat{r}+\kappa\int_{0}^{T}\int_{\hat{\Omega}_{b}}\hat{\mathcal{J}}^{\eta^{\delta}}_{b}\hat{\nabla}^{\eta^{\delta}}_{b}\hat{p}\cdot\hat{\nabla}^{\eta^{\delta}}_{b}\hat{r}-\int_{0}^{T}\int_{\hat{\Gamma}}((\hat{\boldsymbol{u}}-\hat{\zeta}\boldsymbol{e}_{y})\cdot\hat{\boldsymbol{n}}^{\omega})\hat{r}\\ =\int_{\Omega_{f}(0)}\boldsymbol{u}(0)\cdot\boldsymbol{v}(0)+\rho_{p}\int_{\hat{\Gamma}}\partial_{t}\hat{\omega}(0)\cdot\hat{\varphi}(0)+\rho_{b}\int_{\hat{\Omega}_{b}}\partial_{t}\hat{\boldsymbol{\eta}}(0)\cdot\hat{\boldsymbol{\psi}}(0)+c_{0}\int_{\hat{\Omega}_{b}}\hat{p}(0)\cdot\hat{r}(0), (45)

where 𝒥^bηδ\hat{\mathcal{J}}^{\eta^{\delta}}_{b} and 𝒥^Γω\hat{\mathcal{J}}^{\omega}_{\Gamma} are defined in (18), 𝒘^\hat{\boldsymbol{w}} is defined in (20), ^fω\hat{\nabla}^{\omega}_{f} in (19), ^bηδg^\hat{\nabla}^{\eta^{\delta}}_{b}\hat{g} in (21), and ζ^\hat{\zeta} in (51).

5.2 Formal energy inequality

Here we show that the regularized problem is defined in a way that still gives rise to an energy equality, which in fact is the same energy inequality that one would formally obtain for the original problem, except with integrals over the moving Biot domain Ωb(t)\Omega_{b}(t) becoming integrals over the regularized moving Biot domain Ωbδ(t)\Omega^{\delta}_{b}(t). More precisely, we formally prove that a weak solution to the regularized problem satisfies the following energy equality.

Lemma 5.1.

Assuming that a weak solution exists, the following energy equality holds:

EK(T)+EE(T)+0T(DfV(t)+DbV(t)+DfbV(t)+DβV(t))𝑑t=EK(0)+EE(0)E^{K}(T)+{{E}}^{E}(T)+\int_{0}^{T}\left({{D}}^{V}_{f}(t)+{{D}}^{V}_{b}(t)+D^{V}_{f_{b}}(t)+D^{V}_{\beta}(t)\right)dt=E^{K}(0)+{{E}}^{E}(0) (46)

where

EK(t)=12Ωf(t)|𝒖(t)|2+12ρbΩ^b|t𝜼^(t)|2+12c0Ω^b|p^(t)|2+12ρpΓ^|tω^(t)|2\displaystyle{{E}}^{K}(t)=\frac{1}{2}\int_{\Omega_{f}(t)}|\boldsymbol{u}(t)|^{2}+\frac{1}{2}\rho_{b}\int_{\hat{\Omega}_{b}}|\partial_{t}\hat{\boldsymbol{\eta}}(t)|^{2}+\frac{1}{2}c_{0}\int_{\hat{\Omega}_{b}}|\hat{p}(t)|^{2}+\frac{1}{2}\rho_{p}\int_{\hat{\Gamma}}|\partial_{t}\hat{\omega}(t)|^{2}

is the sum of the kinetic energy of the fluid, the kinetic energy of the Biot poroviscoelastic matrix motion, the kinetic energy of the filtrating fluid flow in the Biot medium, and the kinetic energy of the plate motion, EE(t){{E}}^{E}(t) is defined by

EE(t)=2μeΩ^b|𝑫^(𝜼^)(t)|2+2λeΩ^b|^𝜼^(t)|2+Γ^|Δ^ω^(t)|2,\displaystyle{{E}}^{E}(t)=2\mu_{e}\int_{\hat{\Omega}_{b}}|\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}})(t)|^{2}+2\lambda_{e}\int_{\hat{\Omega}_{b}}|\hat{\nabla}\cdot\hat{\boldsymbol{\eta}}(t)|^{2}+\int_{\hat{\Gamma}}|\hat{\Delta}\hat{\omega}(t)|^{2},

which corresponds to the elastic energy of the Biot poroviscoelastic matrix and the elastic energy of the plate, and

DfV(t)=2νΩf(t)|𝑫(𝒖)|2,DbV(t)=2μvΩ^b|𝑫^(t𝜼^)|2+λvΩ^b|^t𝜼^|2,\displaystyle{{D}}^{V}_{f}(t)=2\nu\int_{\Omega_{f}(t)}|\boldsymbol{D}(\boldsymbol{u})|^{2},\quad{{D}}^{V}_{b}(t)=2\mu_{v}\int_{\hat{\Omega}_{b}}|\hat{\boldsymbol{D}}(\partial_{t}\hat{\boldsymbol{\eta}})|^{2}+\lambda_{v}\int_{\hat{\Omega}_{b}}|\hat{\nabla}\cdot\partial_{t}\hat{\boldsymbol{\eta}}|^{2},
DfbV(t)=κΩbδ(t)|p|2,DβV(t)=βΓ(t)|(𝝃𝒖)𝝉|2\displaystyle D^{V}_{f_{b}}(t)=\kappa\int_{\Omega_{b}^{\delta}(t)}|\nabla p|^{2},\quad D^{V}_{\beta}(t)=\beta\int_{\Gamma(t)}|(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}|^{2}

correspond to dissipation due to fluid viscosity, viscosity of the Biot poroviscoelastic matrix, dissipation due to permeability effects, and dissipation due to friction in the Beavers-Joseph-Saffman slip condition.

Proof.

To derive this energy equality we start by substituting (𝒗^,φ^,𝝍^,r^)=(𝒖^,ζ^,t𝜼^,p^)(\hat{\boldsymbol{v}},\hat{\varphi},\hat{\boldsymbol{\psi}},\hat{r})=(\hat{\boldsymbol{u}},\hat{\zeta},\partial_{t}\hat{\boldsymbol{\eta}},\hat{p}) into the regularized weak formulation (45) defined on the fixed reference domain and calculate

12Γ^(𝒖^ζ^𝒆y)𝒏^ω|𝒖^|2+Γ^(12|𝒖^|2p^)(ζ^𝒆y𝒖^)𝒏^ωΓ^((𝒖^ζ^𝒆y)𝒏^ω)p^=0.\frac{1}{2}\int_{\hat{\Gamma}}(\hat{\boldsymbol{u}}-\hat{\zeta}\boldsymbol{e}_{y})\cdot\hat{\boldsymbol{n}}^{\omega}|\hat{\boldsymbol{u}}|^{2}+\int_{\hat{\Gamma}}\left(\frac{1}{2}|\hat{\boldsymbol{u}}|^{2}-\hat{p}\right)(\hat{\zeta}\boldsymbol{e}_{y}-\hat{\boldsymbol{u}})\cdot\hat{\boldsymbol{n}}^{\omega}-\int_{\hat{\Gamma}}((\hat{\boldsymbol{u}}-\hat{\zeta}\boldsymbol{e}_{y})\cdot\hat{\boldsymbol{n}}^{\omega})\hat{p}=0.

Furthermore, using integration by parts one obtains

α(Ω^b𝒥^bηδp^^bηδt𝜼^+Ω^b𝒥^bηδt𝜼^^bηδp^+Γ^(ζ^𝒆y𝒏^ωδ)p^)\displaystyle\alpha\left(\int_{\hat{\Omega}_{b}}\hat{\mathcal{J}}^{\eta^{\delta}}_{b}\hat{p}\hat{\nabla}^{\eta^{\delta}}_{b}\cdot\partial_{t}\hat{\boldsymbol{\eta}}+\int_{\hat{\Omega}_{b}}\hat{\mathcal{J}}^{\eta^{\delta}}_{b}\partial_{t}\hat{\boldsymbol{\eta}}\cdot\hat{\nabla}^{\eta^{\delta}}_{b}\hat{p}+\int_{\hat{\Gamma}}(\hat{\zeta}\boldsymbol{e}_{y}\cdot\hat{\boldsymbol{n}}^{\omega^{\delta}})\hat{p}\right)
=α(Ωbδ(t)p𝝃+Ωbδ(t)𝝃p+Γδ(t)(ζ𝒆y𝒏δ)p)=0,\displaystyle=\alpha\left(\int_{\Omega_{b}^{\delta}(t)}p\nabla\cdot\boldsymbol{\xi}+\int_{\Omega_{b}^{\delta}(t)}\boldsymbol{\xi}\cdot\nabla p+\int_{\Gamma^{\delta}(t)}(\zeta\boldsymbol{e}_{y}\cdot\boldsymbol{n}^{\delta})p\right)=0,

where 𝒏δ\boldsymbol{n}^{\delta} is the upward pointing unit normal vector to Γδ(t)\Gamma^{\delta}(t). Finally, by the Reynold’s transport theorem

0TΩf(t)𝒖t𝒖+120TΓ(t)(ζ𝒆y𝒏)|𝒖|2=12Ωf(T)|𝒖|212Ωf(0)|𝒖|2.\int_{0}^{T}\int_{\Omega_{f}(t)}\boldsymbol{u}\cdot\partial_{t}\boldsymbol{u}+\frac{1}{2}\int_{0}^{T}\int_{\Gamma(t)}(\zeta\boldsymbol{e}_{y}\cdot\boldsymbol{n})|\boldsymbol{u}|^{2}=\frac{1}{2}\int_{\Omega_{f}(T)}|\boldsymbol{u}|^{2}-\frac{1}{2}\int_{\Omega_{f}(0)}|\boldsymbol{u}|^{2}.

By combining these calculations one obtains the final energy estimate:

12Ωf(T)|𝒖(T)|2+2ν0TΩf(t)|𝑫(𝒖)|2+β0TΓ(t)|(𝝃𝒖)𝝉|2+12ρpΓ^|tω^(T)|2+Γ^|Δ^ω^(T)|2+12ρbΩ^b|t𝜼^(T)|2+2μeΩ^b|𝑫^(𝜼^)(T)|2+2λeΩ^b|^𝜼^(T)|2+2μv0TΩ^b|𝑫^(t𝜼^)|2+λv0TΩ^b|^t𝜼^|2+12c0Ω^b|p^(T)|2+κ0TΩbδ(t)|p|2=12Ωf(0)|𝒖(0)|2+12ρpΓ^|tω^(0)|2+Γ^|Δ^ω^(0)|2+12ρbΩ^b|t𝜼^(0)|2+2μeΩ^b|𝑫^(𝜼^)(0)|2+2λeΩ^b|^𝜼^(0)|2+12c0Ω^b|p^(0)|2.\frac{1}{2}\int_{\Omega_{f}(T)}|\boldsymbol{u}(T)|^{2}+2\nu\int_{0}^{T}\int_{\Omega_{f}(t)}|\boldsymbol{D}(\boldsymbol{u})|^{2}+\beta\int_{0}^{T}\int_{\Gamma(t)}|(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}|^{2}+\frac{1}{2}\rho_{p}\int_{\hat{\Gamma}}|\partial_{t}\hat{\omega}(T)|^{2}+\int_{\hat{\Gamma}}|\hat{\Delta}\hat{\omega}(T)|^{2}\\ +\frac{1}{2}\rho_{b}\int_{\hat{\Omega}_{b}}|\partial_{t}\hat{\boldsymbol{\eta}}(T)|^{2}+2\mu_{e}\int_{\hat{\Omega}_{b}}|\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}})(T)|^{2}+2\lambda_{e}\int_{\hat{\Omega}_{b}}|\hat{\nabla}\cdot\hat{\boldsymbol{\eta}}(T)|^{2}+2\mu_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}|\hat{\boldsymbol{D}}(\partial_{t}\hat{\boldsymbol{\eta}})|^{2}\\ +\lambda_{v}\int_{0}^{T}\int_{\hat{\Omega}_{b}}|\hat{\nabla}\cdot\partial_{t}\hat{\boldsymbol{\eta}}|^{2}+\frac{1}{2}c_{0}\int_{\hat{\Omega}_{b}}|\hat{p}(T)|^{2}+\kappa\int_{0}^{T}\int_{\Omega_{b}^{\delta}(t)}|\nabla p|^{2}=\frac{1}{2}\int_{\Omega_{f}(0)}|\boldsymbol{u}(0)|^{2}+\frac{1}{2}\rho_{p}\int_{\hat{\Gamma}}|\partial_{t}\hat{\omega}(0)|^{2}\\ +\int_{\hat{\Gamma}}|\hat{\Delta}\hat{\omega}(0)|^{2}+\frac{1}{2}\rho_{b}\int_{\hat{\Omega}_{b}}|\partial_{t}\hat{\boldsymbol{\eta}}(0)|^{2}+2\mu_{e}\int_{\hat{\Omega}_{b}}|\hat{\boldsymbol{D}}(\hat{\boldsymbol{\eta}})(0)|^{2}+2\lambda_{e}\int_{\hat{\Omega}_{b}}|\hat{\nabla}\cdot\hat{\boldsymbol{\eta}}(0)|^{2}+\frac{1}{2}c_{0}\int_{\hat{\Omega}_{b}}|\hat{p}(0)|^{2}.

5.3 Statement of the main existence result for the regularized problem

We now state the main result on the existence of a weak solution to the regularized problem.

Theorem 5.1.

Let ρb,μe,λe,α,ρp,ν>0\rho_{b},\mu_{e},\lambda_{e},\alpha,\rho_{p},\nu>0 and μv,λv0\mu_{v},\lambda_{v}\geq 0. Consider initial data for the plate displacement ω^0H02(Γ^)\hat{\omega}_{0}\in H_{0}^{2}(\hat{\Gamma}), plate velocity ζ^0L2(Γ^)\hat{\zeta}_{0}\in L^{2}(\hat{\Gamma}), Biot displacement 𝜼^0H1(Ω^b)\hat{\boldsymbol{\eta}}_{0}\in H^{1}(\hat{\Omega}_{b}), Biot velocity 𝝃^0H1(Ω^b)\hat{\boldsymbol{\xi}}_{0}\in H^{1}(\hat{\Omega}_{b}) in the case of a viscoelastic Biot medium μv,λv>0\mu_{v},\lambda_{v}>0 and 𝝃^0L2(Ω^b)\hat{\boldsymbol{\xi}}_{0}\in L^{2}(\hat{\Omega}_{b}) otherwise for the case of a purely elastic Biot medium, Biot pore pressure p^0L2(Ω^b)\hat{p}_{0}\in L^{2}(\hat{\Omega}_{b}), and fluid velocity 𝒖0H1(Ωf(0))\boldsymbol{u}_{0}\in H^{1}(\Omega_{f}(0)) which is divergence-free. Suppose further that |ω^0|R0<R|\hat{\omega}_{0}|\leq R_{0}<R for some R0R_{0}, 𝜼^0|Γ=ω^0𝒆y\hat{\boldsymbol{\eta}}_{0}|_{\Gamma}=\hat{\omega}_{0}\boldsymbol{e}_{y}, and 𝝃^0|Γ=ζ^0𝒆y\hat{\boldsymbol{\xi}}_{0}|_{\Gamma}=\hat{\zeta}_{0}\boldsymbol{e}_{y}, and for some arbitrary but fixed regularization parameter δ>0\delta>0, suppose that Id+𝜼^0δ\text{Id}+\hat{\boldsymbol{\eta}}_{0}^{\delta} is an invertible map with det(𝑰+𝜼^0δ)>0\det(\boldsymbol{I}+\nabla\hat{\boldsymbol{\eta}}_{0}^{\delta})>0. Then, there exists a weak solution (𝒖,ω^,𝜼^,p^)(\boldsymbol{u},\hat{\omega},\hat{\boldsymbol{\eta}},\hat{p}) to the regularized FPSI problem with regularization parameter δ\delta on some time interval [0,T][0,T], for some T>0T>0.

While TT in general depends on δ\delta, we will show that if there exists a smooth solution to the nonregularized FPSI problem, then this time TT for the regularized problem is independent of δ\delta. This will allow us to pass to the limit as δ0\delta\to 0 and show that weak solutions to the regularized FPSI problems constructed in this manuscript, converge to a smooth solution of the original, nonregularized problem, when a smooth solution to the nonregularized problem exists. On the other hand, without the additional assumption of the existence of a strong solution, one cannot draw any conclusions about the limit as δ0\delta\to 0. This assumption plays a crucial role in demonstrating that TT remains independent of δ\delta. Furthermore, as elucidated in Remark 4.1, energy estimates alone are insufficient to take the limit in certain terms in the weak formulation. Therefore, in order to pass to the limit as δ0\delta\to 0 one would need to prove additional regularity estimates (beyond energy estimates) for weak solutions, which appears to be beyond the current state-of-the-art techniques.

Remark 5.5.

The result above is a local result, since it holds up to some time T>0T>0, which needs to be sufficiently small. However, it is easy to show that this T>0T>0 can be made maximal, in the sense that it holds until the time for which Id+𝜼^δ\text{Id}+\hat{\boldsymbol{\eta}}^{\delta} fails to be invertible or |ω^(,x)|=R|\hat{\omega}(\cdot,x)|=R for some xΓ^x\in\hat{\Gamma} when the reticular plate collides with the boundary. This can be shown using a standard method, see e.g., pg. 397-398 of [21], or the proof of Theorem 7.1 in [52].

An important notational convention. For notational simplicity, we will no longer use the “hat” notation to distinguish between functions and domains in the physical or reference configuration: for example, we will denote both the pore pressure pp on Ωb(t)\Omega_{b}(t) and p^\hat{p} on Ω^b\hat{\Omega}_{b} by pp, as the distinction between these two will be clear from context. In addition, we will remove the “hat” convention from the reference domains, and for example, we will denote the reference domain Ω^b\hat{\Omega}_{b} for the Biot medium by Ωb\Omega_{b}. We will follow this notational convention for the rest of the manuscript.

The proof of Theorem 5.1 is constructive, and based on an operator splitting scheme. This is an approach that has been used in constructive existence proofs of weak solutions for a variety of FSI problems, see for example [52].

6 The splitting scheme

The splitting scheme is defined as follows. First, semidiscretize the problem in time by introducing the time step Δt=T/N\Delta t=T/N, and subdivide the time interval [0,T][0,T] into NN subintervals, each of width Δt\Delta t. On each subinterval, we will run two subproblems: (1) a plate subproblem which takes into account the elastodynamics of the reticular plate and updates the plate displacement and the plate velocity, and (2) a fluid/Biot subproblem which updates the fluid velocity, the Biot displacement, the Biot pore pressure, and the plate velocity. Hence, each subinterval [nΔt,(n+1)Δt][n\Delta t,(n+1)\Delta t] involves an iteration of the plate subproblem and then an iteration of the fluid/Biot subproblem, and the solution from the previous subproblem is used as data for the subsequent subproblem. The approximations of the fluid velocity, plate displacement and velocity, and Biot poroviscoelastic material displacement and pressure will be denoted by

(𝒖Nn+i2,ωNn+i2,ζNn+i2,𝜼Nn+i2,pNn+i2), for n=0,1,.,N and i=0,1,(\boldsymbol{u}^{n+\frac{i}{2}}_{N},\omega^{n+\frac{i}{2}}_{N},\zeta^{n+\frac{i}{2}}_{N},\boldsymbol{\eta}^{n+\frac{i}{2}}_{N},p^{n+\frac{i}{2}}_{N}),\qquad\text{ for }n=0,1,....,N\text{ and }i=0,1,

where they are all defined on the given time subinterval [nΔt,(n+1)Δt][n\Delta t,(n+1)\Delta t]. Here, the quantities with the superscript n+1/2n+1/2 denote the resulting approximate solutions obtained after the plate subproblem is solved, and the quantities with the superscript n+1n+1 denote the resulting approximate solutions obtained after the fluid/Biot subproblem is solved. For the splitting scheme we will work on the fixed reference domain and hence, we will semi-discretize the regularized weak formulation (45) on the fixed reference domain. Backward Euler discretization will be used to approximate time derivatives, with the following shorthand notation:

f˙Nn+i2=fNn+i2fNn+i21Δt.\dot{f}^{n+\frac{i}{2}}_{N}=\frac{f^{n+\frac{i}{2}}_{N}-f^{n+\frac{i}{2}-1}_{N}}{\Delta t}. (47)

As a technical comment, in the description of the subproblems below, the backward Euler discretization (47) can potentially give rise to negative subscripts, so when relevant, we will explicitly define f1f_{-1} and f1/2f_{-1/2} depending on the context.

6.1 The plate subproblem

Only the plate displacement and velocity ωNn+12\omega^{n+\frac{1}{2}}_{N} and ζNn+12\zeta^{n+\frac{1}{2}}_{N} are updated in this subproblem, leaving the remaining variables unchanged:

𝒖Nn+12=𝒖Nn,𝜼Nn+12=𝜼Nn,pNn+12=pNn.\boldsymbol{u}^{n+\frac{1}{2}}_{N}=\boldsymbol{u}^{n}_{N},\qquad\boldsymbol{\eta}^{n+\frac{1}{2}}_{N}=\boldsymbol{\eta}^{n}_{N},\qquad p^{n+\frac{1}{2}}_{N}=p^{n}_{N}.

The new plate displacement and velocity are calculated from the following weak formulation of the plate subproblems: find ωNn+12H02(Γ)\omega^{n+\frac{1}{2}}_{N}\in H_{0}^{2}(\Gamma) and ζNn+12H02(Γ)\zeta^{n+\frac{1}{2}}_{N}\in H_{0}^{2}(\Gamma), such that

Γ(ωNn+12ωNn12Δt)ϕ=ΓζNn+12ϕ, for all ϕL2(Γ),\int_{\Gamma}\left(\frac{\omega^{n+\frac{1}{2}}_{N}-\omega^{n-\frac{1}{2}}_{N}}{\Delta t}\right)\cdot\phi=\int_{\Gamma}\zeta^{n+\frac{1}{2}}_{N}\cdot\phi,\qquad\text{ for all }\phi\in L^{2}(\Gamma), (48)
ρpΓ(ζNn+12ζNnΔt)φ+ΓΔωNn+12Δφ=0, for all φH02(Γ).\rho_{p}\int_{\Gamma}\left(\frac{\zeta^{n+\frac{1}{2}}_{N}-\zeta^{n}_{N}}{\Delta t}\right)\cdot\varphi+\int_{\Gamma}\Delta\omega^{n+\frac{1}{2}}_{N}\cdot\Delta\varphi=0,\qquad\text{ for all }\varphi\in H_{0}^{2}(\Gamma). (49)

When n=0n=0, we set ωN12=ω(0)\omega^{-\frac{1}{2}}_{N}=\omega(0) and ζN0=ζ(0)\zeta^{0}_{N}=\zeta(0). In particular, ω(0)𝒆y=𝜼(0)|Γ\omega(0)\boldsymbol{e}_{y}=\boldsymbol{\eta}(0)|_{\Gamma} and ζ(0)𝒆y=𝝃(0)\zeta(0)\boldsymbol{e}_{y}=\boldsymbol{\xi}(0).

Lemma 6.1.

Problem (48), (49) has a unique solution which satisfies the following energy equality:

12ρpΓ|ζNn+12|2+12ρpΓ|ζNn+12ζNn|2+12Γ|ΔωNn+12|2+12Γ|Δ(ωNn+12ωNn12)|2\displaystyle\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+\frac{1}{2}}_{N}|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+\frac{1}{2}}_{N}-\zeta^{n}_{N}|^{2}+\frac{1}{2}\int_{\Gamma}|\Delta\omega^{n+\frac{1}{2}}_{N}|^{2}+\frac{1}{2}\int_{\Gamma}|\Delta(\omega^{n+\frac{1}{2}}_{N}-\omega^{n-\frac{1}{2}}_{N})|^{2}
=12ρpΓ|ζNn12|2+12Γ|ΔωNn12|2.\displaystyle=\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n-\frac{1}{2}}_{N}|^{2}+\frac{1}{2}\int_{\Gamma}|\Delta\omega^{n-\frac{1}{2}}_{N}|^{2}. (50)
Proof.

To prove this, we first notice that (48) immediately implies that

ζNn+12=ωNn+12ωNn12Δt\zeta^{n+\frac{1}{2}}_{N}=\frac{\omega^{n+\frac{1}{2}}_{N}-\omega^{n-\frac{1}{2}}_{N}}{\Delta t} (51)

so that, by substituting into (49), it suffices to find ωNn+12H02(Γ)\omega^{n+\frac{1}{2}}_{N}\in H_{0}^{2}(\Gamma) which satisfies:

ρpΓωNn+12φ+(Δt)2ΓΔωNn+12Δφ=ρpΓ(ωNn12+(Δt)ζNn)φ, for all φH02(Γ).\rho_{p}\int_{\Gamma}\omega^{n+\frac{1}{2}}_{N}\cdot\varphi+(\Delta t)^{2}\int_{\Gamma}\Delta\omega^{n+\frac{1}{2}}_{N}\cdot\Delta\varphi=\rho_{p}\int_{\Gamma}(\omega^{n-\frac{1}{2}}_{N}+(\Delta t)\zeta^{n}_{N})\cdot\varphi,\qquad\text{ for all }\varphi\in H_{0}^{2}(\Gamma).

The bilinear form

B[ω,φ]=ρpΓωφ+(Δt)2ΓΔωΔφB[\omega,\varphi]=\rho_{p}\int_{\Gamma}\omega\cdot\varphi+(\Delta t)^{2}\int_{\Gamma}\Delta\omega\cdot\Delta\varphi

is coercive on H02(Γ)H^{2}_{0}(\Gamma), and

φρpΓ(ωNn12+(Δt)ζNn)φ\varphi\to\rho_{p}\int_{\Gamma}\left(\omega^{n-\frac{1}{2}}_{N}+(\Delta t)\zeta^{n}_{N}\right)\cdot\varphi

is a continuous linear functional on H02(Γ)H^{2}_{0}(\Gamma), since we will have ωNn12H02(Γ)\omega^{n-\frac{1}{2}}_{N}\in H_{0}^{2}(\Gamma) and ζNnL2(Γ)\zeta^{n}_{N}\in L^{2}(\Gamma) by the way our splitting scheme is defined. Thus, by the Lax-Milgram lemma, there exists a unique solution ωNn+12H02(Γ)\omega^{n+\frac{1}{2}}_{N}\in H_{0}^{2}(\Gamma), from which we also recover ζNn+12H02(Γ)\zeta^{n+\frac{1}{2}}_{N}\in H^{2}_{0}(\Gamma) using (51) above.

The energy equality above follows by substituting φ=ζNn+12=ωNn+12ωNn12ΔtH02(Γ)\varphi=\zeta^{n+\frac{1}{2}}_{N}=\frac{\omega^{n+\frac{1}{2}}_{N}-\omega^{n-\frac{1}{2}}_{N}}{\Delta t}\in H_{0}^{2}(\Gamma) into the weak formulation and using the identity

(ab)a=12(|a|2+|ab|2|b|2).(a-b)\cdot a=\frac{1}{2}(|a|^{2}+|a-b|^{2}-|b|^{2}).

6.2 The fluid and Biot subproblem

For the fluid and Biot subproblem, we update the quantities related to the fluid and the Biot medium. Due to the kinematic coupling between the Biot medium displacement and the plate displacement, we must also update the plate velocity, as the dynamics of the Biot medium affect the kinematics of the plate. In this step, only the plate displacement remains unchanged:

ωNn+1=ωNn+12.\omega^{n+1}_{N}=\omega^{n+\frac{1}{2}}_{N}.

To state the weak formulation of the fluid and Biot subproblem, we define the solution and test spaces, respectively:

𝒱Nn+1\displaystyle\mathcal{V}^{n+1}_{N} =\displaystyle= {(𝒖,ζ,𝜼,p)𝒱fωNn×H02(Γ)×Vd×Vp},\displaystyle\{(\boldsymbol{u},\zeta,\boldsymbol{\eta},p)\in\mathcal{V}^{\omega^{n}_{N}}_{f}\times H_{0}^{2}(\Gamma)\times V_{d}\times V_{p}\}, (52)
𝒬Nn+1\displaystyle\mathcal{Q}^{n+1}_{N} =\displaystyle= {(𝒗,φ,𝝍,r)VfωNn×H02(Γ)×Vd×Vp:𝝍=φ𝒆y on Γ},\displaystyle\{(\boldsymbol{v},\varphi,\boldsymbol{\psi},r)\in V^{\omega^{n}_{N}}_{f}\times H_{0}^{2}(\Gamma)\times V_{d}\times V_{p}:\boldsymbol{\psi}=\varphi\boldsymbol{e}_{y}\text{ on }\Gamma\}, (53)

where VfωV^{\omega}_{f}, VdV_{d}, and VpV_{p} are defined in (29), (32), and (34).

The weak formulation now reads: find (𝒖Nn+1,ζNn+1,𝜼Nn+1,pNn+1)𝒱Nn+1(\boldsymbol{u}^{n+1}_{N},\zeta^{n+1}_{N},\boldsymbol{\eta}^{n+1}_{N},p^{n+1}_{N})\in\mathcal{V}^{n+1}_{N} defined on the reference domain, such that for all test functions (𝒗,φ,𝝍,r)𝒬Nn+1(\boldsymbol{v},\varphi,\boldsymbol{\psi},r)\in\mathcal{Q}^{n+1}_{N} defined on the reference domain, the following holds:

Ωf(1+ωNnR)𝒖˙Nn+1𝒗+2νΩf(1+ωNnR)𝑫fωNn(𝒖Nn+1):𝑫fωNn(𝒗)+Γ(12𝒖Nn+1𝒖NnpNn+1)(𝝍𝒗)𝒏ωNn\displaystyle\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\boldsymbol{\dot{u}}^{n+1}_{N}\cdot\boldsymbol{v}+2\nu\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{u}^{n+1}_{N}):\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{v})+\int_{\Gamma}\left(\frac{1}{2}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{u}^{n}_{N}-p^{n+1}_{N}\right)(\boldsymbol{\psi}-\boldsymbol{v})\cdot\boldsymbol{n}^{\omega^{n}_{N}} (54)
+12Ωf(1+ωNnR)[((𝒖NnζNn+12R+yR𝒆y)fωNn𝒖Nn+1)𝒗((𝒖NnζNn+12R+yR𝒆y)fωNn𝒗)𝒖Nn+1]\displaystyle+\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\left[\left(\left(\boldsymbol{u}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\omega^{n}_{N}}_{f}\boldsymbol{u}^{n+1}_{N}\right)\cdot\boldsymbol{v}-\left(\left(\boldsymbol{u}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\omega^{n}_{N}}_{f}\boldsymbol{v}\right)\cdot\boldsymbol{u}^{n+1}_{N}\right]
+12RΩfζNn+12𝒖Nn+1𝒗+12Γ(𝒖Nn+1𝜼˙Nn+1)𝒏ωNn(𝒖Nn𝒗)\displaystyle+\frac{1}{2R}\int_{\Omega_{f}}\zeta^{n+\frac{1}{2}}_{N}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{v}+\frac{1}{2}\int_{\Gamma}(\boldsymbol{u}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}^{\omega^{n}_{N}}(\boldsymbol{u}^{n}_{N}\cdot\boldsymbol{v})
+β𝒥ΓωNnΓ(𝜼˙Nn+1𝒖Nn+1)𝝉ωNn(𝝍𝒗)𝝉ωNn+ρbΩb(𝜼˙Nn+1𝜼˙NnΔt)𝝍\displaystyle+\frac{\beta}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}\int_{\Gamma}(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}(\boldsymbol{\psi}-\boldsymbol{v})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}+\rho_{b}\int_{\Omega_{b}}\left(\frac{\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n}_{N}}{\Delta t}\right)\cdot\boldsymbol{\psi}
+ρpΓ(ζNn+1ζNn+12Δt)φ+2μeΩb𝑫(𝜼Nn+1):𝑫(𝝍)+λeΩb(𝜼Nn+1)(𝝍)\displaystyle+\rho_{p}\int_{\Gamma}\left(\frac{\zeta^{n+1}_{N}-\zeta^{n+\frac{1}{2}}_{N}}{\Delta t}\right)\varphi+2\mu_{e}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{e}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}^{n+1}_{N})(\nabla\cdot\boldsymbol{\psi})
+2μvΩb𝑫(𝜼˙Nn+1):𝑫(𝝍)+λvΩb(𝜼˙Nn+1)(𝝍)αΩb𝒥b(ηNn)δpNn+1b(ηNn)δ𝝍\displaystyle+2\mu_{v}\int_{\Omega_{b}}\boldsymbol{D}(\dot{\boldsymbol{\eta}}^{n+1}_{N}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{v}\int_{\Omega_{b}}(\nabla\cdot\dot{\boldsymbol{\eta}}^{n+1}_{N})(\nabla\cdot\boldsymbol{\psi})-\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\psi}
+c0ΩbpNn+1pNnΔtrαΩb𝒥b(ηNn)δ𝜼˙Nn+1b(ηNn)δrαΓ(𝜼˙Nn+1𝒏(ωNn)δ)r\displaystyle+c_{0}\int_{\Omega_{b}}\frac{p^{n+1}_{N}-p^{n}_{N}}{\Delta t}r-\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r-\alpha\int_{\Gamma}(\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}})r
+κΩb𝒥b(ηNn)δb(ηNn)δpNn+1b(ηNn)δrΓ[(𝒖Nn+1𝜼˙Nn+1)𝒏ωNn]r=0,\displaystyle+\kappa\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r-\int_{\Gamma}[(\boldsymbol{u}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}^{\omega^{n}_{N}}]r=0,

and

Γ(𝜼Nn+1𝜼NnΔt)ϕ=ΓζNn+1𝒆yϕ, for all ϕL2(Γ).\int_{\Gamma}\left(\frac{\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N}}{\Delta t}\right)\cdot\boldsymbol{\phi}=\int_{\Gamma}\zeta^{n+1}_{N}\boldsymbol{e}_{y}\cdot\boldsymbol{\phi},\qquad\text{ for all }\boldsymbol{\phi}\in L^{2}(\Gamma). (55)

We remark that when n=0n=0, the backwards Euler discretization (47) involves a negative subscript in the definition of 𝜼˙N0\dot{\boldsymbol{\eta}}^{0}_{N}, so in this case, we will instead set 𝜼˙N0\dot{\boldsymbol{\eta}}^{0}_{N} to be the initial plate velocity ζ0\zeta_{0}.

Lemma 6.2.

Problem (54), (55) has a unique solution provided that the following assumptions hold:

  1. 1.

    Assumption 1A: Boundedness of the plate displacement away from RR. There exists a positive constant RmaxR_{max} such that

    |ωNk+i2|Rmax<R, for all k=0,1,,n and i=0,1.|\omega^{k+\frac{i}{2}}_{N}|\leq R_{max}<R,\qquad\text{ for all }k=0,1,...,n\text{ and }i=0,1. (56)
  2. 2.

    Assumption 2A: Invertibility of the map from fixed to moving Biot domain. The map

    Id+(𝜼Nn)δ:Ωb(Ωb)Nn,δ is invertible,\text{Id}+(\boldsymbol{\eta}^{n}_{N})^{\delta}:\Omega_{b}\to(\Omega_{b})^{n,\delta}_{N}\qquad\text{ is invertible}, (57)

    where we define (Ωb)Nn,δ(\Omega_{b})^{n,\delta}_{N} to be the image of Ωb\Omega_{b} under the map Id+(𝜼Nn)δ\text{Id}+(\boldsymbol{\eta}^{n}_{N})^{\delta}.

Additionally, the weak solution satisfies the following energy equality:

12Ωf(1+ωNn+1R)|𝒖Nn+1|2+12ρbΩb|𝜼˙Nn+1|2+12c0Ωb|pNn+1|2+μeΩb|𝑫(𝜼Nn+1)|2+12λeΩb|𝜼Nn+1|2\displaystyle\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n+1}_{N}}{R}\right)|\boldsymbol{u}^{n+1}_{N}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\dot{\boldsymbol{\eta}}^{n+1}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p^{n+1}_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N})|^{2}+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}^{n+1}_{N}|^{2}
+12ρpΓ|ζNn+1|2+2μv(Δt)Ωb|𝑫(𝜼˙Nn+1)|2+λv(Δt)Ωb|𝜼˙Nn+1|2+κ(Δt)Ωb𝒥b(ηNn)δ|b(ηNn)δpNn+1|2\displaystyle+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+1}_{N}|^{2}+2\mu_{v}(\Delta t)\int_{\Omega_{b}}|\boldsymbol{D}(\dot{\boldsymbol{\eta}}^{n+1}_{N})|^{2}+\lambda_{v}(\Delta t)\int_{\Omega_{b}}|\nabla\cdot\dot{\boldsymbol{\eta}}^{n+1}_{N}|^{2}+\kappa(\Delta t)\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}|^{2}
+β(Δt)𝒥ΓωNnΓ|(𝜼˙Nn+1𝒖Nn+1)𝝉ωNn|2+12ρbΩb|𝜼˙Nn+1𝜼˙Nn|2+12c0Ωb|pNn+1pNn|2+μeΩb|𝑫(𝜼Nn+1𝜼Nn)|2\displaystyle+\frac{\beta(\Delta t)}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}\int_{\Gamma}|(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p^{n+1}_{N}-p^{n}_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N})|^{2}
+12λeΩb|(𝜼Nn+1𝜼Nn)|2=12Ωf(1+ωNnR)|𝒖Nn|2+12ρbΩb|𝜼˙Nn|2+12c0Ωb|pNn|2+μeΩb|𝑫(𝜼Nn)|2\displaystyle+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot(\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N})|^{2}=\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)|\boldsymbol{u}^{n}_{N}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\dot{\boldsymbol{\eta}}^{n}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p^{n}_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}^{n}_{N})|^{2}
+12λeΩb|𝜼Nn|2+12ρpΓ|ζNn+12|2.\displaystyle+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}^{n}_{N}|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+\frac{1}{2}}_{N}|^{2}.

The proof is based on using the Lax-Milgram Lemma. However, in this case the proof is more involved for two reasons. First, the bilinear form associated with problem (54) and (55) is not coercive on the Hilbert space 𝒱fωNn×Vd×Vp,\mathcal{V}^{\omega^{n}_{N}}_{f}\times V_{d}\times V_{p}, because of a mismatch between the hyperbolic and parabolic scaling in the problem. The second reason is that it is not a priori clear that Korn’s inequality, which is needed in the proof of the existence, holds for the Biot domain. To deal with the first difficulty and recover the coercive structure of the problem, the test functions can be rescaled by the factor Δt\Delta t so that

𝒗(Δt)𝒗,r(Δt)r.\boldsymbol{v}\to(\Delta t)\boldsymbol{v},\qquad r\to(\Delta t)r. (58)

This scaling of the test functions is valid because if (𝒗,φ,𝝍,v)𝒬Nn+1(\boldsymbol{v},\varphi,\boldsymbol{\psi},v)\in\mathcal{Q}^{n+1}_{N}, then the rescaled test function satisfies ((Δt)1𝒗,φ,𝝍,(Δt)1r)𝒬Nn+1((\Delta t)^{-1}\boldsymbol{v},\varphi,\boldsymbol{\psi},(\Delta t)^{-1}r)\in\mathcal{Q}^{n+1}_{N} also. To deal with the second difficulty, one can show by explicit calculation that the following Korn’s inequality holds for this problem. We refer the reader to Section 11.2 and Corollary 11.2.22 in [16] for a more general proof of the Korn inequality.

Proposition 6.1.

Korn’s inequality for the Biot poroviscoelastic domain. For all 𝜼Vd\boldsymbol{\eta}\in V_{d},

Ωb|𝑫(𝜼)|212Ωb|𝜼|2.\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta})|^{2}\geq\frac{1}{2}\int_{\Omega_{b}}|\nabla\boldsymbol{\eta}|^{2}.
Proof.

By a standard approximation argument, it suffices to assume that 𝜼\boldsymbol{\eta} is smooth. Because ηx=0\eta_{x}=0 on Γ\Gamma and because 𝜼=0\boldsymbol{\eta}=0 on the left, top, and right boundaries of Ωb\Omega_{b}, we have from integration by parts, that

Ωbηxyηyx=Ωbηx2ηyxy=Ωbηxxηyy.\int_{\Omega_{b}}\frac{\partial\eta_{x}}{\partial y}\frac{\partial\eta_{y}}{\partial x}=-\int_{\Omega_{b}}\eta_{x}\frac{\partial^{2}\eta_{y}}{\partial x\partial y}=\int_{\Omega_{b}}\frac{\partial\eta_{x}}{\partial x}\frac{\partial\eta_{y}}{\partial y}.

Therefore, by using the inequality a2+2ab+b20a^{2}+2ab+b^{2}\geq 0, we obtain

Ωb|𝑫(𝜼)|2\displaystyle\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta})|^{2} =Ωb(ηxx)2+(ηyy)2+2[12(ηxy+ηyx)]2\displaystyle=\int_{\Omega_{b}}\left(\frac{\partial\eta_{x}}{\partial x}\right)^{2}+\left(\frac{\partial\eta_{y}}{\partial y}\right)^{2}+2\left[\frac{1}{2}\left(\frac{\partial\eta_{x}}{\partial y}+\frac{\partial\eta_{y}}{\partial x}\right)\right]^{2}
=Ωb(ηxx)2+(ηyy)2+12(ηxy+ηyx)2\displaystyle=\int_{\Omega_{b}}\left(\frac{\partial\eta_{x}}{\partial x}\right)^{2}+\left(\frac{\partial\eta_{y}}{\partial y}\right)^{2}+\frac{1}{2}\left(\frac{\partial\eta_{x}}{\partial y}+\frac{\partial\eta_{y}}{\partial x}\right)^{2}
=Ωb(ηxx)2+(ηyy)2+12[(ηxy)2+(ηyx)2]+ηxyηyx\displaystyle=\int_{\Omega_{b}}\left(\frac{\partial\eta_{x}}{\partial x}\right)^{2}+\left(\frac{\partial\eta_{y}}{\partial y}\right)^{2}+\frac{1}{2}\left[\left(\frac{\partial\eta_{x}}{\partial y}\right)^{2}+\left(\frac{\partial\eta_{y}}{\partial x}\right)^{2}\right]+\frac{\partial\eta_{x}}{\partial y}\frac{\partial\eta_{y}}{\partial x}
=Ωb(ηxx)2+ηxxηyy+(ηyy)2+12[(ηxy)2+(ηyx)2]12Ωb|𝜼|2.\displaystyle=\int_{\Omega_{b}}\left(\frac{\partial\eta_{x}}{\partial x}\right)^{2}+\frac{\partial\eta_{x}}{\partial x}\frac{\partial\eta_{y}}{\partial y}+\left(\frac{\partial\eta_{y}}{\partial y}\right)^{2}+\frac{1}{2}\left[\left(\frac{\partial\eta_{x}}{\partial y}\right)^{2}+\left(\frac{\partial\eta_{y}}{\partial x}\right)^{2}\right]\geq\frac{1}{2}\int_{\Omega_{b}}|\nabla\boldsymbol{\eta}|^{2}.

Proof.

Proof of Lemma 6.2. Rewrite the weak formulation (54) and (55) so that all of the functions at the (n+1)(n+1)st time step are on the left hand side while all other quantities are on the right hand side. In addition, we rewrite ζNn+1\zeta^{n+1}_{N} in terms of 𝜼Nn\boldsymbol{\eta}^{n}_{N} and 𝜼Nn+1\boldsymbol{\eta}^{n+1}_{N} by using (55):

ζNn+1𝒆y=𝜼Nn+1𝜼NnΔt|Γ.\zeta^{n+1}_{N}\boldsymbol{e}_{y}=\frac{\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N}}{\Delta t}\Big{|}_{\Gamma}.

After using the rescaling (58) of the test functions, the weak formulation involves the following coercive and continuous bilinear form B:H×HB:H\times H\to\mathbb{R}, where HH is the Hilbert space 𝒱fωNn×Vd×Vp\mathcal{V}^{\omega^{n}_{N}}_{f}\times V_{d}\times V_{p}:

B([𝒖,𝜼,p),(𝒗,𝝍,r)]:=(Δt)2Ωf(1+ωNnR)𝒖𝒗\displaystyle{{B([\boldsymbol{u},\boldsymbol{\eta},p),(\boldsymbol{v},\boldsymbol{\psi},r)]}}:=(\Delta t)^{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\boldsymbol{u}\cdot\boldsymbol{v}
+12(Δt)3Ωf(1+ωNnR)[((𝒖NnζNn+12R+yR𝒆y)ωNn𝒖)𝒗((𝒖NnζNn+12R+yR𝒆y)ωNn𝒗)𝒖]\displaystyle+\frac{1}{2}(\Delta t)^{3}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\left[\left(\left(\boldsymbol{u}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\omega^{n}_{N}}\boldsymbol{u}\right)\cdot\boldsymbol{v}-\left(\left(\boldsymbol{u}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\omega^{n}_{N}}\boldsymbol{v}\right)\cdot\boldsymbol{u}\right]
+(Δt)312RΩfζNn+12𝒖𝒗+12(Δt)3Γ(𝒖(Δt)1𝜼)𝒏ωNn(𝒖Nn𝒗)\displaystyle+(\Delta t)^{3}\cdot\frac{1}{2R}\int_{\Omega_{f}}\zeta^{n+\frac{1}{2}}_{N}\boldsymbol{u}\cdot\boldsymbol{v}+\frac{1}{2}(\Delta t)^{3}\int_{\Gamma}(\boldsymbol{u}-(\Delta t)^{-1}\boldsymbol{\eta})\cdot\boldsymbol{n}^{\omega^{n}_{N}}(\boldsymbol{u}^{n}_{N}\cdot\boldsymbol{v})
+2ν(Δt)3Ωf(1+ωNnR)𝑫fωNn(𝒖):𝑫fωNn(𝒗)+(Δt)2Γ(12𝒖𝒖Nnp)(𝝍(Δt)𝒗)𝒏ωNn\displaystyle+2\nu(\Delta t)^{3}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{u}):\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{v})+(\Delta t)^{2}\int_{\Gamma}\left(\frac{1}{2}\boldsymbol{u}\cdot\boldsymbol{u}^{n}_{N}-p\right)(\boldsymbol{\psi}-(\Delta t)\boldsymbol{v})\cdot\boldsymbol{n}^{\omega^{n}_{N}}
+β𝒥ΓωNn(Δt)2Γ[(Δt)1𝜼𝒖]𝝉ωNn(𝝍(Δt)𝒗)𝝉ωNn+ρbΩb𝜼𝝍+ρpΓ𝜼𝝍\displaystyle+\frac{\beta}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}(\Delta t)^{2}\int_{\Gamma}[(\Delta t)^{-1}\boldsymbol{\eta}-\boldsymbol{u}]\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}(\boldsymbol{\psi}-(\Delta t)\boldsymbol{v})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}+\rho_{b}\int_{\Omega_{b}}\boldsymbol{\eta}\cdot\boldsymbol{\psi}+\rho_{p}\int_{\Gamma}\boldsymbol{\eta}\cdot\boldsymbol{\psi}
+(2μe(Δt)2+2μv(Δt))Ωb𝑫(𝜼):𝑫(𝝍)+(λe(Δt)2+λv(Δt))Ωb(𝜼)(𝝍)\displaystyle+(2\mu_{e}(\Delta t)^{2}+2\mu_{v}(\Delta t))\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}):\boldsymbol{D}(\boldsymbol{\psi})+(\lambda_{e}(\Delta t)^{2}+\lambda_{v}(\Delta t))\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta})(\nabla\cdot\boldsymbol{\psi})
α(Δt)2Ωb𝒥b(ηNn)δpb(ηNn)δ𝝍+c0(Δt)2Ωbprα(Δt)2Ωb𝒥b(ηNn)δ𝜼b(ηNn)δr\displaystyle-\alpha(\Delta t)^{2}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\psi}+c_{0}(\Delta t)^{2}\int_{\Omega_{b}}pr-\alpha(\Delta t)^{2}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\eta}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r
α(Δt)2Γ(𝜼𝒏(ωNn)δ)r+κ(Δt)3Ωb𝒥b(ηNn)δb(ηNn)δpb(ηNn)δr\displaystyle-\alpha(\Delta t)^{2}\int_{\Gamma}(\boldsymbol{\eta}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}})r+\kappa(\Delta t)^{3}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r
(Δt)3Γ[(𝒖(Δt)1𝜼)𝒏ωNn]r.\displaystyle-(\Delta t)^{3}\int_{\Gamma}[(\boldsymbol{u}-(\Delta t)^{-1}\boldsymbol{\eta})\cdot\boldsymbol{n}^{\omega^{n}_{N}}]r.

With this notation, the weak formulation reads: find (𝒖Nn+1,𝜼Nn+1,pNn+1)𝒱fωNn×Vd×Vp(\boldsymbol{u}^{n+1}_{N},\boldsymbol{\eta}^{n+1}_{N},p^{n+1}_{N})\in\mathcal{V}^{\omega^{n}_{N}}_{f}\times V_{d}\times V_{p} such that for all test functions (𝒗,𝝍,r)𝒱fωNn×Vd×Vp(\boldsymbol{v},\boldsymbol{\psi},r)\in\mathcal{V}^{\omega^{n}_{N}}_{f}\times V_{d}\times V_{p},

B[(𝒖Nn+1,𝜼Nn+1,pNn+1),(𝒗,𝝍,r)]=(Δt)2Ωf(1+ωNnR)𝒖Nn𝒗12(Δt)2Γ𝜼Nn𝒏ωNn(𝒖Nn𝒗)+β𝒥ΓωNn(Δt)Γ𝜼Nn𝝉ωNn(𝝍(Δt)𝒗)𝝉ωNn+ρbΩb(2𝜼Nn𝜼Nn1)𝝍+ρpΓ(𝜼Nn+(Δt)ζNn+12𝒆y)𝝍+2μv(Δt)Ωb𝑫(𝜼Nn):𝑫(𝝍)+λv(Δt)Ωb(𝜼Nn)(𝝍)+c0(Δt)2ΩbpNnrα(Δt)2Ωb𝒥b(ηNn)δ𝜼Nnb(ηNn)δrα(Δt)2Γ(𝜼Nn𝒏(ωNn)δ)r+(Δt)2Γ(𝜼Nn𝒏ωNn)r.{{B[(\boldsymbol{u}^{n+1}_{N},\boldsymbol{\eta}^{n+1}_{N},p^{n+1}_{N}),(\boldsymbol{v},\boldsymbol{\psi},r)]}}=(\Delta t)^{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\boldsymbol{u}^{n}_{N}\cdot\boldsymbol{v}-\frac{1}{2}(\Delta t)^{2}\int_{\Gamma}\boldsymbol{\eta}^{n}_{N}\cdot\boldsymbol{n}^{\omega^{n}_{N}}(\boldsymbol{u}^{n}_{N}\cdot\boldsymbol{v})\\ +\frac{\beta}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}(\Delta t)\int_{\Gamma}\boldsymbol{\eta}^{n}_{N}\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}(\boldsymbol{\psi}-(\Delta t)\boldsymbol{v})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}+\rho_{b}\int_{\Omega_{b}}(2\boldsymbol{\eta}^{n}_{N}-\boldsymbol{\eta}^{n-1}_{N})\cdot\boldsymbol{\psi}+\rho_{p}\int_{\Gamma}(\boldsymbol{\eta}^{n}_{N}+(\Delta t)\zeta^{n+\frac{1}{2}}_{N}\boldsymbol{e}_{y})\cdot\boldsymbol{\psi}\\ +2\mu_{v}(\Delta t)\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}^{n}_{N}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{v}(\Delta t)\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}^{n}_{N})(\nabla\cdot\boldsymbol{\psi})\\ +c_{0}(\Delta t)^{2}\int_{\Omega_{b}}p^{n}_{N}r-\alpha(\Delta t)^{2}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\eta}^{n}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r-\alpha(\Delta t)^{2}\int_{\Gamma}(\boldsymbol{\eta}^{n}_{N}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}})r+(\Delta t)^{2}\int_{\Gamma}(\boldsymbol{\eta}^{n}_{N}\cdot\boldsymbol{n}^{\omega^{n}_{N}})r. (59)

We now show that the bilinear form B[(𝒖,𝜼,p),(𝒗,𝝍,r)]B[(\boldsymbol{u},\boldsymbol{\eta},p),(\boldsymbol{v},\boldsymbol{\psi},r)] is coercive and continuous as a bilinear form on the Hilbert space 𝒱fωNn×Vd×Vp,\mathcal{V}^{\omega^{n}_{N}}_{f}\times V_{d}\times V_{p}, with the inner product given by

(𝒖,𝜼,p),(𝒗,𝝍,r)=Ωf(𝒖𝒗+𝒖:𝒗)+Ωb(𝜼𝝍+𝜼:𝝍)+Ωb(pr+pr).\langle(\boldsymbol{u},\boldsymbol{\eta},p),(\boldsymbol{v},\boldsymbol{\psi},r)\rangle=\int_{\Omega_{f}}(\boldsymbol{u}\cdot\boldsymbol{v}+\nabla\boldsymbol{u}:\nabla\boldsymbol{v})+\int_{\Omega_{b}}(\boldsymbol{\eta}\cdot\boldsymbol{\psi}+\nabla\boldsymbol{\eta}:\nabla\boldsymbol{\psi})+\int_{\Omega_{b}}(p\cdot r+\nabla p\cdot\nabla r).

We focus on establishing coercivity, since continuity follows by standard arguments. To show coercivity we calculate B[(𝒖,𝒗,p),(𝒖,𝜼,p)]B[(\boldsymbol{u},\boldsymbol{v},p),(\boldsymbol{u},\boldsymbol{\eta},p)]. In this calculation we note that after integration by parts, the sum of the following terms becomes zero:

α(Δt)2Ωb𝒥b(ηNn)δpb(ηNn)δ𝜼α(Δt)2Ωb𝒥b(ηNn)δ𝜼b(ηNn)δpα(Δt)2Γ(𝜼𝒏(ωNn)δ)p=0.-\alpha(\Delta t)^{2}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\eta}-\alpha(\Delta t)^{2}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\eta}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p-\alpha(\Delta t)^{2}\int_{\Gamma}\left(\boldsymbol{\eta}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}}\right)p=0.

Indeed, to see this, we bring the integrals back to the time-dependent physical domain, which we can do as long as (ηNn)δ(\boldsymbol{\eta}^{n}_{N})^{\delta} is a bijection from Ωb\Omega_{b} to (Ωb)Nn,δ(\Omega_{b})^{n,\delta}_{N}, which is provided by Assumption 2A (57), and perform the following computation:

α(Δt)2Ωb𝒥b(ηNn)δpb(ηNn)δ𝜼α(Δt)2Ωb𝒥b(ηNn)δ𝜼b(ηNn)δpα(Δt)2Γ(𝜼𝒏(ωNn)δ)p\displaystyle-\alpha(\Delta t)^{2}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\eta}-\alpha(\Delta t)^{2}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\eta}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p-\alpha(\Delta t)^{2}\int_{\Gamma}\left(\boldsymbol{\eta}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}}\right)p
=α(Δt)2((Ωb)Nn,δp𝜼+(Ωb)Nn,δ𝜼p+ΓNn,δ(𝜼𝒏)p)=0,\displaystyle=-\alpha(\Delta t)^{2}\left(\int_{(\Omega_{b})^{n,\delta}_{N}}p\nabla\cdot\boldsymbol{\eta}+\int_{(\Omega_{b})^{n,\delta}_{N}}\boldsymbol{\eta}\cdot\nabla p+\int_{\Gamma^{n,\delta}_{N}}(\boldsymbol{\eta}\cdot\boldsymbol{n})p\right)=0,

where we used integration by parts, the fact that 𝒏\boldsymbol{n} points outwards from Ωf\Omega_{f} and hence inwards towards Ωb\Omega_{b}, and also use that 𝜼=0\boldsymbol{\eta}=0 on the left, right, and top boundaries of Ωb\Omega_{b}. Combining this with the fact that (Δt)ζNn+12=ωNn+12ωNn12=ωNn+1ωNn(\Delta t)\zeta^{n+\frac{1}{2}}_{N}=\omega^{n+\frac{1}{2}}_{N}-\omega^{n-\frac{1}{2}}_{N}=\omega^{n+1}_{N}-\omega^{n}_{N}, we obtain

B[(𝒖,𝜼,p),(𝒖,𝜼,p)]:=(Δt)2Ωf(1+ωNn+ωNn+12R)|𝒖|2+2ν(Δt)3Ωf(1+ωNnR)|𝑫fωNn(𝒖)|2+β𝒥ΓωNn(Δt)Γ|(𝜼(Δt)𝒖)𝝉ωNn|2+ρbΩb|𝜼|2+ρpΓ|𝜼|2+(2μe(Δt)2+2μv(Δt))Ωb|𝑫(𝜼)|2+(λe(Δt)2+λv(Δt))Ωb|𝜼|2+c0(Δt)2Ωb|p|2+κ(Δt)3Ωb𝒥b(ηNn)δ|b(ηNn)δp|2.{{B[(\boldsymbol{u},\boldsymbol{\eta},p),(\boldsymbol{u},\boldsymbol{\eta},p)]}}:=(\Delta t)^{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}+\omega^{n+1}_{N}}{2R}\right)|\boldsymbol{u}|^{2}+2\nu(\Delta t)^{3}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\left|\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{u})\right|^{2}\\ +\frac{\beta}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}(\Delta t)\int_{\Gamma}\left|(\boldsymbol{\eta}-(\Delta t)\boldsymbol{u})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}\right|^{2}+\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\eta}|^{2}+\rho_{p}\int_{\Gamma}|\boldsymbol{\eta}|^{2}+(2\mu_{e}(\Delta t)^{2}+2\mu_{v}(\Delta t))\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta})|^{2}\\ +(\lambda_{e}(\Delta t)^{2}+\lambda_{v}(\Delta t))\int_{\Omega_{b}}\left|\nabla\cdot\boldsymbol{\eta}\right|^{2}+c_{0}(\Delta t)^{2}\int_{\Omega_{b}}|p|^{2}+\kappa(\Delta t)^{3}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p|^{2}.

Coercivity of this form follows from the fact that |ωNk+i2|<R|\omega^{k+\frac{i}{2}}_{N}|<R, see Assumption 1A in (56), and Korn inequality, see Proposition 6.1, once we handle the last term and show that

κ(Δt)3Ωb𝒥b(ηNn)δ|b(ηNn)δp|2cΩb|p|2,\kappa(\Delta t)^{3}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p|^{2}\geq c\int_{\Omega_{b}}|\nabla p|^{2},

for some positive constant c>0c>0. To show this, we first recall the definitions

𝒥b(ηNn)δ=det(𝑰+(𝜼Nn)δ),b(ηNn)δp=p(𝑰+(𝜼Nn)δ)1.\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}=\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}),\qquad\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p=\nabla p\cdot(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})^{-1}.

Then, letting |||\cdot| denote the matrix norm, we have

κ(Δt)3Ωb𝒥b(ηNn)δ|b(ηNn)δp|2κ(Δt)3Ωb𝒥b(ηNn)δ|𝑰+(𝜼Nn)δ|2|p|2.\kappa(\Delta t)^{3}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p|^{2}\geq\kappa(\Delta t)^{3}\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}|^{-2}|\nabla p|^{2}. (60)

Assumption 2A (57) implies that 𝑰+(𝜼Nn)δ\boldsymbol{I}+(\boldsymbol{\eta}^{n}_{N})^{\delta} is an invertible map from Ωb\Omega_{b} to (Ωb)Nn,δ(\Omega_{b})^{n,\delta}_{N}, and we further note that |𝑰+(𝜼Nn)δ||\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}| is continuous on Ωb¯\overline{\Omega_{b}} and hence is bounded from above. Thus, |𝑰+(𝜼Nn)δ|2c0>0|\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}|^{-2}\geq c_{0}>0 for some positive constant c0c_{0}. The assumption that 𝑰+(𝜼Nn)δ\boldsymbol{I}+(\boldsymbol{\eta}^{n}_{N})^{\delta} is invertible implies that det(𝑰+(𝜼Nn)δ)>0\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})>0. However, since this determinant is a continuous function on the compact set Ωb¯\overline{\Omega_{b}}, we conclude that there exists a positive constant c1>0c_{1}>0 such that det(𝑰+(𝜼Nn)δ)c1>0\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})\geq c_{1}>0. This establishes coercivity.

Existence of a unique weak solution (𝒖Nn+1,𝜼Nn+1,pNn+1)𝒱fωNn×Vd×Vp(\boldsymbol{u}^{n+1}_{N},\boldsymbol{\eta}^{n+1}_{N},p^{n+1}_{N})\in\mathcal{V}^{\omega^{n}_{N}}_{f}\times V_{d}\times V_{p} now follows from the Lax-Milgram lemma. From here, we recover ζNn+1\zeta^{n+1}_{N}, by using ζNn+1𝒆y=𝜼Nn+1𝜼NnΔt|Γ\displaystyle\zeta^{n+1}_{N}\boldsymbol{e}_{y}=\frac{\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N}}{\Delta t}\Big{|}_{\Gamma}. Note that 𝜼Nn+1𝜼NnΔt|Γ\displaystyle\frac{\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N}}{\Delta t}\Big{|}_{\Gamma} points in the yy direction because the trace of any function 𝜼Vd\boldsymbol{\eta}\in V_{d} on Γ\Gamma points in the yy direction by definition, see (32).

Energy equality: We substitute 𝒗=𝒖Nn+1\boldsymbol{v}=\boldsymbol{u}^{n+1}_{N}, φ=ζNn+1\varphi=\zeta^{n+1}_{N}, 𝝍=𝜼˙Nn+1\boldsymbol{\psi}=\boldsymbol{\dot{\eta}}^{n+1}_{N}, and r=pNn+1r=p^{n+1}_{N} into (54), and use the identity

(ab)a=12(|a|2+|ab|2|b|2).(a-b)\cdot a=\frac{1}{2}(|a|^{2}+|a-b|^{2}-|b|^{2}).

Since ωNn+1=ωNn+12\omega^{n+1}_{N}=\omega^{n+\frac{1}{2}}_{N} and (Δt)ζNn+12=ωNn+12ωNn(\Delta t)\zeta^{n+\frac{1}{2}}_{N}=\omega^{n+\frac{1}{2}}_{N}-\omega^{n}_{N}, we obtain the following energy equality:

12Ωf(1+ωNn+1R)|𝒖Nn+1|2+12ρbΩb|𝜼˙Nn+1|2+12c0Ωb|pNn+1|2+μeΩb|𝑫(𝜼Nn+1)|2+12λeΩb|𝜼Nn+1|2\displaystyle\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n+1}_{N}}{R}\right)|\boldsymbol{u}^{n+1}_{N}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\dot{\boldsymbol{\eta}}^{n+1}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p^{n+1}_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N})|^{2}+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}^{n+1}_{N}|^{2}
+12ρpΓ|ζNn+1|2+2μv(Δt)Ωb|𝑫(𝜼˙Nn+1)|2+λv(Δt)Ωb|𝜼˙Nn+1|2+κ(Δt)Ωb𝒥b(ηNn)δ|b(ηNn)δpNn+1|2\displaystyle+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+1}_{N}|^{2}+2\mu_{v}(\Delta t)\int_{\Omega_{b}}|\boldsymbol{D}(\dot{\boldsymbol{\eta}}^{n+1}_{N})|^{2}+\lambda_{v}(\Delta t)\int_{\Omega_{b}}|\nabla\cdot\dot{\boldsymbol{\eta}}^{n+1}_{N}|^{2}+\kappa(\Delta t)\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}|^{2}
+β(Δt)𝒥ΓωNnΓ|(𝜼˙Nn+1𝒖Nn+1)𝝉ωNn|2+12ρbΩb|𝜼˙Nn+1𝜼˙Nn|2+12c0Ωb|pNn+1pNn|2+μeΩb|𝑫(𝜼Nn+1𝜼Nn)|2\displaystyle+\frac{\beta(\Delta t)}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}\int_{\Gamma}|(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p^{n+1}_{N}-p^{n}_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N})|^{2}
+12λeΩb|(𝜼Nn+1𝜼Nn)|2=12Ωf(1+ωNnR)|𝒖Nn|2+12ρbΩb|𝜼˙Nn|2+12c0Ωb|pNn|2+μeΩb|𝑫(𝜼Nn)|2\displaystyle+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot(\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N})|^{2}=\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)|\boldsymbol{u}^{n}_{N}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\dot{\boldsymbol{\eta}}^{n}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p^{n}_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}^{n}_{N})|^{2}
+12λeΩb|𝜼Nn|2+12ρpΓ|ζNn+12|2,\displaystyle+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}^{n}_{N}|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+\frac{1}{2}}_{N}|^{2},

where the terms containing parameter α\alpha cancel out after bringing the integrals back to the time-dependent domain, integrating by parts, and recalling that the normal vector points inward towards the Biot domain:

αΩb𝒥b(ηNn)δpNn+1b(ηNn)δ𝜼˙Nn+1αΩb𝒥b(ηNn)δ𝜼˙Nn+1b(ηNn)δpNn+1αΓ(𝜼˙Nn+1𝒏(ωNn)δ)pNn+1=α(Ωb)Nn,δpNn+1(𝜼˙Nn+1)α(Ωb)Nn,δ𝜼˙Nn+1pNn+1αΓNn,δ(𝜼˙Nn+1𝒏)pNn+1=0.-\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\dot{\eta}}^{n+1}_{N}-\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}-\alpha\int_{\Gamma}\left(\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}}\right)p^{n+1}_{N}\\ =-\alpha\int_{(\Omega_{b})^{n,\delta}_{N}}p^{n+1}_{N}(\nabla\cdot\boldsymbol{\dot{\eta}}^{n+1}_{N})-\alpha\int_{(\Omega_{b})^{n,\delta}_{N}}\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\nabla p^{n+1}_{N}-\alpha\int_{\Gamma^{n,\delta}_{N}}(\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\boldsymbol{n})p^{n+1}_{N}=0.

This completes the proof of the Lemma. ∎

6.3 The coupled semi-discrete problem: weak formulation and energy

To obtain uniform energy estimates for approximate solutions of our semidiscretized scheme it is useful to present the scheme in monolithic form:

Ωf(1+ωNnR)𝒖˙Nn+1𝒗+2νΩf(1+ωNnR)𝑫fωNn(𝒖Nn+1):𝑫fωNn(𝒗)+Γ(12𝒖Nn+1𝒖NnpNn+1)(𝝍𝒗)𝒏ωNn\displaystyle\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\boldsymbol{\dot{u}}^{n+1}_{N}\cdot\boldsymbol{v}+2\nu\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{u}^{n+1}_{N}):\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{v})+\int_{\Gamma}\left(\frac{1}{2}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{u}^{n}_{N}-p^{n+1}_{N}\right)(\boldsymbol{\psi}-\boldsymbol{v})\cdot\boldsymbol{n}^{\omega^{n}_{N}} (61)
+12Ωf(1+ωNnR)[((𝒖NnζNn+12R+yR𝒆y)fωNn𝒖Nn+1)𝒗((𝒖NnζNn+12R+yR𝒆y)fωNn𝒗)𝒖Nn+1]\displaystyle+\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\left[\left(\left(\boldsymbol{u}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\omega^{n}_{N}}_{f}\boldsymbol{u}^{n+1}_{N}\right)\cdot\boldsymbol{v}-\left(\left(\boldsymbol{u}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\omega^{n}_{N}}_{f}\boldsymbol{v}\right)\cdot\boldsymbol{u}^{n+1}_{N}\right]
+12RΩfζNn+12𝒖Nn+1𝒗+12Γ(𝒖Nn+1𝜼˙Nn+1)𝒏ωNn(𝒖Nn𝒗)+β𝒥ΓωNnΓ(𝜼˙Nn+1𝒖Nn+1)𝝉ωNn(𝝍𝒗)𝝉ωNn\displaystyle+\frac{1}{2R}\int_{\Omega_{f}}\zeta^{n+\frac{1}{2}}_{N}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{v}+\frac{1}{2}\int_{\Gamma}(\boldsymbol{u}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}^{\omega^{n}_{N}}(\boldsymbol{u}^{n}_{N}\cdot\boldsymbol{v})+\frac{\beta}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}\int_{\Gamma}(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}(\boldsymbol{\psi}-\boldsymbol{v})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}
+ρbΩb(𝜼˙Nn+1𝜼˙NnΔt)𝝍+ρpΓ(ζNn+1ζNnΔt)φ+2μeΩb𝑫(𝜼Nn+1):𝑫(𝝍)+λeΩb(𝜼Nn+1)(𝝍)\displaystyle+\rho_{b}\int_{\Omega_{b}}\left(\frac{\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n}_{N}}{\Delta t}\right)\cdot\boldsymbol{\psi}+\rho_{p}\int_{\Gamma}\left(\frac{\zeta^{n+1}_{N}-\zeta^{n}_{N}}{\Delta t}\right)\varphi+2\mu_{e}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{e}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}^{n+1}_{N})(\nabla\cdot\boldsymbol{\psi})
+2μvΩb𝑫(𝜼˙Nn+1):𝑫(𝝍)+λvΩb(𝜼˙Nn+1)(𝝍)αΩb𝒥b(ηNn)δpNn+1b(ηNn)δ𝝍+c0ΩbpNn+1pNnΔtr\displaystyle+2\mu_{v}\int_{\Omega_{b}}\boldsymbol{D}(\dot{\boldsymbol{\eta}}^{n+1}_{N}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{v}\int_{\Omega_{b}}(\nabla\cdot\dot{\boldsymbol{\eta}}^{n+1}_{N})(\nabla\cdot\boldsymbol{\psi})-\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\psi}+c_{0}\int_{\Omega_{b}}\frac{p^{n+1}_{N}-p^{n}_{N}}{\Delta t}r
αΩb𝒥b(ηNn)δ𝜼˙Nn+1b(ηNn)δrαΓ(𝜼˙Nn+1𝒏(ωNn)δ)r+κΩb𝒥b(ηNn)δb(ηNn)δpNn+1b(ηNn)δr\displaystyle-\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r-\alpha\int_{\Gamma}(\boldsymbol{\dot{\eta}}^{n+1}_{N}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}})r+\kappa\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r
Γ[(𝒖Nn+1𝜼˙Nn+1)𝒏ωNn]r+ΓΔωNn+12Δφ=0,(𝒗,φ,𝝍,r)𝒬Nn+1,\displaystyle-\int_{\Gamma}[(\boldsymbol{u}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}^{\omega^{n}_{N}}]r+\int_{\Gamma}\Delta\omega^{n+\frac{1}{2}}_{N}\cdot\Delta\varphi=0,\ \forall(\boldsymbol{v},\varphi,\boldsymbol{\psi},r)\in\mathcal{Q}^{n+1}_{N},
Γ(ωNn+12ωNn12Δt)ϕ=ΓζNn+12ϕ,Γ(𝜼Nn+1𝜼NnΔt)ϕ=ΓζNn+1𝒆yϕ,ϕ,ϕL2(Γ).\int_{\Gamma}\left(\frac{\omega^{n+\frac{1}{2}}_{N}-\omega^{n-\frac{1}{2}}_{N}}{\Delta t}\right)\phi=\int_{\Gamma}\zeta^{n+\frac{1}{2}}_{N}\phi,\qquad\int_{\Gamma}\left(\frac{\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N}}{\Delta t}\right)\cdot\boldsymbol{\phi}=\int_{\Gamma}\zeta^{n+1}_{N}\boldsymbol{e}_{y}\cdot\boldsymbol{\phi},\ \forall\phi,\boldsymbol{\phi}\in L^{2}(\Gamma). (62)

Next, we will obtain uniform energy estimates for the approximate solutions generated from the splitting scheme. To do this, we define the discrete energy ENn+i2E^{n+\frac{i}{2}}_{N} and discrete dissipation DNn+1D^{n+1}_{N} as follows:

ENn+i2\displaystyle E^{n+\frac{i}{2}}_{N} =12Ωf(1+ωNnR)|𝒖Nn+i2|2+12ρbΩb|𝜼˙Nn+i2|2+12c0Ωb|pNn+i2|2+μeΩb|𝑫(𝜼Nn+i2)|2,.\displaystyle=\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)|\boldsymbol{u}^{n+\frac{i}{2}}_{N}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\dot{\eta}}^{n+\frac{i}{2}}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p^{n+\frac{i}{2}}_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}^{n+\frac{i}{2}}_{N})|^{2},. (63)
+12λeΩb|𝜼Nn+i2|2+12ρpΓ|ζNn+i2|2+12Γ|ΔωNn+i2|2,i=0,1.\displaystyle+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}^{n+\frac{i}{2}}_{N}|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+\frac{i}{2}}_{N}|^{2}+\frac{1}{2}\int_{\Gamma}|\Delta\omega^{n+\frac{i}{2}}_{N}|^{2},\ i=0,1.
DNn+1\displaystyle D^{n+1}_{N} =2ν(Δt)Ωf(1+ωNnR)|𝑫fωNn(𝒖Nn+1)|2+2μv(Δt)Ωb|𝑫(𝜼˙Nn+1)|2+λv(Δt)Ωb|𝜼˙Nn+1|2\displaystyle=2\nu(\Delta t)\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\left|\boldsymbol{D}^{\omega^{n}_{N}}_{f}(\boldsymbol{u}^{n+1}_{N})\right|^{2}+2\mu_{v}(\Delta t)\int_{\Omega_{b}}|\boldsymbol{D}(\dot{\boldsymbol{\eta}}^{n+1}_{N})|^{2}+\lambda_{v}(\Delta t)\int_{\Omega_{b}}|\nabla\cdot\dot{\boldsymbol{\eta}}^{n+1}_{N}|^{2}
+κ(Δt)Ωb𝒥b(ηNn)δ|b(ηNn)δpNn+1|2+β(Δt)𝒥ΓωNnΓ|(𝜼˙Nn+1𝒖Nn+1)𝝉ωNn|2.\displaystyle+\kappa(\Delta t)\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}|^{2}+\frac{\beta(\Delta t)}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}\int_{\Gamma}\left|(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}\right|^{2}.

Then, the semidiscrete weak formulation (61) and (62) implies the following uniform estimates on the discretized energy and dissipation.

Lemma 6.3.

The following discrete energy equalities hold for the semi-discretized formulation (61), (62):

ENn+12+12ρpΓ|ζNn+12ζNn|2+12Γ|Δ(ωNn+12ωNn12)|2=ENnE^{n+\frac{1}{2}}_{N}+\frac{1}{2}\rho_{p}\int_{\Gamma}\left|\zeta^{n+\frac{1}{2}}_{N}-\zeta^{n}_{N}\right|^{2}+\frac{1}{2}\int_{\Gamma}\left|\Delta(\omega^{n+\frac{1}{2}}_{N}-\omega^{n-\frac{1}{2}}_{N})\right|^{2}=E^{n}_{N} (64)
ENn+1+DNn+1+12Ωf(1+ωNnR)|𝒖Nn+1𝒖Nn|2+12ρbΩb|𝜼˙Nn+1𝜼˙Nn|2+12c0Ωb|pNn+1pNn|2+μeΩb|𝑫(𝜼Nn+1𝜼Nn)|2+12λeΩb|(𝜼Nn+1𝜼Nn)|2+12ρpΓ|ζNn+1ζNn+12|2=ENn+12,E^{n+1}_{N}+D^{n+1}_{N}+\frac{1}{2}\int_{\Omega_{f}}\left(1+\frac{\omega^{n}_{N}}{R}\right)\left|\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}^{n}_{N}\right|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}\left|\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n}_{N}\right|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}\left|p^{n+1}_{N}-p^{n}_{N}\right|^{2}\\ +\mu_{e}\int_{\Omega_{b}}\left|\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N})\right|^{2}+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}\left|\nabla\cdot(\boldsymbol{\eta}^{n+1}_{N}-\boldsymbol{\eta}^{n}_{N})\right|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta^{n+1}_{N}-\zeta^{n+\frac{1}{2}}_{N}|^{2}=E^{n+\frac{1}{2}}_{N}, (65)

where the discrete energy ENn+i2E^{n+\frac{i}{2}}_{N} and the discrete dissipation DNn+1D^{n+1}_{N} are defined in (63).

We remark that the terms not included in the definition of ENn+i2E^{n+\frac{i}{2}}_{N} and DNn+1D^{n+1}_{N}, appearing in (64) and (65), are numerical dissipation terms.

These energy identities immediately imply that ENn+i2E^{n+\frac{i}{2}}_{N} and n=1NDNn\sum_{n=1}^{N}D^{n}_{N} are uniformly bounded by a constant CC independent of nn and NN.

The semidiscretized splitting scheme defines semidiscretized approximations of the solution to the regularized problem at discrete time points. To work with approximate functions and show that they converge to the solution of the continuous problem, we need to extend the semidiscrete approximations to the entire time interval and investigate uniform boundedness of those approximate solution functions. This is done next.

7 Approximate solutions

Now that we have defined the numerical solutions at each time step, we collect the solutions into approximate solutions defined on the whole time interval [0,T][0,T], for which we will obtain uniform estimates from our previous energy estimates.

We define the following two extensions of the approximate functions to the entire interval [0,T][0,T]:

  • Piecewise constant approximate solutions, for (n1)Δt<tnΔt(n-1)\Delta t<t\leq n\Delta t:

    𝒖N(t)=𝒖Nn,𝜼N(t)=𝜼Nn,pN(t)=pNn,ωN(t)=ωNn12,ζN(t)=ζNn12,ζN(t)=ζNn;\boldsymbol{u}_{N}(t)=\boldsymbol{u}^{n}_{N},\quad\boldsymbol{\eta}_{N}(t)=\boldsymbol{\eta}^{n}_{N},\quad p_{N}(t)=p^{n}_{N},\quad\omega_{N}(t)=\omega^{n-\frac{1}{2}}_{N},\quad\zeta_{N}(t)=\zeta^{n-\frac{1}{2}}_{N},\quad\zeta_{N}^{*}(t)=\zeta^{n}_{N}; (66)
  • Linear interpolations:

    𝜼¯N(nΔt)=𝜼Nn,p¯N(nΔt)=pNn,ω¯N(nΔt)=ωNn12, for n=0,1,,N,\boldsymbol{\overline{\eta}}_{N}(n\Delta t)=\boldsymbol{\eta}^{n}_{N},\qquad\overline{p}_{N}(n\Delta t)=p^{n}_{N},\qquad\overline{\omega}_{N}(n\Delta t)=\omega^{n-\frac{1}{2}}_{N},\qquad\text{ for }n=0,1,...,N, (67)

    where we formally set ωN12=ω0\omega^{-\frac{1}{2}}_{N}=\omega_{0}.

Note that by construction, we have that

tω¯N=ζN,t𝜼¯N|Γ=ζN𝒆y.\partial_{t}\overline{\omega}_{N}=\zeta_{N},\qquad\partial_{t}\boldsymbol{\overline{\eta}}_{N}|_{\Gamma}=\zeta^{*}_{N}\boldsymbol{e}_{y}. (68)

From the preceding energy estimates, we have the following lemma on uniform boundedness.

Lemma 7.1.

Uniform boundedness of approximate solutions. Assume:

  1. 1.

    Assumption 1B: Uniform boundedness of plate displacements. There exists a positive constant RmaxR_{max} such that for all NN,

    |ωNn12|Rmax<R, for all n=0,1,,N,|\omega^{n-\frac{1}{2}}_{N}|\leq R_{max}<R,\qquad\text{ for all }n=0,1,...,N, (69)
    |(𝜼Nn)δ|Γ|Rmax<R, for all n=0,1,,N.\left|(\boldsymbol{\eta}^{n}_{N})^{\delta}|_{\Gamma}\right|\leq R_{max}<R,\qquad\text{ for all }n=0,1,...,N. (70)
  2. 2.

    Assumption 2B: Uniform invertibility of the Lagrangian map (Jacobian). There exists a positive constant c0c_{0} such that for all NN,

    det(𝑰+(𝜼Nn)δ)c0>0, for all n=0,1,,N.\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})\geq c_{0}>0,\qquad\text{ for all }n=0,1,...,N. (71)
  3. 3.

    Assumption 2C: Uniform boundedness of the Lagrangian map (matrix norm). There exists positive constants c1c_{1} and c2c_{2} such that for all NN,

    |(𝑰+(𝜼Nn)δ)1|c1,|𝑰+(𝜼Nn)δ|c2, for all n=0,1,,N.|(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})^{-1}|\leq c_{1},\qquad|\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}|\leq c_{2},\qquad\text{ for all }n=0,1,...,N. (72)

Then for all NN:

  • 𝒖N\boldsymbol{u}_{N} is uniformly bounded in L(0,T;L2(Ωf))L^{\infty}(0,T;L^{2}(\Omega_{f})) and L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})).

  • 𝜼N\boldsymbol{\eta}_{N} is uniformly bounded in L(0,T;H1(Ωb))L^{\infty}(0,T;H^{1}(\Omega_{b})).

  • pNp_{N} is uniformly bounded in L(0,T;L2(Ωb))L^{\infty}(0,T;L^{2}(\Omega_{b})) and L2(0,T;H1(Ωb))L^{2}(0,T;H^{1}(\Omega_{b})).

  • ωN\omega_{N} is uniformly bounded in L(0,T;H02(Γ))L^{\infty}(0,T;H_{0}^{2}(\Gamma)).

In addition, we have the following estimates on the linear interpolations.

  • 𝜼¯N\boldsymbol{\overline{\eta}}_{N} is uniformly bounded in W1,(0,T;L2(Ωb))W^{1,\infty}(0,T;L^{2}(\Omega_{b})).

  • 𝝎¯N\boldsymbol{\overline{\omega}}_{N} is uniformly bounded in W1,(0,T;L2(Γ))W^{1,\infty}(0,T;L^{2}(\Gamma)).

Remark 7.1.

A crucial remark about invertibility. At first, it would appear that to show the uniform boundedness results above, we also need to have a fourth assumption, which is Assumption 2A (57) from before, that the map Id+(𝜼Nn)δ:Ωb2\text{Id}+(\boldsymbol{\eta}^{n}_{N})^{\delta}:\Omega_{b}\to\mathbb{R}^{2} is injective (and is hence a bijection onto its image), for each n=0,1,,Nn=0,1,...,N and for all NN. However, this is implied by an injectivity theorem, see Ciarlet [25] Theorem 5-5-2. Note also that Assumption 1A (56) from before is automatically satisfied once we verify Assumption 1B (69), (70). In particular, this injectivity theorem is as follows. Since det(𝑰+(𝜼Nn)δ)>0\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})>0 by Assumption 2B (71), it suffices to show that Id+(𝜼Nn)δ=𝝋0\text{Id}+(\boldsymbol{\eta}^{n}_{N})^{\delta}=\boldsymbol{\varphi}_{0} on Ωb\partial\Omega_{b}, for some injective mapping 𝝋0:Ωb¯2\boldsymbol{\varphi}_{0}:\overline{\Omega_{b}}\to\mathbb{R}^{2}, for example a standard ALE mapping 𝝋0(x,y)=(x,y+(1yR)ω)\boldsymbol{\varphi}_{0}(x,y)=\left(x,y+\left(1-\frac{y}{R}\right)\omega\right) can be used. This implies the very useful fact that (Id+(𝜼Nn)δ)(Ωb¯)=𝝋0(Ωb¯)(\text{Id}+(\boldsymbol{\eta}^{n}_{N})^{\delta})(\overline{\Omega_{b}})=\boldsymbol{\varphi}_{0}(\overline{\Omega_{b}}), which means that the deformed configuration is fully determined by the behavior on the boundary.

Proof.

The uniform boundedness of approximate solutions follows from the uniform energy estimates. More precisely, the uniform boundedness of 𝒖N\boldsymbol{u}_{N} in L(0,T;L2(Ωf))L^{\infty}(0,T;L^{2}(\Omega_{f})) follows from Assumption 1B (69). The uniform boundedness of 𝒖N\boldsymbol{u}_{N} in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})) follows from Korn’s inequality on the fluid domain. The uniform boundedness of 𝜼N\boldsymbol{\eta}_{N} in L(0,T;H1(Ωb))L^{\infty}(0,T;H^{1}(\Omega_{b})) follows from combining the uniform energy estimates with Korn’s inequality, stated in Proposition 6.1. To establish the uniform boundedness of pNp_{N} in L2(0,T;H1(Ωb))L^{2}(0,T;H^{1}(\Omega_{b})), we note that the discrete energy estimates in Lemma 6.3 imply the following uniform discrete dissipation bound:

n=0N1κ(Δt)Ωb𝒥b(ηNn)δ|b(ηNn)δpNn+1|2C,{{\sum_{n=0}^{N-1}}}\kappa(\Delta t)\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}|^{2}\leq C,

for some constant CC uniform in NN, where 𝒥b(ηNn)δ=det(𝑰+(𝜼Nn)δ),\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}=\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}), and b(ηNn)δr=r(𝑰+(𝜼Nn)δ)1 on Ωb.\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r=\nabla r\cdot(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})^{-1}\ \text{ on }\Omega_{b}. By Assumption 2B (71), we conclude that

(Δt)n=0N1Ωb|b(ηNn)δpNn+1|2C.(\Delta t){{\sum_{n=0}^{N-1}}}\int_{\Omega_{b}}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}|^{2}\leq C.

Since on Ωb\Omega_{b}, we have that pNn+1=b(ηNn)δpNn+1(𝑰+(𝜼Nn)δ)\nabla p^{n+1}_{N}=\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\cdot(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}), we use Assumption 2C (72), which implies |𝑰+(𝜼Nn)δ|c2|\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}|\leq c_{2}, and obtain the estimate

(Δt)n=0N1Ωb|pNn+1|2|𝑰+(𝜼Nn)δ|2(Δt)n=0N1Ωb|b(ηNn)δpNn+1|2C,(\Delta t){{\sum_{n=0}^{N-1}}}\int_{\Omega_{b}}|\nabla p^{n+1}_{N}|^{2}\leq|\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}|^{2}\cdot(\Delta t){{\sum_{n=0}^{N-1}}}\int_{\Omega_{b}}|\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}|^{2}\leq C,

for a constant CC independent of NN. Thus, pNp_{N} is uniformly bounded in L2(0,T;H1(Ωb))L^{2}(0,T;H^{1}(\Omega_{b})), since by the definition of the piecewise constant approximate solution pNp_{N} in (66), we have that

pNL2(0,T;H1(Ωb))2=(Δt)n=0N1Ωb|pNn+1|2.||p_{N}||^{2}_{L^{2}(0,T;H^{1}(\Omega_{b}))}=(\Delta t)\sum_{n=0}^{N-1}\int_{\Omega_{b}}|\nabla p_{N}^{n+1}|^{2}.

The above uniform boundedness result implies the following weak convergence results.

Proposition 7.1.

Assume that the three assumptions listed in Lemma 7.1 hold. Then, there exists a subsequence such that the following weak convergence results hold:

  • 𝒖N𝒖\boldsymbol{u}_{N}\rightharpoonup\boldsymbol{u} weakly* in L(0,T;L2(Ωf))L^{\infty}(0,T;L^{2}(\Omega_{f})),   𝒖N𝒖\boldsymbol{u}_{N}\rightharpoonup\boldsymbol{u} weakly in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})),

  • 𝜼N𝜼\boldsymbol{\eta}_{N}\rightharpoonup\boldsymbol{\eta} weakly* in L(0,T;H1(Ωb))L^{\infty}(0,T;H^{1}(\Omega_{b})),   𝜼¯N𝜼¯\overline{\boldsymbol{\eta}}_{N}\rightharpoonup\overline{\boldsymbol{\eta}} weakly* in W1,(0,T;L2(Ωb))W^{1,\infty}(0,T;L^{2}(\Omega_{b})),

  • pNpp_{N}\rightharpoonup p weakly* in L(0,T;L2(Ωb))L^{\infty}(0,T;L^{2}(\Omega_{b})),   pNpp_{N}\rightharpoonup p weakly in L2(0,T;H1(Ωb))L^{2}(0,T;H^{1}(\Omega_{b})),

  • ωNω\omega_{N}\rightharpoonup\omega weakly* in L(0,T;H02(Γ))L^{\infty}(0,T;H_{0}^{2}(\Gamma)),   ω¯Nω¯\overline{\omega}_{N}\rightharpoonup\overline{\omega} weakly* in W1,(0,T;L2(Γ))W^{1,\infty}(0,T;L^{2}(\Gamma)).

Furthermore, 𝜼=𝜼¯\boldsymbol{\eta}=\boldsymbol{\overline{\eta}} and ω=ω¯\omega=\overline{\omega}.

To use these results and to be able to construct approximate solutions, it is essential to show that the assumptions from Lemma 7.1 hold. This is given by the following lemma.

Lemma 7.2.

Suppose that the initial data satisfies |ω0|R0<R|\omega_{0}|\leq R_{0}<R for some R0R_{0}, and suppose that 𝜼0\boldsymbol{\eta}_{0} has the property that Id+(𝜼0)δ\text{Id}+(\boldsymbol{\eta}_{0})^{\delta} is invertible with det(𝑰+(𝜼0)δ)c0>0\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}_{0})^{\delta})\geq c_{0}>0 on Ωb\Omega_{b} for some positive constant c0c_{0}. Then, there exists a sufficiently small time T>0T>0 such that for all NN, all three assumptions in Lemma 7.1 hold and the splitting scheme is well defined until time TT.

Proof.

First, notice that the assumptions on the initial data immediately imply that the three assumptions from Lemma 7.1 hold for the initial data, i.e., for n=0n=0. In particular, there exist constants α0\alpha_{0}, α1\alpha_{1}, and α2\alpha_{2} such that

det(𝑰+(𝜼0)δ)α0>0,\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}_{0})^{\delta})\geq\alpha_{0}>0, (73)
|𝑰+(𝜼0)δ|α1>0,|(𝑰+(𝜼0)δ)1|α2>0.|\boldsymbol{I}+\nabla(\boldsymbol{\eta}_{0})^{\delta}|\geq\alpha_{1}>0,\qquad|(\boldsymbol{I}+\nabla(\boldsymbol{\eta}_{0})^{\delta})^{-1}|\geq\alpha_{2}>0. (74)

This is because det(𝑰+(𝜼0)δ)\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}_{0})^{\delta}), |𝑰+(𝜼0)δ||\boldsymbol{I}+\nabla(\boldsymbol{\eta}_{0})^{\delta}|, and |(𝑰+(𝜼0)δ)1||(\boldsymbol{I}+\nabla(\boldsymbol{\eta}_{0})^{\delta})^{-1}| are positive continuous functions on the compact set Ωb¯\overline{\Omega_{b}}.

Next, we want to define an appropriate time T>0T>0 such that the three assumptions hold uniformly for all NN and nΔtn\Delta t up to time TT. To do this, we use the energy estimates. Define the initial energy determined by the initial data by E0E_{0}. Then, by the uniform energy estimates, we have that

ENk+12E0,ENk+1E0, for all k=0,1,,N1.E^{k+\frac{1}{2}}_{N}\leq E_{0},\qquad E^{k+1}_{N}\leq E_{0},\qquad\text{ for all }k=0,1,...,N-1.

Therefore, after completing both subproblems of the scheme on the time step [kΔt,(k+1)Δt][k\Delta t,(k+1)\Delta t], we obtain that

𝜼˙NnL2(Ωb)C, for n=0,1,,k+1,||\boldsymbol{\dot{\eta}}^{n}_{N}||_{L^{2}(\Omega_{b})}\leq C,\qquad\text{ for }n=0,1,...,k+1, (75)
ωNn+12H02(Γ)C, for n=0,1,,k,||\omega^{n+\frac{1}{2}}_{N}||_{H_{0}^{2}(\Gamma)}\leq C,\qquad\text{ for }n=0,1,...,k, (76)
ζNn+i2L2(Γ)C, for 0n+i2k+1 and i=0,1,||\zeta^{n+\frac{i}{2}}_{N}||_{L^{2}(\Gamma)}\leq C,\qquad\text{ for }0\leq n+\frac{i}{2}\leq k+1\qquad\text{ and }i=0,1, (77)

for a constant CC depending only on the initial energy E0E_{0}.

Step 1 (Uniform bound on the plate displacements ωNn12\omega^{n-\frac{1}{2}}_{N}). We first find a condition on TT such that Assumption 1B (69) is satisfied. Suppose that the linear interpolation ω¯N\overline{\omega}_{N} is defined up to time (k+1)Δt(k+1)\Delta t, where we recall that the linear interpolation is defined via (67). Then, by (76) and (77) and the fact that tω¯N=ζN\partial_{t}\overline{\omega}_{N}=\zeta_{N} from (68), we have

ω¯NW1,(0,(k+1)Δt;L2(Γ))C,||\overline{\omega}_{N}||_{W^{1,\infty}(0,(k+1)\Delta t;L^{2}(\Gamma))}\leq C, (78)
ω¯NL(0,(k+1)Δt;H02(Γ))C,||\overline{\omega}_{N}||_{L^{\infty}(0,(k+1)\Delta t;H_{0}^{2}(\Gamma))}\leq C, (79)

where CC depends only on E0E_{0} and is independent of NN. Thus, following the method in [52] (see in particular equation (73) in [52]), we obtain by an interpolation inequality that for all t,t+τ[0,(k+1)Δt]t,t+\tau\in[0,(k+1)\Delta t] with τ>0\tau>0,

ω¯N(t+τ)ω¯N(t)H1(Γ)Cω¯N(t+τ)ω¯N(t)L2(Γ)1/2ω¯N(t+τ)ω¯N(t)H2(Γ)1/2.||\overline{\omega}_{N}(t+\tau)-\overline{\omega}_{N}(t)||_{H^{1}(\Gamma)}\leq C||\overline{\omega}_{N}(t+\tau)-\overline{\omega}_{N}(t)||^{1/2}_{L^{2}(\Gamma)}||\overline{\omega}_{N}(t+\tau)-\overline{\omega}_{N}(t)||^{1/2}_{H^{2}(\Gamma)}. (80)

Here, we used a Sobolev interpolation inequality, see for example Theorem 4.17 (pg. 79) of [1]. By the Lipschitz continuity of ω¯N\overline{\omega}_{N} taking values in L2(Γ)L^{2}(\Gamma) given by (78) and by the boundedness of ω¯N\overline{\omega}_{N} in H02(Γ)H_{0}^{2}(\Gamma) given by (79),

ω¯N(t+τ)ω¯N(t)H1(Γ)Cτ1/2||\overline{\omega}_{N}(t+\tau)-\overline{\omega}_{N}(t)||_{H^{1}(\Gamma)}\leq C\cdot\tau^{1/2} (81)

for a constant CC depending only on E0E_{0} (and in particular, not depending on kk or NN). Therefore, setting t=0t=0 and τ=(k+1)Δt\tau=(k+1)\Delta t and using the continuous embedding of H1(Γ)H^{1}(\Gamma) into C(Γ)C(\Gamma) (see e.g. [1, Chapter V] for Sobolev embeddings),

ωNk+1ω0C(Γ)C[(k+1)Δt]1/2CT1/2,||\omega^{k+1}_{N}-\omega_{0}||_{C(\Gamma)}\leq C\cdot[(k+1){{\Delta t}}]^{1/2}\leq C\cdot T^{1/2}, (82)

where CC depends only E0E_{0}. Because |ω0|<R|\omega_{0}|<R, we can choose T>0T>0 sufficiently small so that

CT1/2<R||ω0||C(Γ).C\cdot T^{1/2}<R-||\omega_{0}||_{C(\Gamma)}. (83)

This will give the first part of Assumption 1B, which is (69).

Step 2 (Bound on the trace of the Biot displacements ηNn\boldsymbol{\eta}^{n}_{N} and the Lagrangian map). Next, we find a condition on TT so that the remaining assumptions (70), (71), and (72) are satisfied. We do this by controlling the behavior of the structure displacement 𝜼\boldsymbol{\eta}. First note that

𝜼Nk+1𝜼0L2(Ωb)(Δt)n=1k+1𝜼˙NnL2(Ωb)C(k+1)(Δt)CT,||\boldsymbol{\eta}^{k+1}_{N}-\boldsymbol{\eta}_{0}||_{L^{2}(\Omega_{b})}\leq(\Delta t)\sum_{n=1}^{k+1}||\boldsymbol{\dot{\eta}}^{n}_{N}||_{L^{2}(\Omega_{b})}\leq C(k+1)(\Delta t)\leq CT,

for CC depending only on E0E_{0}, where the first inequality follows from the triangle inequality and the definition of 𝜼˙Nn=ηNnηNn1Δt\displaystyle\dot{\boldsymbol{\eta}}^{n}_{N}=\frac{\eta^{n}_{N}-\eta^{n-1}_{N}}{\Delta t}, and the second inequality follows from (75). By the odd extension defined in Definition 5.1,

𝜼Nk+1𝜼0L2(Ω~b)C(𝜼Nk+1𝜼0L2(Ωb)+ωNk+1ω0L2(Γ))CT,||\boldsymbol{\eta}^{k+1}_{N}-\boldsymbol{\eta}_{0}||_{L^{2}(\tilde{\Omega}_{b})}\leq C\left(||\boldsymbol{\eta}^{k+1}_{N}-\boldsymbol{\eta}_{0}||_{L^{2}(\Omega_{b})}+||\omega^{k+1}_{N}-\omega_{0}||_{L^{2}(\Gamma)}\right)\leq CT,

for a constant CC depending only on E0E_{0}, where the estimate ωNk+1ω0L2(Γ)CT||\omega^{k+1}_{N}-\omega_{0}||_{L^{2}(\Gamma)}\leq CT follows from the Lipschitz estimate (78). By regularization, we then have that for a constant depending only on δ\delta and E0E_{0},

(𝜼Nk+1)δ(𝜼0)δH3(Ωb)C(δ,E0)T.||(\boldsymbol{\eta}_{N}^{k+1})^{\delta}-(\boldsymbol{\eta}_{0})^{\delta}||_{H^{3}(\Omega_{b})}\leq C(\delta,E_{0})\cdot T.

By using the trace theorem and the continuous embedding of H2(Γ)H^{2}(\Gamma) into C(Γ)C(\Gamma), we thus conclude that

||(𝜼Nk+1)δ|Γ(𝜼0)δ|Γ||C(Γ)C(δ,E0)T.||(\boldsymbol{\eta}_{N}^{k+1})^{\delta}|_{\Gamma}-(\boldsymbol{\eta}_{0})^{\delta}|_{\Gamma}||_{C(\Gamma)}\leq C(\delta,E_{0})\cdot T. (84)

Since H2(Ωb)H^{2}(\Omega_{b}) embeds continuously into C(Ωb)C(\Omega_{b}), we also have that

||(𝜼Nk+1)δ(𝜼0)δ||C(Ωb)C(δ,E0)T.||\nabla(\boldsymbol{\eta}_{N}^{k+1})^{\delta}-\nabla(\boldsymbol{\eta}_{0})^{\delta}||_{C(\Omega_{b})}\leq C(\delta,E_{0})\cdot T. (85)

Note that det(𝑰+𝑨)\det(\boldsymbol{I}+\boldsymbol{A}) is a continuous function of the entries of 𝑨\boldsymbol{A}. Also note that the matrix norms |𝑰+𝑨||\boldsymbol{I}+\boldsymbol{A}| and |(𝑰+𝑨)1||(\boldsymbol{I}+\boldsymbol{A})^{-1}| are continuous functions of the matrix 𝑨\boldsymbol{A}. Furthermore, we emphasize that the constant C(δ,E0)C(\delta,E_{0}) depends only on δ\delta and E0E_{0} and hence is independent of kk and NN. This dependence on δ\delta is allowable, since for this existence proof, δ\delta is an arbitrary but fixed regularization parameter.

Thus, there exists TT sufficiently small so that by (84) and (85), the remaining assumptions (70), (71), and (72) are satisfied, since these assumptions are all satisfied for the initial displacement 𝜼0\boldsymbol{\eta}_{0}. Furthermore, we can choose the constants c0c_{0}, c1c_{1}, c2c_{2}, and RmaxR_{max} (defined in the statement of those assumptions) independently of NN and n=0,1,,Nn=0,1,...,N, because of the fact that the constant C(δ,E0)C(\delta,E_{0}) in our estimates does not depend on kk (satisfying (k+1)ΔtT(k+1)\Delta t\leq T) or NN. ∎

8 Compactness arguments

We next want to pass to the limit in the semidiscrete formulation for the approximate solutions, stated in (61) and (62). Because this is a nonlinear problem with geometric nonlinearities, we must obtain stronger convergence than just weak and weak* convergence in Proposition 7.1, in order to pass to the limit. To do this, we will use compactness arguments of two types: the classical Aubin-Lions compactness theorem for functions defined on fixed domains, and generalized Aubin-Lions compactness arguments introduced in [57] for functions defined on moving domains, see also [52]. We will first deal with compactness arguments for the plate displacement and the Biot domain displacement. Then, we will deal with compactness arguments for the fluid velocity defined on moving domains.

8.1 Compactness for Biot poroelastic medium displacement

We show strong convergence of the Biot structure displacements 𝜼¯N\overline{\boldsymbol{\eta}}_{N} by using a standard Aubin-Lions compactness argument. In particular, we have the following strong convergence result for the Biot medium displacement:

Lemma 8.1.

The following compact embedding holds true W1,(0,T;L2(Ωb))L(0,T;H1(Ωb))C(0,T;L2(Ωb)),W^{1,\infty}(0,T;L^{2}(\Omega_{b}))\cap L^{\infty}(0,T;H^{1}(\Omega_{b}))\subset\subset C(0,T;L^{2}(\Omega_{b})), which implies the existence of a subsequence such that

𝜼¯N𝜼stronglyinC(0,T;L2(Ωb)).\boldsymbol{\overline{\eta}}_{N}\to\boldsymbol{\eta}\ {\rm{strongly\ in}}\ C(0,T;L^{2}(\Omega_{b})).
Proof.

The compact embedding above is a direct consequence of the standard Aubin-Lions compactness lemma in the case of p=p=\infty, which gives a stronger compact embedding into C(0,T;L2(Ωb))C(0,T;L^{2}(\Omega_{b})) rather than just L(0,T;L2(Ωb))L^{\infty}(0,T;L^{2}(\Omega_{b})). The fact that we can find a strongly convergent subsequence follows from this compact embedding, once we recall that {𝜼¯N}N=1\{\overline{\boldsymbol{\eta}}_{N}\}_{N=1}^{\infty} are uniformly bounded in the Banach space W1,(0,T;L2(Ωb))L(0,T;H1(Ωb))W^{1,\infty}(0,T;L^{2}(\Omega_{b}))\cap L^{\infty}(0,T;H^{1}(\Omega_{b})) by the uniform energy estimates. ∎

8.2 Compactness for the plate displacement

The uniform boundedness of the linear interpolation of the plate displacement ω¯N\overline{\omega}_{N} in W1,(0,T;L2(Γ))W^{1,\infty}(0,T;L^{2}(\Gamma)) and L(0,T;H02(Γ))L^{\infty}(0,T;H_{0}^{2}(\Gamma)) implies strong convergence of ω¯N\overline{\omega}_{N} in C(0,T;Hs(Γ))C(0,T;H^{s}(\Gamma)). Even though the plate displacements are uniformly bounded in L(0,T;H02(Γ))L^{\infty}(0,T;H_{0}^{2}(\Gamma)) we only get convergence in C(0,T;Hs(Γ))C(0,T;H^{s}(\Gamma)) for 0<s<20<s<2. This is because we will be using Arzela-Ascoli theorem and hence we will lose regularity due to the compact embedding of H2(Γ)H^{2}(\Gamma) into Hs(Γ)H^{s}(\Gamma) for 0<s<20<s<2. The precise statement of the compactness results for the approximate plate displacements is as follows:

Proposition 8.1.

Given arbitrary 0<s<20<s<2, there exists a subsequence such that the following strong convergences hold:

ω¯Nω, in C(0,T;Hs(Γ)),\overline{\omega}_{N}\to\omega,\qquad\text{ in }C(0,T;H^{s}(\Gamma)),
ωNω, in L(0,T;Hs(Γ)).\omega_{N}\to\omega,\qquad\text{ in }L^{\infty}(0,T;H^{s}(\Gamma)).
Proof.

Using the same argument as in Step 1 of the proof of Lemma 7.2, one can show the following uniform estimate for the linear interpolations ω¯N\overline{\omega}_{N} and τ>0\tau>0, t,t+τ[0,T]t,t+\tau\in[0,T]:

ω¯N(t+τ)ω¯N(t)H2α(Γ)ω¯N(t+τ)ω¯N(t)L2(Γ)1αω¯N(t+τ)ω¯N(t)H2(Γ)αCτ1α, for 0<α<1,\begin{array}[]{rl}||\overline{\omega}_{N}(t+\tau)-\overline{\omega}_{N}(t)||_{H^{2\alpha}(\Gamma)}&\leq||\overline{\omega}_{N}(t+\tau)-\overline{\omega}_{N}(t)||^{1-\alpha}_{L^{2}(\Gamma)}||\overline{\omega}_{N}(t+\tau)-\overline{\omega}_{N}(t)||^{\alpha}_{H^{2}(\Gamma)}\\ &\leq C\tau^{1-\alpha},\qquad\text{ for }0<\alpha<1,\end{array} (86)

where the constant CC is independent of NN, but can depend on the choice of α\alpha. The first inequality in (86) follows from an interpolation estimate for Sobolev spaces (see Theorem 4.17, pg. 79 of [1]) and as in the proof of estimate (80) from Lemma 7.2, the second inequality follows from the uniform Lipschitz estimate (78) and the uniform boundedness estimate (79). Because the constant CC in (86) is independent of NN, this estimate implies that for a given arbitrary α(0,1)\alpha\in(0,1), the functions ω¯N\overline{\omega}_{N} are uniformly bounded as functions in C0,1α(0,T;H2α(Γ))C^{0,1-\alpha}(0,T;H^{2\alpha}(\Gamma)). Hence, the strong convergence of ω¯N\overline{\omega}_{N} follows directly from the Arzela-Ascoli theorem and the fact that H2αH^{2\alpha} embeds compactly into any H2αϵH^{2\alpha-\epsilon} for ϵ>0\epsilon>0, once we choose α(0,1)\alpha\in(0,1) and ϵ>0\epsilon>0 appropriately so that 2αϵ=s2\alpha-\epsilon=s for a given arbitrary 0<s<20<s<2. Hence, we obtain the desired strong convergence, as the equicontinuity condition for the Arzela-Ascoli theorem follows from the above estimate.

To show a similar strong convergence result for ωN\omega_{N}, we must show that

ωN(t)ω¯N(t)L(0,T;Hs(Γ))0,||\omega_{N}(t)-\overline{\omega}_{N}(t)||_{L^{\infty}(0,T;H^{s}(\Gamma))}\to 0,

for arbitrary 0<s<20<s<2. Once we observe that ω¯N(nΔt)=ωN(t)\overline{\omega}_{N}(n\Delta t)=\omega_{N}(t) for nΔtt<(n+1)Δtn\Delta t\leq t<(n+1)\Delta t, this follows immediately from the above Hölder continuity estimate (86), as

ωN(t)ω¯N(t)L(0,T;Hs(Γ))C(Δt)1s20, as N.||\omega_{N}(t)-\overline{\omega}_{N}(t)||_{L^{\infty}(0,T;H^{s}(\Gamma))}\leq C(\Delta t)^{1-\frac{s}{2}}\to 0,\qquad\text{ as }N\to\infty.

Thus, ωN\omega_{N} and ω¯N\overline{\omega}_{N} have the same limit in L(0,T;Hs(Γ))L^{\infty}(0,T;H^{s}(\Gamma)) for 0<s<20<s<2.

Next, we will obtain compactness for the Biot velocity, plate velocity, pore pressure, and fluid velocity. Because the test space (53) has the pore pressure and fluid velocity decoupled from the Biot/plate velocity, we can handle the compactness argument for each of these quantities separately. In particular, we recall the definition of the discrete test space from (53) and note that we can decouple this test space into three smaller test spaces, one for the Biot/plate displacement/velocity, one for the pore pressure, and one for the fluid velocity. In the next section we show compactness results for the Biot velocity and plate velocity, which must be treated together since they are coupled by a kinematic coupling condition at the plate interface Γ\Gamma.

8.3 Compactness for the Biot velocity and plate velocity

Here, we will state and prove a compactness result for the Biot and plate velocities (𝝃N,ζN)(\boldsymbol{\xi}_{N},\zeta_{N}), by showing the existence of convergent subsequences that converge in L2(0,T;Hs(Ωb)×Hs(Γ))L^{2}(0,T;H^{-s}(\Omega_{b})\times H^{-s}(\Gamma)) for 1/2<s<0-1/2<s<0. We remark that we must consider negative spatial Sobolev spaces for the Biot/plate velocities for the following two reasons:

  • First, our existence result in Theorem 5.1 includes the purely elastic case in which the viscoelasticity coefficients μv,λv\mu_{v},\lambda_{v} are allowed to be zero. Hence, we can only expect the Biot velocities in the finite-energy spaces to have spatial regularity of at most L2(Ωb)L^{2}(\Omega_{b}).

  • Second, for the plate velocities ζN\zeta_{N}, we must consider negative spatial Sobolev spaces on Γ\Gamma since by the coupling conditions (11) and (12), it is not true that the plate velocities ζN\zeta_{N} are equal to the traces of the fluid velocities 𝒖NH1(Ωf)\boldsymbol{u}_{N}\in H^{1}(\Omega_{f}), which is typically the case in FSI with purely elastic structures and no-slip condition. Therefore, we do not get any higher regularity of the plate velocities than what we get from the finite energy spaces, which implies that the plate velocities ζN\zeta_{N} are only at most L2(Γ)L^{2}(\Gamma).

The main compactness result for the Biot/plate velocities is as follows:

Theorem 8.1.

For 1/2<s<0-1/2<s<0, there exists a subsequence such that

(𝝃N,ζN)(𝝃,ζ) strongly in L2(0,T;Hs(Ωb)×Hs(Γ)).(\boldsymbol{\xi}_{N},\zeta_{N})\to(\boldsymbol{\xi},\zeta)\ \text{ strongly in }L^{2}(0,T;H^{-s}(\Omega_{b})\times H^{-s}(\Gamma)).
Proof.

We will establish this result by using a compactness criterion for piecewise constant functions due to Dreher and Jüngel [31]. To simplify arguments, we define a slightly more regular Biot/plate velocity test space:

𝒬v={(𝝍,φ)(VdH2(Ωb))×H02(Γ):𝝍=φ𝒆y on Γ}.\mathcal{Q}_{v}=\{(\boldsymbol{\psi},\varphi)\in(V_{d}\cap H^{2}(\Omega_{b}))\times H_{0}^{2}(\Gamma):\boldsymbol{\psi}=\varphi\boldsymbol{e}_{y}\text{ on }\Gamma\}. (87)

We will use the following chain of embeddings

L2(Ωb)×L2(Γ)Hs(Ωb)×Hs(Γ)𝒬v,L^{2}(\Omega_{b})\times L^{2}(\Gamma)\subset\subset H^{-s}(\Omega_{b})\times H^{-s}(\Gamma)\subset\mathcal{Q}_{v}^{\prime},

where the first embedding is compact, as required for the Dreher-Jüngel compactness criterion [31].

Let τΔt\tau_{\Delta t} denote the time shift τΔtf(t,)=f(tΔt,)\tau_{\Delta t}f(t,\cdot)=f(t-\Delta t,\cdot) for a function ff defined on [0,T][0,T]. As required by the Dreher-Jüngel compactness criterion [31], to obtain compactness we must verify that the following inequality is satisfied for a uniform constant CC and for all Δt=T/N\Delta t=T/N:

τΔt(𝝃N,ζN)(𝝃N,ζN)ΔtL1(Δt,T;𝒬v)+(𝝃N,ζN)L(0,T;L2(Ωb)×L2(Γ))C.\left|\left|\frac{\tau_{\Delta t}(\boldsymbol{\xi}_{N},\zeta_{N})-(\boldsymbol{\xi}_{N},\zeta_{N})}{\Delta t}\right|\right|_{L^{1}({{\Delta t}},T;\mathcal{Q}_{v}^{\prime})}+||(\boldsymbol{\xi}_{N},\zeta_{N})||_{L^{\infty}(0,T;L^{2}(\Omega_{b})\times L^{2}(\Gamma))}\leq C. (88)

The second term in this inequality is uniformly bounded by Lemma 7.1, which gives exactly the uniform boundednenss of (𝝃N,ζN)(\boldsymbol{\xi}_{N},\zeta_{N}) in L(0,T;L2(Ωb)×L2(Γ))L^{\infty}(0,T;L^{2}(\Omega_{b})\times L^{2}(\Gamma)).

To deal with the first term in (88) we use the coupled semidiscrete formulation (61), (62) and set the test functions 𝒗\boldsymbol{v} and rr for the fluid velocity and Biot pore pressure to be zero because we are considering only the Biot and plate velocities. We obtain that for all test functions (𝝍,φ)𝒬v(\boldsymbol{\psi},\varphi)\in\mathcal{Q}_{v}, where 𝒬v\mathcal{Q}_{v} is defined in (87), the following holds:

ρbΩb(𝝃Nn+1𝝃NnΔt)𝝍+ρpΓ(ζNn+1ζNnΔt)φ\displaystyle\rho_{b}\int_{\Omega_{b}}\left(\frac{\boldsymbol{\xi}^{n+1}_{N}-\boldsymbol{\xi}^{n}_{N}}{\Delta t}\right)\cdot\boldsymbol{\psi}+\rho_{p}\int_{\Gamma}\left(\frac{\zeta^{n+1}_{N}-\zeta^{n}_{N}}{\Delta t}\right)\cdot\varphi
=Γ(12𝒖Nn+1𝒖NnpNn+1)(𝝍𝒏ωNn)Γβ𝒥ΓωNn(ζNn+1𝒆y𝒖Nn+1)𝝉ωNn(𝝍𝝉ωNn)\displaystyle=-\int_{\Gamma}\left(\frac{1}{2}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{u}^{n}_{N}-p^{n+1}_{N}\right)(\boldsymbol{\psi}\cdot\boldsymbol{n}^{\omega^{n}_{N}})-\int_{\Gamma}\frac{\beta}{\mathcal{J}^{\omega^{n}_{N}}_{\Gamma}}(\zeta^{n+1}_{N}\boldsymbol{e}_{y}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}^{\omega^{n}_{N}}(\boldsymbol{\psi}\cdot\boldsymbol{\tau}^{\omega^{n}_{N}})
2μeΩb𝑫(𝜼Nn+1):𝑫(𝝍)λeΩb(𝜼Nn+1)(𝝍)2μvΩb𝑫(𝝃Nn+1):𝑫(𝝍)\displaystyle-2\mu_{e}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}^{n+1}_{N}):\boldsymbol{D}(\boldsymbol{\psi})-\lambda_{e}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}^{n+1}_{N})(\nabla\cdot\boldsymbol{\psi})-2\mu_{v}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\xi}^{n+1}_{N}):\boldsymbol{D}(\boldsymbol{\psi})
λvΩb(𝝃Nn+1)(𝝍)+αΩb𝒥b(ηNn)δpNn+1b(ηNn)δ𝝍ΓΔωNn+12Δφ.\displaystyle-\lambda_{v}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\xi}^{n+1}_{N})(\nabla\cdot\boldsymbol{\psi})+\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\psi}-\int_{\Gamma}\Delta\omega^{n+\frac{1}{2}}_{N}\cdot\Delta\varphi.

The estimate for the first term in (88) will follow if we can estimate the right-hand side in terms of the 𝒬v{\mathcal{Q}_{v}^{\prime}} norm. For this purpose consider an arbitrary (𝝍,φ)𝒬v1||(\boldsymbol{\psi},\varphi)||_{\mathcal{Q}_{v}}\leq 1, so that 𝝍H2(Ωb)1||\boldsymbol{\psi}||_{H^{2}(\Omega_{b})}\leq 1 and φH02(Γ)1||\varphi||_{H_{0}^{2}(\Gamma)}\leq 1. By the uniform estimates in Lemma 7.1 and the regularity of the test functions in (87), it is clear that the terms on the right hand side are all uniformly bounded by a constant CC, independent of (𝝍,φ)𝒬v1||(\boldsymbol{\psi},\varphi)||_{\mathcal{Q}_{v}}\leq 1, except possibly the term

αΩb𝒥b(ηNn)δpNn+1b(ηNn)δ𝝍.\displaystyle\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\psi}.

To estimate this term we recall the definitions

𝒥b(ηNn)δ=det(𝑰+(𝜼Nn)δ),(ηNn)δ𝝍=tr[𝝍(I+(𝜼Nn)δ)1].\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}=\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}),\qquad\nabla^{(\eta^{n}_{N})^{\delta}}\cdot\boldsymbol{\psi}=\text{tr}\left[\nabla\boldsymbol{\psi}\cdot(I+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})^{-1}\right].

By assumption 2C (72) and the fact that 𝝍H1(Ωb)1||\boldsymbol{\psi}||_{H^{1}(\Omega_{b})}\leq 1, we have that (ηNn)δ𝝍L2(Ωb)||\nabla^{(\eta^{n}_{N})^{\delta}}\cdot\boldsymbol{\psi}||_{L^{2}(\Omega_{b})} is uniformly bounded, while by the boundedness of 𝜼Nn\boldsymbol{\eta}^{n}_{N} in H1(Ωb)H^{1}(\Omega_{b}), we have that |𝒥b(ηNn)δ|C|\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}|\leq C. Therefore, using the fact that pNp_{N} is uniformly bounded in L(0,T;L2(Ωb))L^{\infty}(0,T;L^{2}(\Omega_{b})), we obtain the desired estimate

|αΩb𝒥b(ηNn)δpNn+1b(ηNn)δ𝝍|C.\left|\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}\cdot\boldsymbol{\psi}\right|\leq C.

Finally, we conclude that

(𝝃Nn+1,ζNn+1)(𝝃Nn,ζNn)Δt𝒬vC, for a constant C that is independent of n and N,\left|\left|\frac{(\boldsymbol{\xi}^{n+1}_{N},\zeta^{n+1}_{N})-(\boldsymbol{\xi}^{n}_{N},\zeta^{n}_{N})}{\Delta t}\right|\right|_{\mathcal{Q}_{v}^{\prime}}\leq C,\ \text{ for a constant $C$ that is independent of $n$ and $N$},

and since

n=1N1(Δt)(𝝃Nn+1,ζNn+1)(𝝃Nn,ζNn)Δt𝒬v(Δt)n=1N1CCT,\sum_{n=1}^{N-1}(\Delta t)\left|\left|\frac{(\boldsymbol{\xi}^{n+1}_{N},\zeta^{n+1}_{N})-(\boldsymbol{\xi}^{n}_{N},\zeta^{n}_{N})}{\Delta t}\right|\right|_{\mathcal{Q}_{v}^{\prime}}\\ \leq(\Delta t)\sum_{n=1}^{N-1}C\leq CT,

we conclude that (88) holds for a uniform constant CC. This establishes the desired result.

8.4 Compactness for the pore pressure

Theorem 8.2.

There exists a subsequence such that

pNpstrongly in L2(0,T;L2(Ωb)).p_{N}\to p\ \text{strongly in $L^{2}(0,T;L^{2}(\Omega_{b}))$}.
Proof.

The proof is based on a similar application of the Dreher-Jüngel compactness criterion for piecewise constant functions [31] as in the previous compactness result. We first observe that we have the following chain of embeddings H1(Ωb)L2(Ωb)(VpH2(Ωb))H^{1}(\Omega_{b})\subset\subset L^{2}(\Omega_{b})\subset(V_{p}\cap H^{2}(\Omega_{b}))^{\prime}, and so by the Dreher-Jüngel compactness criterion [31] it suffices to show that the following inequality holds for a constant CC independent of NN:

τΔtpNpNΔtL1(Δt,T;(VpH2(Ωb)))+pNL2(0,T;H1(Ωb))C.\left|\left|\frac{\tau_{\Delta t}p_{N}-p_{N}}{\Delta t}\right|\right|_{L^{1}(\Delta t,T;(V_{p}\cap H^{2}(\Omega_{b}))^{\prime})}+||p_{N}||_{L^{2}(0,T;H^{1}(\Omega_{b}))}\leq C. (89)

To obtain this estimate, we observe that the approximate solutions for the pore pressure satisfy the following weak formulation for all test functions rVpr\in V_{p}, where VpV_{p} is defined by (34):

c0Ωb(pNn+1pNnΔt)rαΩb𝒥b(ηNn)δ𝜼˙Nn+1b(ηNn)δrαΓ(𝜼˙Nn+1𝒏(ωNn)δ)r\displaystyle c_{0}\int_{\Omega_{b}}\left(\frac{p^{n+1}_{N}-p^{n}_{N}}{\Delta t}\right)\cdot r-\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\dot{\boldsymbol{\eta}}^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r-\alpha\int_{\Gamma}(\dot{\boldsymbol{\eta}}^{n+1}_{N}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}})r
+κΩb𝒥b(ηNn)δb(ηNn)δpNn+1b(ηNn)δrΓ[(𝒖Nn+1𝜼˙Nn+1)𝒏ωNn]r=0.\displaystyle+\kappa\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r-\int_{\Gamma}[(\boldsymbol{u}^{n+1}_{N}-\dot{\boldsymbol{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}^{\omega^{n}_{N}}]r=0.

We use more regularity for the test space VpH2(Ωb)V_{p}\cap H^{2}(\Omega_{b}) to make the following estimates simpler. We compute that for any rVpH2(Ωb)r\in V_{p}\cap H^{2}(\Omega_{b}) we have

c0Ωb(pNn+1pNnΔt)r=αΩb𝒥b(ηNn)δ𝝃Nn+1b(ηNn)δr+αΓ(ζNn+1𝒆y𝒏(ωNn)δ)r\displaystyle c_{0}\int_{\Omega_{b}}\left(\frac{p^{n+1}_{N}-p^{n}_{N}}{\Delta t}\right)\cdot r=\alpha\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\boldsymbol{\xi}^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r+\alpha\int_{\Gamma}(\zeta^{n+1}_{N}\boldsymbol{e}_{y}\cdot\boldsymbol{n}^{(\omega^{n}_{N})^{\delta}})r
κΩb𝒥b(ηNn)δb(ηNn)δpNn+1b(ηNn)δr+Γ[(𝒖Nn+1ζNn+1𝒆y)𝒏ωNn]r.\displaystyle-\kappa\int_{\Omega_{b}}\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}\cdot\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r+\int_{\Gamma}[(\boldsymbol{u}^{n+1}_{N}-\zeta^{n+1}_{N}\boldsymbol{e}_{y})\cdot\boldsymbol{n}^{\omega^{n}_{N}}]r.

We estimate the right hand side for rVpH2(Ωb)1||r||_{V_{p}\cap H^{2}(\Omega_{b})}\leq 1. Recall that 𝒥b(ηNn)δ=det(𝑰+(𝜼Nn)δ)\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b}=\det(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta}),

b(ηNn)δr=(rx~,ry~)(𝑰+(𝜼Nn)δ)1, and b(ηNn)δpNn+1=(pNn+1x~,pNn+1y~)(𝑰+(𝜼Nn)δ)1.\nabla^{(\eta^{n}_{N})^{\delta}}_{b}r=\left(\frac{\partial r}{\partial\tilde{x}},\frac{\partial r}{\partial\tilde{y}}\right)\cdot(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})^{-1},\quad\text{ and }\quad\nabla^{(\eta^{n}_{N})^{\delta}}_{b}p^{n+1}_{N}=\left(\frac{\partial p^{n+1}_{N}}{\partial\tilde{x}},\frac{\partial p^{n+1}_{N}}{\partial\tilde{y}}\right)\cdot(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})^{-1}.

We have by Assumption 2C (72) that |(𝑰+(𝜼Nn)δ)1||(\boldsymbol{I}+\nabla(\boldsymbol{\eta}^{n}_{N})^{\delta})^{-1}| is uniformly bounded, and furthermore, 𝒥b(ηNn)δ\mathcal{J}^{(\eta^{n}_{N})^{\delta}}_{b} is positive and bounded above. By combining these facts with standard estimates we obtain that

pNn+1pNnΔt(VpH2(Ωb))C for a constant C that is independent of n and N.\left|\left|\frac{p^{n+1}_{N}-p^{n}_{N}}{\Delta t}\right|\right|_{(V_{p}\cap H^{2}(\Omega_{b}))^{\prime}}\leq C\ \text{ for a constant $C$ that is independent of $n$ and $N$.}

Combining this with the fact that pNp_{N} is uniformly bounded in L2(0,T;H1(Ωb))L^{2}(0,T;H^{1}(\Omega_{b})) gives the desired estimate (89). ∎

8.5 Compactness for the fluid velocity

We will obtain convergence of the fluid velocity along a subsequence by using a generalized Aubin-Lions compactness theorem for functions defined on moving domains, stated as Theorem 3.1 in [57]. To help the reader, we state Theorem 3.1 from [57] at the end of this manuscript, in the appendix, Section A.3. The reason we must use the generalized Aubin-Lions compactness theorem is that the approximate fluid velocities are defined on different time-dependent fluid domains. To prepare for an application of the generalized Aubin-Lions compactness argument we will map our approximate fluid problem back onto the physical domain

Ωf,Nn={(x,y)2:0xL,RyωNn(x)},\Omega^{n}_{f,N}=\{(x,y)\in\mathbb{R}^{2}:0\leq x\leq L,-R\leq y\leq\omega^{n}_{N}(x)\},

where we redefine the fluid velocity solution and test spaces as follows:

VNn+1={𝒖H1(Ωf,Nn):𝒖=0 on Ωf,Nn,𝒖=0 on Ωf,NnΓNn},QNn=VNn+1H3(Ωf,Nn).V^{n+1}_{N}=\{\boldsymbol{u}\in H^{1}(\Omega^{n}_{f,N}):\nabla\cdot\boldsymbol{u}=0\text{ on }\Omega^{n}_{f,N},\boldsymbol{u}=0\text{ on }\partial\Omega^{n}_{f,N}\setminus\Gamma^{n}_{N}\},\quad Q^{n}_{N}=V^{n+1}_{N}\cap H^{3}(\Omega^{n}_{f,N}). (90)

The approximate fluid velocity 𝒖Nn+1VNn+1\boldsymbol{u}^{n+1}_{N}\in V^{n+1}_{N} on the physical domain satisfies the following semidiscrete formulation:

Ωf,Nn𝒖Nn+1𝒖~NnΔt𝒗+2νΩf,Nn𝑫(𝒖Nn+1):𝑫(𝒗)\displaystyle\int_{\Omega^{n}_{f,N}}\frac{\boldsymbol{u}^{n+1}_{N}-\tilde{\boldsymbol{u}}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}+2\nu\int_{\Omega^{n}_{f,N}}\boldsymbol{D}(\boldsymbol{u}_{N}^{n+1}):\boldsymbol{D}(\boldsymbol{v})
+12Ωf,Nn[((𝒖~NnζNn+12R+yR+ωNn𝒆y)𝒖Nn+1)𝒗((𝒖~NnζNn+12R+yR+ωNn𝒆y)𝒗)𝒖Nn+1]\displaystyle+\frac{1}{2}\int_{\Omega^{n}_{f,N}}\left[\left(\left(\tilde{\boldsymbol{u}}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R+\omega^{n}_{N}}\boldsymbol{e}_{y}\right)\cdot\nabla\boldsymbol{u}^{n+1}_{N}\right)\cdot\boldsymbol{v}-\left(\left(\tilde{\boldsymbol{u}}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R+\omega^{n}_{N}}\boldsymbol{e}_{y}\right)\cdot\nabla\boldsymbol{v}\right)\cdot\boldsymbol{u}^{n+1}_{N}\right]
+12RΩf,NnRR+ωNnζNn+12𝒖Nn+1𝒗+12ΓNn(𝒖Nn+1𝜼˙Nn+1)𝒏(𝒖~Nn𝒗)\displaystyle+\frac{1}{2R}\int_{\Omega^{n}_{f,N}}\frac{R}{R+\omega^{n}_{N}}\zeta^{n+\frac{1}{2}}_{N}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{v}+\frac{1}{2}\int_{\Gamma^{n}_{N}}(\boldsymbol{u}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}(\tilde{\boldsymbol{u}}^{n}_{N}\cdot\boldsymbol{v})
ΓNn(12𝒖Nn+1𝒖~NnpNn+1)(𝒗𝒏)βΓNn(𝜼˙Nn+1𝒖Nn+1)𝝉(𝒗𝝉)=0,𝒗QNn\displaystyle-\int_{\Gamma^{n}_{N}}\left(\frac{1}{2}\boldsymbol{u}^{n+1}_{N}\cdot\tilde{\boldsymbol{u}}^{n}_{N}-p^{n+1}_{N}\right)(\boldsymbol{v}\cdot\boldsymbol{n})-\beta\int_{\Gamma^{n}_{N}}(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}(\boldsymbol{v}\cdot\boldsymbol{\tau})=0,\ \forall\boldsymbol{v}\in Q^{n}_{N} (91)

where

𝒖~Nn=𝒖Nn𝚽fωNn1(𝚽fωNn)1,\tilde{\boldsymbol{u}}^{n}_{N}=\boldsymbol{u}^{n}_{N}\circ\boldsymbol{\Phi}^{\omega^{n-1}_{N}}_{f}\circ(\boldsymbol{\Phi}^{\omega^{n}_{N}}_{f})^{-1}, (92)

𝒖Nn\boldsymbol{u}^{n}_{N} is originally defined on Ωf,Nn1\Omega^{n-1}_{f,N}, and the ALE map 𝚽fωNn:ΩfΩf,Nn\boldsymbol{\Phi}^{\omega^{n}_{N}}_{f}:\Omega_{f}\to\Omega^{n}_{f,N} is defined by (16).

To be able to compare functions on different physical domains we introduce a maximal domain ΩfM\Omega^{M}_{f} which contains all the physical domains. The existence of such a domain, and the extensions of the velocity functions onto the maximal domain are discussed next.

8.5.1 Extension to maximal domain

We consider the following maximal fluid domain which contains all the physical fluid domains:

ΩfM={(x,y)2:0xL,RyM(x)},\Omega^{M}_{f}=\{(x,y)\in\mathbb{R}^{2}:0\leq x\leq L,-R\leq y\leq M(x)\},

where the function M(x)M(x) is obtained from the following proposition, established in Lemma 2.5 in [68] and Lemma 4.5 in [57] in the context of fluid-structure interaction between an incompressible viscous Newtonian fluid and an elastic Koiter shell:

Proposition 8.2.

There exists smooth functions m(x)m(x) and M(x)M(x) defined on Γ=[0,L]\Gamma=[0,L], satisfying m(0)=m(L)=M(0)=M(L)=0m(0)=m(L)=M(0)=M(L)=0, such that

m(x)ωNn(x)M(x), for all x[0,L],N, and n=0,1,,N.m(x)\leq\omega^{n}_{N}(x)\leq M(x),\qquad\text{ for all }x\in[0,L],N,\text{ and }n=0,1,...,N.

Furthermore, there exist smooth functions mNn,l(x)m^{n,l}_{N}(x) and MNn,l(x)M^{n,l}_{N}(x) defined for positive integers NN, n=0,1,,N1n=0,1,...,N-1 and l=0,1,,Nnl=0,1,...,N-n, such that

  1. 1.

    mNn,l(x)ωNn+i(x)MNn,l(x),m^{n,l}_{N}(x)\leq\omega^{n+i}_{N}(x)\leq M^{n,l}_{N}(x),\qquad for all x[0,L]x\in[0,L] and i=0,1,,li=0,1,...,l.

  2. 2.

    MNn,l(x)mNn,l(x)ClΔt,M^{n,l}_{N}(x)-m^{n,l}_{N}(x)\leq C\sqrt{l\Delta t},\qquad for all x[0,L]x\in[0,L].

  3. 3.

    MNn,l(x)mNn,l(x)L2(Γ)C(lΔt)||M^{n,l}_{N}(x)-m^{n,l}_{N}(x)||_{L^{2}(\Gamma)}\leq C(l\Delta t),

where CC is independent of nn, ll, and NN. Finally, the functions MNn,l(x)M^{n,l}_{N}(x) and mNn,l(x)m^{n,l}_{N}(x) for all nn, ll, and NN, are Lipschitz continuous with a Lipschitz constant that is uniformly bounded above by some constant L>0L>0 independent of nn, ll, and NN.

Once the maximal fluid domain is defined, we can extend the fluid velocities 𝒖Nn\boldsymbol{u}^{n}_{N} from Ωf,Nn\Omega^{n}_{f,N} to this common maximal domain ΩfM\Omega^{M}_{f}, using extensions by zero in ΩfM(Ωf,Nn)c\Omega^{M}_{f}\cap(\Omega^{n}_{f,N})^{c}. Notice that since ωNn(x)\omega^{n}_{N}(x) are all uniformly Lipschitz, the extensions by zero of the H1H^{1} functions 𝒖Nn\boldsymbol{u}^{n}_{N} defined on Lipschitz domains to ΩfM\Omega^{M}_{f} are uniformly bounded in Hs(ΩfM)H^{s}(\Omega^{M}_{f}) for all ss such that 0<s<1/20<s<1/2. Indeed, we have the following lemma, which follows from Theorem 2.7 in [50].

Lemma 8.2.

The approximate fluid velocities {𝒖N}N=1\{\boldsymbol{u}_{N}\}_{N=1}^{\infty} defined on the maximal fluid domain ΩfM\Omega^{M}_{f} by extension by zero are uniformly bounded in L2(0,T;Hs(ΩfM))L^{2}(0,T;H^{s}(\Omega^{M}_{f})) for s(0,1/2)s\in(0,1/2).

8.5.2 Velocity convergence via a generalized Aubin-Lions compactness argument

We now show strong convergence as NN\to\infty along a subsequence of the approximate fluid velocities 𝒖N\boldsymbol{u}_{N}, which are now functions in time defined on the fixed maximal domain ΩfM\Omega^{M}_{f}.

Proposition 8.3.

The sequence 𝒖N\boldsymbol{u}_{N} is relatively compact in L2(0,T;L2(ΩfM))L^{2}(0,T;L^{2}(\Omega^{M}_{f})).

Proof.

The proof is based on using the generalized Aubin-Lions compactness theorem for problems on moving domains, which is Theorem 3.1 of [57], restated in this manuscript for the reader’s convenience as Theorem A.1 in Section A.3. For this purpose we define the Hilbert spaces VV and HH from the statement of the theorem to be

H=L2(ΩfM),V=Hs(ΩfM), for 0<s<1/2,H=L^{2}(\Omega^{M}_{f}),\qquad V=H^{s}(\Omega^{M}_{f}),\qquad\text{ for }0<s<1/2,

where we note that indeed VHV\subset\subset H as required by Theorem 3.1 in [57]. Additionally, the spaces (VΔtn,QΔtn)(V^{n}_{\Delta t},Q^{n}_{\Delta t}) from the statement of the theorem correspond to our spaces (VNn,QNn)(V^{n}_{N},Q^{n}_{N}) defined by (90). Notice that VNn×QNnV^{n}_{N}\times Q^{n}_{N} embeds continuously into V×VV\times V as required by the statement of Theorem 3.1 in [57], where the embedding can be achieved by the extension by zero operator to the maximal domain ΩfM\Omega^{M}_{f}, uniformly in nn and NN.

To obtain compactness of the sequence 𝒖N\boldsymbol{u}_{N} in L2(0,T,H)L^{2}(0,T,H), by Theorem 3.1 in [57], seven properties need to be satisfied by the sequence 𝒖N\boldsymbol{u}_{N} and the spaces VNnV^{n}_{N} and QNnQ^{n}_{N}. They are called Properties A1-3, B, and C1-3.

The proof that approximate solutions 𝒖N\boldsymbol{u}_{N} satisfy Properties A1-3 and C1-3 is analogous to the corresponding proof in [57] (Section 4.2). The main difficulty is to verify Property B, which is a condition on equicontinuity of 𝒖N\boldsymbol{u}_{N}, stated as follows:

Property B, [57]. There exists a constant C>0C>0 independent of NN such that

PNn𝒖Nn+1𝒖NnΔt(QNn)C(1+𝒖Nn+1VNn+1), for all n=0,1,,N1,\left|\left|P^{n}_{N}\frac{\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}^{n}_{N}}{\Delta t}\right|\right|_{(Q^{n}_{N})^{\prime}}\leq C\left(1+||\boldsymbol{u}^{n+1}_{N}||_{V^{n+1}_{N}}\right),\qquad\text{ for all }n=0,1,...,N-1, (93)

where PNnP^{n}_{N} denotes the orthogonal projection onto the closed subspace QNn¯H\overline{Q^{n}_{N}}^{H} of the Hilbert space HH.

The sequence 𝒖N\boldsymbol{u}_{N} constructed in this manuscript, however, does not satisfy this property. Nevertheless, 𝒖N\boldsymbol{u}_{N} satisfy the following generalized Property B which implies the desired equicontinuity under which the generalized Aubin-Lions theorem from [57] still holds:

Generalized Property B. There exist a constant CC independent of nn and NN, an exponent pp, 1p<21\leq p<2, and a sequence of nonnegative numbers {aNn}n=0N1\{a^{n}_{N}\}_{n=0}^{N-1} for each NN, satisfying (Δt)n=0N1|aNn|2C(\Delta t)\sum_{n=0}^{N-1}|a^{n}_{N}|^{2}\leq C uniformly in NN, such that

PNn𝒖Nn+1𝒖NnΔt(QNn)C(aNn+𝒖NnVNn+𝒖Nn+1VNn+1)p, for all n=0,1,,N1.\left|\left|P^{n}_{N}\frac{\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}^{n}_{N}}{\Delta t}\right|\right|_{(Q^{n}_{N})^{\prime}}\leq C\left(a^{n}_{N}+||\boldsymbol{u}^{n}_{N}||_{V^{n}_{N}}+||\boldsymbol{u}^{n+1}_{N}||_{V^{n+1}_{N}}\right)^{p},\qquad\text{ for all }n=0,1,...,N-1. (94)

Recall the statement of [57, Theorem 3.1], which can be found in the Appendix of the present manuscript in Section A.3, Theorem A.1.

Theorem 8.3 (Generalized Aubin-Lions Compactness Result II).

Assume that properties A1-3 and C1-3 of [57, Theorem 3.1] are satisfied. Furthermore, assume that Generalized Property B above is satisfied. Then conclusion of [57, Theorem 3.1] holds, namely, {𝐮N}\{{\bf u}_{N}\} is relatively compact in L2(0,T;H)L^{2}(0,T;H).

Proof.

We just need to prove that the essential equicontinuity estimate in the proof of [57, Theorem 3.1.] still holds under the modified assumption. In particular, for the original form of Property B in (93), one has from Lemma 3.1 in [57] the following equicontinuity estimate for a constant C>0C>0 that is independent of NN:

PΔtn,l(𝒖Nn+l𝒖Nn)(QNn,l)ClΔt.||P^{n,l}_{\Delta t}(\boldsymbol{u}^{n+l}_{N}-\boldsymbol{u}^{n}_{N})||_{(Q^{n,l}_{N})^{\prime}}\leq C\sqrt{l\Delta t}.

With the generalized form of Property B that we use above in (94), the same arguments as in the proof of Lemma 3.1 in [57] will still give rise to the following equicontinuity estimate for a constant C>0C>0 that is independent of NN:

PΔtn,l(𝒖Nn+l𝒖Nn)(QNn,l)C(lΔt)1p2,||P^{n,l}_{\Delta t}(\boldsymbol{u}^{n+l}_{N}-\boldsymbol{u}^{n}_{N})||_{(Q^{n,l}_{N})^{\prime}}\leq C(l\Delta t)^{1-\frac{p}{2}},

where the generalized Aubin-Lions compactness theorem on moving domains still holds with this new equicontinuity estimate. This is because 1p<21\leq p<2 and hence, C(lΔt)1p2C(l\Delta t)^{1-\frac{p}{2}} still converges to zero as Δt0\Delta t\to 0. ∎

We can now complete the proof of Proposition 8.3 by verifying that our sequence 𝒖Nn\boldsymbol{u}^{n}_{N} indeed satisfies the Generalized Property B. Verification that 𝐮Nn\boldsymbol{u}^{n}_{N} satisfies the Generalized Property B. First, recall that by definition,

PNn𝒖Nn+1𝒖NnΔt(QNn)=max𝒗QNn1|Ωf,Nn𝒖Nn+1𝒖NnΔt𝒗𝑑𝒙|.\left|\left|P^{n}_{N}\frac{\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}^{n}_{N}}{\Delta t}\right|\right|_{(Q^{n}_{N})^{\prime}}=\max_{||\boldsymbol{v}||_{Q^{n}_{N}}\leq 1}\left|\int_{\Omega^{n}_{f,N}}\frac{\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right|. (95)

To estimate the right hand-side, we use

|Ωf,Nn𝒖Nn+1𝒖nΔt𝒗𝑑𝒙||Ωf,Nn𝒖Nn+1𝒖~NnΔt𝒗𝑑𝒙|+|Ωf,Nn𝒖~Nn𝒖NnΔt𝒗𝑑𝒙|.\left|\int_{\Omega^{n}_{f,N}}\frac{\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}^{n}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right|\leq\left|\int_{\Omega^{n}_{f,N}}\frac{\boldsymbol{u}^{n+1}_{N}-\tilde{\boldsymbol{u}}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right|+\left|\int_{\Omega^{n}_{f,N}}\frac{\tilde{\boldsymbol{u}}^{n}_{N}-\boldsymbol{u}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right|. (96)

To estimate the first term on the right hand-side we use the semidiscrete formulation for the fluid velocity on the physical domain given by (8.5) to obtain

|Ωf,Nn𝒖Nn+1𝒖~NnΔt𝒗d𝒙|2ν|Ωf,Nn𝑫(𝒖Nn+1):𝑫(𝒗)|\displaystyle\left|\int_{\Omega^{n}_{f,N}}\frac{\boldsymbol{u}^{n+1}_{N}-\tilde{\boldsymbol{u}}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right|\leq 2\nu\left|\int_{\Omega^{n}_{f,N}}\boldsymbol{D}(\boldsymbol{u}_{N}^{n+1}):\boldsymbol{D}(\boldsymbol{v})\right|
+12|Ωf,Nn[((𝒖~NnζNn+12R+yR+ωNn𝒆y)𝒖Nn+1)𝒗((𝒖~NnζNn+12R+yR+ωNn𝒆y)𝒗)𝒖Nn+1]|\displaystyle+\frac{1}{2}\left|\int_{\Omega^{n}_{f,N}}\left[\left(\left(\tilde{\boldsymbol{u}}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R+\omega^{n}_{N}}\boldsymbol{e}_{y}\right)\cdot\nabla\boldsymbol{u}^{n+1}_{N}\right)\cdot\boldsymbol{v}-\left(\left(\tilde{\boldsymbol{u}}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R+\omega^{n}_{N}}\boldsymbol{e}_{y}\right)\cdot\nabla\boldsymbol{v}\right)\cdot\boldsymbol{u}^{n+1}_{N}\right]\right|
+12R|Ωf,NnRR+ωNnζNn+12𝒖Nn+1𝒗|+12|ΓNn(𝒖Nn+1𝜼˙Nn+1)𝒏(𝒖~Nn𝒗)|\displaystyle+\frac{1}{2R}\left|\int_{\Omega^{n}_{f,N}}\frac{R}{R+\omega^{n}_{N}}\zeta^{n+\frac{1}{2}}_{N}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{v}\right|+\frac{1}{2}\left|\int_{\Gamma^{n}_{N}}(\boldsymbol{u}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}(\tilde{\boldsymbol{u}}^{n}_{N}\cdot\boldsymbol{v})\right|
+|ΓNn(12𝒖Nn+1𝒖~NnpNn+1)(𝒗𝒏)|+β|ΓNn(𝜼˙Nn+1𝒖Nn+1)𝝉(𝒗𝝉)|.\displaystyle+\left|\int_{\Gamma^{n}_{N}}\left(\frac{1}{2}\boldsymbol{u}^{n+1}_{N}\cdot\tilde{\boldsymbol{u}}^{n}_{N}-p^{n+1}_{N}\right)(\boldsymbol{v}\cdot\boldsymbol{n})\right|+\beta\left|\int_{\Gamma^{n}_{N}}(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}(\boldsymbol{v}\cdot\boldsymbol{\tau})\right|. (97)

We can bound the terms on the right hand-side uniformly in nn, NN, and 𝒗QNn1||\boldsymbol{v}||_{Q^{n}_{N}}\leq 1 as follows. By the boundedness of 𝒖Nn+1\boldsymbol{u}^{n+1}_{N} in the uniform energy estimates we immediately have

2ν|Ωf,Nn𝑫(𝒖Nn+1):𝑫(𝒗)|C||𝒖Nn+1||H1(Ωf,Nn).2\nu\left|\int_{\Omega^{n}_{f,N}}\boldsymbol{D}(\boldsymbol{u}^{n+1}_{N}):\boldsymbol{D}(\boldsymbol{v})\right|\leq C||\boldsymbol{u}^{n+1}_{N}||_{H^{1}(\Omega^{n}_{f,N})}.

The second term on the right hand-side of the above inequality is bounded as follows. First notice that because 𝒗QNn1||\boldsymbol{v}||_{Q^{n}_{N}}\leq 1, and by the definition of QNnQ^{n}_{N} in (90), we have that 𝒗\boldsymbol{v} is bounded in H3(Ωf,Nn)H^{3}(\Omega^{n}_{f,N}), and hence, 𝒗\boldsymbol{v} and 𝒗\nabla\boldsymbol{v} are bounded in L(Ωf,Nn)L^{\infty}(\Omega^{n}_{f,N}). Furthermore, by the boundedness of the fluid velocity 𝒖Nn\boldsymbol{u}^{n}_{N} on the reference domain due to the uniform energy estimate, and by the uniform boundedness of the Jacobian of the ALE map ΦfωNn\Phi^{\omega^{n}_{N}}_{f}, we obtain the following bound:

12|Ωf,Nn[((𝒖~NnζNn+12R+yR+ωNn𝒆y)𝒖Nn+1)𝒗((𝒖~NnζNn+12R+yR+ωNn𝒆y)𝒗)𝒖Nn+1]|\displaystyle\frac{1}{2}\left|\int_{\Omega^{n}_{f,N}}\left[\left(\left(\tilde{\boldsymbol{u}}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R+\omega^{n}_{N}}\boldsymbol{e}_{y}\right)\cdot\nabla\boldsymbol{u}^{n+1}_{N}\right)\cdot\boldsymbol{v}-\left(\left(\tilde{\boldsymbol{u}}^{n}_{N}-\zeta^{n+\frac{1}{2}}_{N}\frac{R+y}{R+\omega^{n}_{N}}\boldsymbol{e}_{y}\right)\cdot\nabla\boldsymbol{v}\right)\cdot\boldsymbol{u}^{n+1}_{N}\right]\right|
C(𝒖~NnL2(Ωf,Nn)+ζn+12L2(Γ))𝒖Nn+1H1(Ωf,Nn)𝒗H3(Ωf,Nn)C𝒖Nn+1H1(Ωf,Nn).\displaystyle\leq C\left(||\tilde{\boldsymbol{u}}^{n}_{N}||_{L^{2}(\Omega^{n}_{f,N})}+||\zeta^{n+\frac{1}{2}}||_{L^{2}(\Gamma)}\right)||\boldsymbol{u}^{n+1}_{N}||_{H^{1}(\Omega^{n}_{f,N})}\cdot||\boldsymbol{v}||_{H^{3}(\Omega^{n}_{f,N})}\leq C||\boldsymbol{u}^{n+1}_{N}||_{H^{1}(\Omega^{n}_{f,N})}.

Similarly, the next term in (8.5.2) is bounded as follows:

12R|Ωf,NnRR+ωNnζNn+12𝒖Nn+1𝒗|CζNn+12L2(Γ)𝒖Nn+1L2(Ωf,Nn)𝒗H3(Ωf,Nn)C𝒖Nn+1L2(Ωf,Nn).\frac{1}{2R}\left|\int_{\Omega^{n}_{f,N}}\frac{R}{R+\omega^{n}_{N}}\zeta^{n+\frac{1}{2}}_{N}\boldsymbol{u}^{n+1}_{N}\cdot\boldsymbol{v}\right|\leq C||\zeta^{n+\frac{1}{2}}_{N}||_{L^{2}(\Gamma)}||\boldsymbol{u}^{n+1}_{N}||_{L^{2}(\Omega^{n}_{f,N})}\cdot||\boldsymbol{v}||_{H^{3}(\Omega^{n}_{f,N})}\leq C||\boldsymbol{u}^{n+1}_{N}||_{L^{2}(\Omega^{n}_{f,N})}.

To bound the next term we observe that 𝜼˙Nn+1L2(Γ)||\boldsymbol{\dot{\eta}}^{n+1}_{N}||_{L^{2}(\Gamma)} is bounded uniformly and furthermore, the arc length element on ΓNn\Gamma^{n}_{N} is uniformly bounded pointwise since ηNn\eta^{n}_{N} is uniformly bounded in H02(Γ)H_{0}^{2}(\Gamma). Therefore, by using the trace inequality on Ωf\Omega_{f} we have the following estimate:

12\displaystyle\frac{1}{2} |ΓNn(𝒖Nn+1𝜼˙Nn+1)𝒏(𝒖~Nn𝒗)|\displaystyle\left|\int_{\Gamma^{n}_{N}}(\boldsymbol{u}^{n+1}_{N}-\boldsymbol{\dot{\eta}}^{n+1}_{N})\cdot\boldsymbol{n}(\tilde{\boldsymbol{u}}^{n}_{N}\cdot\boldsymbol{v})\right|
C(𝒖Nn+1L4(Γ)𝒖NnL4(Γ)𝒗L2(Γ)+𝜼˙Nn+1L2(Γ)𝒖NnL4(Γ)𝒗L4(Γ))\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{L^{4}(\Gamma)}\cdot||\boldsymbol{u}^{n}_{N}||_{L^{4}(\Gamma)}\cdot||\boldsymbol{v}||_{L^{2}(\Gamma)}+||\boldsymbol{\dot{\eta}}^{n+1}_{N}||_{L^{2}(\Gamma)}\cdot||\boldsymbol{u}^{n}_{N}||_{L^{4}(\Gamma)}\cdot||\boldsymbol{v}||_{L^{4}(\Gamma)}\right)
C(𝒖Nn+1H1/4(Γ)𝒖NnH1/4(Γ)𝒗H1(Ωf)+𝜼˙Nn+1L2(Γ)𝒖NnH1/4(Γ)𝒗H1/4(Γ))\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{H^{1/4}(\Gamma)}\cdot||\boldsymbol{u}^{n}_{N}||_{H^{1/4}(\Gamma)}\cdot||\boldsymbol{v}||_{H^{1}(\Omega_{f})}+||\boldsymbol{\dot{\eta}}^{n+1}_{N}||_{L^{2}(\Gamma)}\cdot||\boldsymbol{u}^{n}_{N}||_{H^{1/4}(\Gamma)}\cdot||\boldsymbol{v}||_{H^{1/4}(\Gamma)}\right)
C(𝒖Nn+1H3/4(Ωf)𝒖NnH3/4(Ωf)+𝜼˙Nn+1L2(Γ)𝒖NnH3/4(Ωf))𝒗H1(Ωf)\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{H^{3/4}(\Omega_{f})}\cdot||\boldsymbol{u}^{n}_{N}||_{H^{3/4}(\Omega_{f})}+||\boldsymbol{\dot{\eta}}^{n+1}_{N}||_{L^{2}(\Gamma)}\cdot||\boldsymbol{u}^{n}_{N}||_{H^{3/4}(\Omega_{f})}\right)\cdot||\boldsymbol{v}||_{H^{{{1}}}(\Omega_{f})}
C(𝒖Nn+1L2(Ωf)1/4𝒖Nn+1H1(Ωf)3/4𝒖NnL2(Ωf)1/4𝒖NnH1(Ωf)3/4+𝜼˙Nn+1L2(Γ)𝒖NnL2(Ωf)1/4𝒖NnH1(Ωf)3/4)\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{L^{2}(\Omega_{f})}^{1/4}||\boldsymbol{u}^{n+1}_{N}||_{H^{1}(\Omega_{f})}^{3/4}||\boldsymbol{u}^{n}_{N}||_{L^{2}(\Omega_{f})}^{1/4}||\boldsymbol{u}^{n}_{N}||_{H^{1}(\Omega_{f})}^{3/4}+||\boldsymbol{\dot{\eta}}^{n+1}_{N}||_{L^{2}(\Gamma)}||\boldsymbol{u}^{n}_{N}||_{L^{2}(\Omega_{f})}^{1/4}||\boldsymbol{u}^{n}_{N}||_{H^{1}(\Omega_{f})}^{3/4}\right)
C(𝒖Nn+1H1(Ωf)3/4𝒖NnH1(Ωf)3/4+𝒖NnH1(Ωf)3/4)C[1+(𝒖NnVNn+𝒖Nn+1VNn)3/2].\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{H^{1}(\Omega_{f})}^{3/4}\cdot||\boldsymbol{u}^{n}_{N}||_{H^{1}(\Omega_{f})}^{3/4}+||\boldsymbol{u}^{n}_{N}||_{H^{1}(\Omega_{f})}^{3/4}\right)\leq C\left[1+\left(||\boldsymbol{u}^{n}_{N}||_{V^{n}_{N}}+||\boldsymbol{u}^{n+1}_{N}||_{V^{n}_{N}}\right)^{3/2}\right].

More precisely, in the above chain of inequalities, we used the following results to justify the steps:

  • In the first inequality, we used Hölder’s inequality with the exponents 22, 44, and 44, the fact that 𝒖~Nn\tilde{\boldsymbol{u}}^{n}_{N} and 𝒖Nn\boldsymbol{u}^{n}_{N} have the same trace along the interface Γ\Gamma by (92), and the uniform boundedness of the Jacobian 𝒥ΓωNn\mathcal{J}^{\omega^{n}_{N}}_{\Gamma} defined in (18).

  • In the second inequality, we use Sobolev embedding on the one-dimensional domain Γ\Gamma.

  • In the third inequality, we use the trace theorem on Ωf\Omega_{f} to bound the trace along Γ\Gamma.

  • In the fourth inequality, we use interpolation, and the fact that since the test function satisfies 𝒗QNn1||\boldsymbol{v}||_{Q^{n}_{N}}\leq 1 where the test space QNnQ^{n}_{N} is defined in (90), we have that 𝒗H1(Ωf)||\boldsymbol{v}||_{H^{1}(\Omega_{f})} is uniformly bounded on the reference domain independently of nn and NN by the uniform bounds on ωN\omega_{N}. The uniform bounds on ωN\omega_{N} ensure uniform boundedness of 𝒥fω\mathcal{J}^{\omega}_{f} defined in (18), which is the term that appears when transforming integrals back to the reference domain.

  • In the fifth inequality, we use uniform boundedness of 𝒖NnL2(Ωf)||\boldsymbol{u}^{n}_{N}||_{L^{2}(\Omega_{f})} and 𝜼˙NnL2(Γ)||\dot{\boldsymbol{\eta}}^{n}_{N}||_{L^{2}(\Gamma)} which is implied by the previous uniform energy estimates, see Lemma 7.1.

  • We use the inequality ab12(a2+b2)ab\leq\frac{1}{2}(a^{2}+b^{2}).

Next, we estimate the second-to-last term in (8.5.2) in a way similar to the immediately preceding term, using similar justifications as given directly above:

|ΓNn\displaystyle\Big{|}\int_{\Gamma^{n}_{N}} (12𝒖Nn+1𝒖~NnpNn+1)(𝒗𝒏)|\displaystyle\left(\frac{1}{2}\boldsymbol{u}^{n+1}_{N}\cdot\tilde{\boldsymbol{u}}^{n}_{N}-p^{n+1}_{N}\right)(\boldsymbol{v}\cdot\boldsymbol{n})\Big{|}
C(𝒖Nn+1L4(Γ)𝒖NnL4(Γ)𝒗L2(Γ)+pNn+1L2(Γ)𝒗L2(Γ))\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{L^{4}(\Gamma)}\cdot||\boldsymbol{u}^{n}_{N}||_{L^{4}(\Gamma)}\cdot||\boldsymbol{v}||_{L^{2}(\Gamma)}+||p^{n+1}_{N}||_{L^{2}(\Gamma)}\cdot||\boldsymbol{v}||_{L^{2}(\Gamma)}\right)
C(𝒖Nn+1H1/4(Γ)𝒖NnH1/4(Γ)𝒗H1(Ωf)+pNn+1H1(Ωb)𝒗H1(Ωf))\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{H^{1/4}(\Gamma)}\cdot||\boldsymbol{u}^{n}_{N}||_{H^{1/4}(\Gamma)}\cdot||\boldsymbol{v}||_{H^{1}(\Omega_{f})}+||p^{n+1}_{N}||_{H^{1}(\Omega_{b})}\cdot||\boldsymbol{v}||_{H^{1}(\Omega_{f})}\right)
C(𝒖Nn+1H3/4(Ωf)𝒖NnH3/4(Ωf)+pNn+1H1(Ωb))\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{H^{3/4}(\Omega_{f})}\cdot||\boldsymbol{u}^{n}_{N}||_{H^{3/4}(\Omega_{f})}+||p^{n+1}_{N}||_{H^{1}(\Omega_{b})}\right)
C(𝒖Nn+1L2(Ωf)1/4𝒖Nn+1H1(Ωf)3/4𝒖NnL2(Ωf)1/4𝒖NnH1(Ωf)3/4+pNn+1H1(Ωb))\displaystyle\leq C\left(||\boldsymbol{u}^{n+1}_{N}||_{L^{2}(\Omega_{f})}^{1/4}||\boldsymbol{u}^{n+1}_{N}||_{H^{1}(\Omega_{f})}^{3/4}\cdot||\boldsymbol{u}^{n}_{N}||_{L^{2}(\Omega_{f})}^{1/4}||\boldsymbol{u}^{n}_{N}||_{H^{1}(\Omega_{f})}^{3/4}+||p^{n+1}_{N}||_{H^{1}(\Omega_{b})}\right)
C[1+(pNn+1H1(Ωb)+𝒖NnVNn+𝒖Nn+1VNn+1)3/2].\displaystyle\leq C\left[1+\left(||p^{n+1}_{N}||_{H^{1}(\Omega_{b})}+||\boldsymbol{u}^{n}_{N}||_{V^{n}_{N}}+||\boldsymbol{u}^{n+1}_{N}||_{V^{n+1}_{N}}\right)^{3/2}\right].

Finally, we estimate the last term

β|ΓNn(𝜼˙Nn+1𝒖Nn+1)𝝉(𝒗𝝉)|C(𝜼˙Nn+1L2(Γ)𝒗L2(Γ)+𝒖Nn+1L2(Γ)𝒗L2(Γ))\displaystyle\beta\left|\int_{\Gamma^{n}_{N}}(\boldsymbol{\dot{\eta}}^{n+1}_{N}-\boldsymbol{u}^{n+1}_{N})\cdot\boldsymbol{\tau}(\boldsymbol{v}\cdot\boldsymbol{\tau})\right|\leq C\left(||\dot{\boldsymbol{\eta}}^{n+1}_{N}||_{L^{2}(\Gamma)}\cdot||\boldsymbol{v}||_{L^{2}(\Gamma)}+||\boldsymbol{u}^{n+1}_{N}||_{L^{2}(\Gamma)}\cdot||\boldsymbol{v}||_{L^{2}(\Gamma)}\right)
C(1+𝒖Nn+1H1(Ωf)).\displaystyle\leq C\left(1+||\boldsymbol{u}^{n+1}_{N}||_{H^{1}(\Omega_{f})}\right).

Therefore, we obtain the final estimate of the first term in (96) which implies the existence of a constant CC independent of nn and NN, such that

max𝒗QNn1|Ωf,Nn𝒖Nn+1𝒖~NnΔt𝒗𝑑𝒙|C(aNn+𝒖NnVNn+𝒖Nn+1VNn+1)3/2, for aNn:=1+pNn+1H1(Ωb), where (Δt)n=0N1|aNn|22[(Δt)N+pNL2(0,T;H1(Ωb))2]C.\max_{||\boldsymbol{v}||_{Q^{n}_{N}}\leq 1}\left|\int_{\Omega^{n}_{f,N}}\frac{\boldsymbol{u}^{n+1}_{N}-\tilde{\boldsymbol{u}}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right|\leq C\left(a^{n}_{N}+||\boldsymbol{u}^{n}_{N}||_{V^{n}_{N}}+||\boldsymbol{u}^{n+1}_{N}||_{V^{n+1}_{N}}\right)^{3/2},\\ \text{ for }a^{n}_{N}:=1+||p^{n+1}_{N}||_{H^{1}(\Omega_{b})},\text{ where }(\Delta t)\sum_{n=0}^{N-1}|a^{n}_{N}|^{2}\leq 2\left[(\Delta t)N+||p_{N}||^{2}_{L^{2}(0,T;H^{1}(\Omega_{b}))}\right]\leq C. (98)

To complete the estimate (96), it remains to show that the second term |Ωf,Nn𝒖~Nn𝒖NnΔt𝒗𝑑𝒙|\displaystyle\left|\int_{\Omega^{n}_{f,N}}\frac{\tilde{\boldsymbol{u}}^{n}_{N}-\boldsymbol{u}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right| is uniformly bounded. This follows from the same estimates as those presented in [57] which show that there exists a constant CC independent of nn and NN, such that

max𝒗QNn1|Ωf,Nn𝒖~Nn𝒖NnΔt𝒗𝑑𝒙|C.\max_{||\boldsymbol{v}||_{Q^{n}_{N}}\leq 1}\left|\int_{\Omega_{f,N}^{n}}\frac{\tilde{\boldsymbol{u}}^{n}_{N}-\boldsymbol{u}^{n}_{N}}{\Delta t}\cdot\boldsymbol{v}d\boldsymbol{x}\right|\leq C. (99)

Combining (98) and (99) with (95) and (96) establishes Generalized Property B and completes the proof of Proposition 8.3. ∎

9 Passing to the limit in the regularized weak formulation

We have so far established the following strong convergence results:

𝜼N𝜼, in C(0,T;L2(Ωb)),\displaystyle\boldsymbol{\eta}_{N}\to\boldsymbol{\eta},\quad\text{ in }C(0,T;L^{2}(\Omega_{b})),
ωNω, in L(0,T;Hs(Γ)) for 0<s<2,\displaystyle\omega_{N}\to\omega,\quad\text{ in }L^{\infty}(0,T;H^{s}(\Gamma))\text{ for }0<s<2,
ζNζ, in L2(0,T;Hs(Γ)), for 1/2<s<0,\displaystyle\zeta_{N}^{*}\to\zeta,\quad\text{ in }L^{2}(0,T;H^{-s}(\Gamma)),\text{ for }-1/2<s<0,
ζNζ, in L2(0,T;Hs(Γ)), for 1/2<s<0,\displaystyle\zeta_{N}\to\zeta,\quad\text{ in }L^{2}(0,T;H^{-s}(\Gamma)),\text{ for }-1/2<s<0,
𝝃N𝝃, in L2(0,T;Hs(Ωb)), for 1/2<s<0,\displaystyle\boldsymbol{\xi}_{N}\to\boldsymbol{\xi},\quad\text{ in }L^{2}(0,T;H^{-s}(\Omega_{b})),\text{ for }-1/2<s<0,
𝒖N𝒖, in L2(0,T;L2(ΩfM)),pNp, in L2(0,T;L2(Ωb)),\displaystyle\boldsymbol{u}_{N}\to\boldsymbol{u},\quad\text{ in }L^{2}(0,T;L^{2}(\Omega^{M}_{f})),\qquad p_{N}\to p,\quad\text{ in }L^{2}(0,T;L^{2}(\Omega_{b})),

where ζN\zeta_{N}^{*} and ζN\zeta_{N} converge to the same limit in L2(0,T;Hs(Γ))L^{2}(0,T;H^{-s}(\Gamma)) for 1/2<s<0-1/2<s<0 due to the numerical dissipation estimates n=1NζNnζNn12L2(Γ)2C\sum_{n=1}^{N}||\zeta^{n}_{N}-\zeta^{n-\frac{1}{2}}_{N}||_{L^{2}(\Gamma)}^{2}\leq C, which imply that ζNζNL2(0,T;L2(Γ))0||\zeta_{N}-\zeta_{N}^{*}||_{L^{2}(0,T;L^{2}(\Gamma))}\to 0.

These strong convergence results will be used to pass to the limit in the semidiscrete formulation of the coupled problem (61) and show that the limit satisfies the weak formulation of the regularized problem. Before we can do this, there are two more convergence results that need to be established. One is a strong convergence result for the traces for the fluid velocity on the boundary of the fluid domain, and the other is a convergence result for the test functions, which are defined on approximate moving domains.

We start with the convergence result for the trace of the fluid velocity 𝒖^N|Γ\hat{\boldsymbol{u}}_{N}|_{\Gamma} along Γ\Gamma.

9.1 Strong convergence of the fluid velocity traces on Γ\Gamma

Proposition 9.1.

The traces 𝒖^N|Γ\hat{\boldsymbol{u}}_{N}|_{\Gamma} of the approximate fluid velocities on Γ\Gamma converge to the trace of the limiting fluid velocity on Γ\Gamma as NN\to\infty:

𝒖^N|Γ𝒖^|Γ, in L2(0,T;Hs12(Γ)), for s(0,1),\hat{\boldsymbol{u}}_{N}|_{\Gamma}\to\hat{\boldsymbol{u}}|_{\Gamma},\qquad\text{ in }L^{2}(0,T;H^{s-\frac{1}{2}}(\Gamma)),\qquad\text{ for }s\in(0,1),

where 𝒖^N=𝒖NΦfτΔtωN\hat{\boldsymbol{u}}_{N}=\boldsymbol{u}_{N}\circ\Phi^{\tau_{\Delta t}\omega_{N}}_{f} and 𝒖^=𝒖Φfω\hat{\boldsymbol{u}}=\boldsymbol{u}\circ\Phi^{\omega}_{f}.

To prove Proposition 9.1, we will use the following elementary lemma.

Lemma 9.1.

Suppose that the functions {fn}n=1\{f_{n}\}_{n=1}^{\infty} and ff are all uniformly bounded in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})) and fnff_{n}\to f in L2(0,T;L2(Ωf))L^{2}(0,T;L^{2}(\Omega_{f})). Then, fnff_{n}\to f in L2(0,T;Hs(Ωf))L^{2}(0,T;H^{s}(\Omega_{f})) for s(0,1)s\in(0,1) and hence fn|Γf|Γf_{n}|_{\Gamma}\to f|_{\Gamma} in L2(0,T;Hs12(Γ))L^{2}(0,T;H^{s-\frac{1}{2}}(\Gamma)) for s(1/2,1)s\in(1/2,1).

Proof of Lemma 9.1.

For s(0,1)s\in(0,1), we compute using Sobolev interpolation that

fnfL2(0,T;Hs(Ωf))2=0T(fnf)(t)Hs(Ωf)2𝑑t0T(fnf)(t)L2(Ωf)2(1s)(fnf)(t)H1(Ωf)2s𝑑tfnfL2(0,T;L2(Ωf))2(1s)fnfL2(0,T;H1(Ωf))2s.||f_{n}-f||_{L^{2}(0,T;H^{s}(\Omega_{f}))}^{2}=\int_{0}^{T}||(f_{n}-f)(t)||_{H^{s}(\Omega_{f})}^{2}dt\\ \leq\int_{0}^{T}||(f_{n}-f)(t)||_{L^{2}(\Omega_{f})}^{2(1-s)}\cdot||(f_{n}-f)(t)||_{H^{1}(\Omega_{f})}^{2s}dt\leq||f_{n}-f||_{L^{2}(0,T;L^{2}(\Omega_{f}))}^{2(1-s)}\cdot||f_{n}-f||_{L^{2}(0,T;H^{1}(\Omega_{f}))}^{2s}.

The result then follows from the fact that fnfL2(0,T;H1(Ωf))C||f_{n}-f||_{L^{2}(0,T;H^{1}(\Omega_{f}))}\leq C for a constant CC that does not depend on NN, the assumption that fnfL2(0,T;L2(Ωf))0||f_{n}-f||_{L^{2}(0,T;L^{2}(\Omega_{f}))}\to 0 as NN\to\infty, and the trace embedding which gives that ||fn|Γf|Γ||L2(0,T;Hs12(Γ))2||fnf||L2(0,T;Hs(Ωf))2||f_{n}|_{\Gamma}-f|_{\Gamma}||_{L^{2}(0,T;H^{s-\frac{1}{2}}(\Gamma))}^{2}\leq||f_{n}-f||_{L^{2}(0,T;H^{s}(\Omega_{f}))}^{2} for s(1/2,1)s\in(1/2,1). ∎

We can use the elementary lemma above to show the desired strong convergence of the fluid velocity traces.

Proof of Proposition 9.1.

We would like to combine the fact that 𝒖N𝒖\boldsymbol{u}_{N}\to\boldsymbol{u} in L2(0,T;L2(ΩfM))L^{2}(0,T;L^{2}(\Omega^{M}_{f})) with the fact that 𝒖N\boldsymbol{u}_{N} and 𝒖\boldsymbol{u} are all uniformly bounded in L2(0,T;H1(Ωf(t)))L^{2}(0,T;H^{1}(\Omega_{f}(t))) for all NN, to deduce strong convergence of the traces of the fluid velocities using Lemma 9.1. We do this in the following steps.

Step 1. We show that 𝒖^N𝒖^\hat{\boldsymbol{u}}_{N}\to\hat{\boldsymbol{u}} on L2(0,T;L2(Ωf))L^{2}(0,T;L^{2}(\Omega_{f})), for 𝒖^N\hat{\boldsymbol{u}}_{N} and 𝒖^\hat{\boldsymbol{u}} defined on the reference fluid domain.

To prove this, we compute 𝒖^N𝒖^L2(0,T;L2(Ωf))2||\hat{\boldsymbol{u}}_{N}-\hat{\boldsymbol{u}}||_{L^{2}(0,T;L^{2}(\Omega_{f}))}^{2} using the functions 𝒖N\boldsymbol{u}_{N} and 𝒖\boldsymbol{u} which are defined on the maximal domain ΩfM\Omega^{M}_{f}:

𝒖^N𝒖^L2(0,T;L2(Ωf))2=0TΩf|𝒖N(t,x,y+(1+yR)τΔtωN)𝒖(t,x,y+(1+yR)ω)|2\displaystyle||\hat{\boldsymbol{u}}_{N}-\hat{\boldsymbol{u}}||_{L^{2}(0,T;L^{2}(\Omega_{f}))}^{2}=\int_{0}^{T}\int_{\Omega_{f}}\left|\boldsymbol{u}_{N}\left(t,x,y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}\right)-\boldsymbol{u}\left(t,x,y+\left(1+\frac{y}{R}\right)\omega\right)\right|^{2}
2(I1+I2),\displaystyle\leq 2(I_{1}+I_{2}),

where

I1=0TΩf|𝒖N(t,x,y+(1+yR)τΔtωN)𝒖(t,x,y+(1+yR)τΔtωN)|2,I_{1}=\int_{0}^{T}\int_{\Omega_{f}}\left|\boldsymbol{u}_{N}\left(t,x,y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}\right)-\boldsymbol{u}\left(t,x,y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}\right)\right|^{2},
I2=0TΩf|𝒖(t,x,y+(1+yR)τΔtωN)𝒖(t,x,y+(1+yR)ω)|2.I_{2}=\int_{0}^{T}\int_{\Omega_{f}}\left|\boldsymbol{u}\left(t,x,y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}\right)-\boldsymbol{u}\left(t,x,y+\left(1+\frac{y}{R}\right)\omega\right)\right|^{2}.

We show that I10I_{1}\to 0 as NN\to\infty by using the fact that 1+ωNnR1+\frac{\omega^{n}_{N}}{R} is uniformly bounded from above by a positive constant, and the fact that ΩfM\Omega^{M}_{f} contains all of the domains Ωf,Nn\Omega^{n}_{f,N}, so that we can estimate:

I1=n=0N1nΔt(n+1)ΔtΩf,Nn(1+ωNnR)|𝒖Nn+1𝒖|2Cn=0N1nΔt(n+1)ΔtΩf,Nn|𝒖Nn+1𝒖|2\displaystyle I_{1}=\sum_{n=0}^{N-1}\int_{n\Delta t}^{(n+1)\Delta t}\int_{\Omega^{n}_{f,N}}\left(1+\frac{\omega^{n}_{N}}{R}\right)|\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}|^{2}\leq C\sum_{n=0}^{N-1}\int_{n\Delta t}^{(n+1)\Delta t}\int_{\Omega^{n}_{f,N}}|\boldsymbol{u}^{n+1}_{N}-\boldsymbol{u}|^{2}
C𝒖N𝒖L2(0,T;L2(ΩfM))20.\displaystyle\leq C||\boldsymbol{u}_{N}-\boldsymbol{u}||^{2}_{L^{2}(0,T;L^{2}(\Omega^{M}_{f}))}\to 0.

For I2I_{2}, we break up the integral into two parts:

I2=I2,1+I2,2,I_{2}=I_{2,1}+I_{2,2},

where

I2,1=0T0LRmin(0,y(t,x))|𝒖(t,x,y+(1+yR)τΔtωN)𝒖(t,x,y+(1+yR)ω)|2,\displaystyle I_{2,1}=\int_{0}^{T}\int_{0}^{L}\int_{-R}^{\min(0,y^{*}(t,x))}\left|\boldsymbol{u}\left(t,x,y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}\right)-\boldsymbol{u}\left(t,x,y+\left(1+\frac{y}{R}\right)\omega\right)\right|^{2},
I2,2=0T0Lmin(0,y(t,x))0|𝒖(t,x,y+(1+yR)τΔtωN)|2,\displaystyle I_{2,2}=\int_{0}^{T}\int_{0}^{L}\int_{\min(0,y^{*}(t,x))}^{0}\left|\boldsymbol{u}\left(t,x,y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}\right)\right|^{2},

for y(t,x)=ωτΔtωNR+τΔtωNy^{*}(t,x)=\frac{\omega-\tau_{\Delta t}\omega_{N}}{R+\tau_{\Delta t}\omega_{N}}. We can interpret y(t,x)y^{*}(t,x) as the yy value for which y+(1+yR)τΔtωN=ωy+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}=\omega. Now, note that

I2,1\displaystyle I_{2,1} 0T0LRmin(0,y(t,x))(y+(1+yR)τΔtωNy+(1+yR)ω|y𝒖(t,x,y)|𝑑y)2\displaystyle\leq\int_{0}^{T}\int_{0}^{L}\int_{-R}^{\min(0,y^{*}(t,x))}\left(\int_{y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}}^{y+\left(1+\frac{y}{R}\right)\omega}|\partial_{y}\boldsymbol{u}(t,x,y^{\prime})|dy^{\prime}\right)^{2}
0T0LRmin(0,y(t,x))(y+(1+yR)τΔtωNy+(1+yR)ω|y𝒖(t,x,y)|2𝑑y)(1+yR)|ωτΔtωN|,\displaystyle\leq\int_{0}^{T}\int_{0}^{L}\int_{-R}^{\min(0,y^{*}(t,x))}\left(\int_{y+\left(1+\frac{y}{R}\right)\tau_{\Delta t}\omega_{N}}^{y+\left(1+\frac{y}{R}\right)\omega}|\partial_{y}\boldsymbol{u}(t,x,y^{\prime})|^{2}dy^{\prime}\right)\cdot\left(1+\frac{y}{R}\right)\cdot|\omega-\tau_{\Delta t}\omega_{N}|,

where we applied Cauchy-Schwarz to the inner dydy^{\prime} integral. We note that τΔtωNω\tau_{\Delta t}\omega_{N}\to\omega pointwise uniformly on [0,T]×Γ[0,T]\times\Gamma as NN\to\infty by the convergence ω¯Nω\overline{\omega}_{N}\to\omega in C(0,T;Hs(Γ))C(0,T;H^{s}(\Gamma)) for 0<s<20<s<2 (from Proposition 8.1) and the estimate (81). Combining this with the fact that 𝒖L2(0,T;L2(Ωfω(t)))||\nabla\boldsymbol{u}||_{L^{2}(0,T;L^{2}(\Omega^{\omega}_{f}(t)))} is bounded, we have that I2,10I_{2,1}\to 0 as NN\to\infty.

Next, by the fundamental theorem of calculus and Jensen’s inequality,

I2,2\displaystyle I_{2,2} 0T0L|min(0,y(t,x))|maxw[R,ω(t,x)]|𝒖(t,x,w)|2\displaystyle\leq\int_{0}^{T}\int_{0}^{L}|\min(0,y^{*}(t,x))|\cdot\max_{w\in[-R,\omega(t,x)]}|\boldsymbol{u}(t,x,w)|^{2}
C0T0L|min(0,y(t,x))|Rω(t,x)|r𝒖(t,x,y)|2𝑑y,\displaystyle\leq{{C}}\int_{0}^{T}\int_{0}^{L}|\min(0,y^{*}(t,x))|\cdot\int_{-R}^{\omega(t,x)}|\partial_{r}\boldsymbol{u}(t,x,y^{\prime})|^{2}dy^{\prime},

for a constant CC that is independent of NN, so we conclude that I2,20I_{2,2}\to 0 as NN\to\infty by the fact that |min(0,y(t,x))|0|\min(0,y^{*}(t,x))|\to 0 uniformly on [0,T]×Γ[0,T]\times\Gamma, and by the boundedness of 𝒖L2(0,T;L2(Ωfω(t))||\nabla\boldsymbol{u}||_{L^{2}(0,T;L^{2}(\Omega^{\omega}_{f}(t))}. Thus, we have that 𝒖^N𝒖^L2(0,T;L2(Ωf)))0||\hat{\boldsymbol{u}}_{N}-\hat{\boldsymbol{u}}||_{L^{2}(0,T;L^{2}(\Omega_{f})))}\to 0. Note that this maximum is well defined for almost every (t,x)(t,x), since Sobolev functions are absolutely continuous for almost every line {(t,x,y):Ryω(t,x)}\{(t,x,y):-R\leq y\leq\omega(t,x)\} with fixed (t,x)(t,x), see [33, Theorem 2, Section 4.9].

Step 2. We claim that the functions 𝒖^N\hat{\boldsymbol{u}}_{N} for positive integers NN and 𝒖^\hat{\boldsymbol{u}} are all uniformly bounded in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})). Recall from Lemma 7.1 that the approximate solutions 𝒖^N\hat{\boldsymbol{u}}_{N} are uniformly bounded in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})). Since 𝒖^\hat{\boldsymbol{u}} is the strong limit of 𝒖^N\hat{\boldsymbol{u}}_{N} in L2(0,T;L2(Ωf))L^{2}(0,T;L^{2}(\Omega_{f})) and 𝒖^N\hat{\boldsymbol{u}}_{N} converge weakly in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})) along a subsequence to a weak limit which hence must also be 𝒖^\hat{\boldsymbol{u}}, we conclude that 𝒖^\hat{\boldsymbol{u}} is also in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})), which establishes the desired result of this step.

Step 3. From Step 1, we have that 𝒖^N𝒖^\hat{\boldsymbol{u}}_{N}\to\hat{\boldsymbol{u}} in L2(0,T;L2(Ωf))L^{2}(0,T;L^{2}(\Omega_{f})) and from Step 2, the functions 𝒖^N\hat{\boldsymbol{u}}_{N} and 𝒖^\hat{\boldsymbol{u}} are bounded in L2(0,T;H1(Ωf))L^{2}(0,T;H^{1}(\Omega_{f})) independently of NN, so we can conclude the proof of Proposition 9.1 by using Lemma 9.1. ∎

9.2 Convergence of the test functions on approximate fluid domains

The main difficulty in passing to the limit will be the test functions for the fluid velocity. In particular, on the fixed reference domain Ωf\Omega_{f} for the fluid, we note that the test functions for the fluid velocity in (39) satisfy fω𝒗=0 on Ωf\nabla^{\omega}_{f}\cdot\boldsymbol{v}=0\text{ on }\Omega_{f}, where ω\omega is the solution for the plate displacement. However, the test functions for the fluid velocity in the semidiscrete formulation in the semidiscrete test space 𝒬Nn+1\mathcal{Q}^{n+1}_{N}, defined by (53), satisfy ωNn𝒗=0 on Ωf\nabla^{\omega^{n}_{N}}\cdot\boldsymbol{v}=0\text{ on }\Omega_{f}. Hence, we need a way of comparing test functions in 𝒬Nn+1\mathcal{Q}^{n+1}_{N} to test functions in the actual test space 𝒱testω\mathcal{V}^{\omega}_{\text{test}}.

To do this, recall that we have defined the maximal domain ΩfM\Omega^{M}_{f} that contains all of the numerical fluid domains Ωf,Nn\Omega_{f,N}^{n}. We then propose to work with the test functions that are defined on ΩfM\Omega^{M}_{f}, and are constructed in such a way that the restrictions of those test functions to the domain defined by the plate displacement ω\omega, and composed with the ALE mapping Φfω\Phi^{\omega}_{f} defined in (16), gives a space of test functions 𝒳fω\mathcal{X}^{\omega}_{f} that is dense in the fluid velocity test space 𝒱fω\mathcal{V}^{\omega}_{f}. The space of all such test functions defined on ΩfM\Omega^{M}_{f} is denoted by 𝒳\mathcal{X} and it is defined as follows.

The test space 𝒳\mathcal{X}: The test space 𝒳\mathcal{X} consists of functions 𝒗Cc1([0,T);H1(ΩfM))\boldsymbol{v}\in C_{c}^{1}([0,T);H^{1}(\Omega^{M}_{f})) satisfying the following properties for each t[0,T)t\in[0,T):

  1. 1.

    For each t[0,T)t\in[0,T), 𝒗(t)\boldsymbol{v}(t) is a smooth vector-valued function on ΩfM\Omega^{M}_{f}.

  2. 2.

    𝒗(t)=0\nabla\cdot\boldsymbol{v}(t)=0 on ΩfM\Omega^{M}_{f} for all t[0,T)t\in[0,T).

  3. 3.

    𝒗(t)=0\boldsymbol{v}(t)=0 on ΩfMΓM\partial\Omega^{M}_{f}\setminus\Gamma_{M} for all t[0,T)t\in[0,T), where ΓM={(x,M(x)):0xL}\Gamma_{M}=\{(x,M(x)):0\leq x\leq L\} is the top boundary of the maximal fluid domain ΩfM\Omega^{M}_{f}.

Given 𝒗𝒳\boldsymbol{v}\in\mathcal{X}, define

𝒗~=𝒗|Ωfω𝚽fωand𝒗~N=𝒗|ΩfωN𝚽fωN.\tilde{\boldsymbol{v}}=\boldsymbol{v}|_{\Omega^{\omega}_{f}}\circ\boldsymbol{\Phi}^{\omega}_{f}\quad{\rm and}\quad\tilde{\boldsymbol{v}}_{N}=\boldsymbol{v}|_{\Omega^{\omega_{N}}_{f}}\circ\boldsymbol{\Phi}^{\omega_{N}}_{f}. (100)

The test functions 𝒗~\tilde{\boldsymbol{v}} are dense in the fluid velocity test space 𝒱fω\mathcal{V}^{\omega}_{f} associated with the fixed domain formulation, and the test functions 𝒗~N\tilde{\boldsymbol{v}}_{N} restricted to [nΔt,(n+1)Δt)[n\Delta t,(n+1)\Delta t) are dense in VfωNnV^{\omega^{n}_{N}}_{f}, where VfωNnV^{\omega^{n}_{N}}_{f} is the velocity test space for the semidiscretized problem(s) given in (53). Therefore, for each fixed NN, we can consider the semidiscrete formulation with the test function 𝒗~N\tilde{\boldsymbol{v}}_{N}, which we emphasize is discontinuous in time, due to the jumps in ωN\omega_{N} at each nΔtn\Delta t. To pass to the limit as NN\to\infty we can use the same approach as in Lemma 7.1 in [52] and Lemma 2.8 in [68], to obtain the following strong convergence results of the velocity test functions 𝒗~N\tilde{\boldsymbol{v}}_{N} and their gradients, which will allow us to pass to the limit in the semidiscrete weak formulations:

Proposition 9.2.

Consider 𝒗𝒳\boldsymbol{v}\in\mathcal{X}, and 𝒗~\tilde{\boldsymbol{v}} and 𝒗~N\tilde{\boldsymbol{v}}_{N} defined in (100). Then

𝒗~N𝒗,𝒗~N𝒗,\tilde{\boldsymbol{v}}_{N}\to\boldsymbol{v},\qquad\nabla\tilde{\boldsymbol{v}}_{N}\to\nabla\boldsymbol{v},

pointwise, uniformly on [0,T]×Ωf[0,T]\times\Omega_{f}, as NN\to\infty.

Remark 9.1.

We emphasize that we were able to construct such a test space 𝒳{\cal{X}} because in the definition of the full test space 𝒱testω\mathcal{V}^{\omega}_{\text{test}} in (39), the only component of the test space whose definition depends on the plate displacement is the fluid velocity, and fortunately, this fluid velocity component of the test space is decoupled from the other components. This is a feature of fluid-poroelastic structure interaction problems. In the purely elastic case of FSI, the fluid velocity test space is coupled to that of the structure, and the construction of the test functions that converge on the approximate fluid domains in more involving, see e.g., [21, 52, 54].

9.3 Passing to the limit

We are now in a position to pass to the limit in the semidiscrete formulation. To do this, we consider any test function (𝒗~N,φ,𝝍,r)(\tilde{\boldsymbol{v}}_{N},\varphi,\boldsymbol{\psi},r) for a given 𝒗𝒳\boldsymbol{v}\in\mathcal{X} and for each n=0,1,,N1n=0,1,...,N-1, we test the semidiscrete formulation (61) with (𝒗~N(t),φ(t),𝝍(t),r(t))(\tilde{\boldsymbol{v}}_{N}(t),\varphi(t),\boldsymbol{\psi}(t),r(t)) for each t[nΔt,(n+1)Δt)t\in[n\Delta t,(n+1)\Delta t), integrate the resulting expressions in time from t=nΔtt=n\Delta t to t=(n+1)Δtt=(n+1)\Delta t, and then finally sum over n=0,1,,N1n=0,1,...,N-1 to get an integral over the entire time interval [0,T][0,T]. Then, using the definition of the approximate solutions from Section 7, we thus obtain that for all (𝒗~N,φ,𝝍,r)(\tilde{\boldsymbol{v}}_{N},\varphi,\boldsymbol{\psi},r) in the test space with 𝒗𝒳\boldsymbol{v}\in\mathcal{X}, the following holds:

0TΩf(1+τΔtωNR)t𝒖¯N𝒗~N+120TΩf(1+τΔtωNR)[((τΔt𝒖NζNR+yR𝒆y)fτΔtωN𝒖N)𝒗~N\displaystyle\int_{0}^{T}\int_{\Omega_{f}}\left(1+\frac{\tau_{\Delta t}\omega_{N}}{R}\right)\partial_{t}\overline{\boldsymbol{u}}_{N}\cdot\tilde{\boldsymbol{v}}_{N}+\frac{1}{2}\int_{0}^{T}\int_{\Omega_{f}}\left(1+\frac{\tau_{\Delta t}\omega_{N}}{R}\right)\Bigg{[}\left(\left(\tau_{\Delta t}\boldsymbol{u}_{N}-\zeta_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\tau_{\Delta t}\omega_{N}}_{f}\boldsymbol{u}_{N}\right)\cdot\tilde{\boldsymbol{v}}_{N}
((τΔt𝒖NζNR+yR𝒆y)fτΔtωN𝒗~N)𝒖N]+12R0TΩfζN𝒖N𝒗~N\displaystyle-\left(\left(\tau_{\Delta t}\boldsymbol{u}_{N}-\zeta_{N}\frac{R+y}{R}\boldsymbol{e}_{y}\right)\cdot\nabla^{\tau_{\Delta t}\omega_{N}}_{f}\tilde{\boldsymbol{v}}_{N}\right)\cdot\boldsymbol{u}_{N}\Bigg{]}+\frac{1}{2R}\int_{0}^{T}\int_{\Omega_{f}}\zeta_{N}\boldsymbol{u}_{N}\cdot\tilde{\boldsymbol{v}}_{N}
+120TΓ(𝒖NζN𝒆y)𝒏τΔtωN(τΔt𝒖N𝒗~N)+2ν0TΩf(1+τΔtωNR)𝑫fτΔtωN(𝒖N):𝑫fτΔtωN(𝒗~N)\displaystyle+\frac{1}{2}\int_{0}^{T}\int_{\Gamma}(\boldsymbol{u}_{N}-\zeta_{N}^{*}\boldsymbol{e}_{y})\cdot\boldsymbol{n}^{\tau_{\Delta t}\omega_{N}}(\tau_{\Delta t}\boldsymbol{u}_{N}\cdot\tilde{\boldsymbol{v}}_{N})+2\nu\int_{0}^{T}\int_{\Omega_{f}}\left(1+\frac{\tau_{\Delta t}\omega_{N}}{R}\right)\boldsymbol{D}^{\tau_{\Delta t}\omega_{N}}_{f}(\boldsymbol{u}_{N}):\boldsymbol{D}^{\tau_{\Delta t}\omega_{N}}_{f}(\tilde{\boldsymbol{v}}_{N})
+0TΓ(12𝒖NτΔt𝒖NpN)(𝝍𝒗~N)𝒏τΔtωN+β𝒥ΓτΔtωN0TΓ(ζN𝒆y𝒖N)𝝉τΔtωN(𝝍𝒗~N)𝝉τΔtωN\displaystyle+\int_{0}^{T}\int_{\Gamma}\left(\frac{1}{2}\boldsymbol{u}_{N}\cdot\tau_{\Delta t}\boldsymbol{u}_{N}-p_{N}\right)(\boldsymbol{\psi}-\tilde{\boldsymbol{v}}_{N})\cdot\boldsymbol{n}^{\tau_{\Delta t}\omega_{N}}+\frac{\beta}{\mathcal{J}^{\tau_{\Delta t}\omega_{N}}_{\Gamma}}\int_{0}^{T}\int_{\Gamma}(\zeta_{N}^{*}\boldsymbol{e}_{y}-\boldsymbol{u}_{N})\cdot\boldsymbol{\tau}^{\tau_{\Delta t}\omega_{N}}(\boldsymbol{\psi}-\tilde{\boldsymbol{v}}_{N})\cdot\boldsymbol{\tau}^{\tau_{\Delta t}\omega_{N}}
+ρb0TΩb(𝝃NτΔt𝝃NΔt)𝝍+ρp0TΓtζ¯Nφ+2μe0TΩb𝑫(𝜼N):𝑫(𝝍)\displaystyle+\rho_{b}\int_{0}^{T}\int_{\Omega_{b}}\left(\frac{\boldsymbol{\xi}_{N}-\tau_{\Delta t}\boldsymbol{\xi}_{N}}{\Delta t}\right)\cdot\boldsymbol{\psi}+\rho_{p}\int_{0}^{T}\int_{\Gamma}\partial_{t}\overline{\zeta}_{N}\cdot\varphi+2\mu_{e}\int_{0}^{T}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}_{N}):\boldsymbol{D}(\boldsymbol{\psi})
+λe0TΩb(𝜼N)(𝝍)+2μv0TΩb𝑫(𝝃N):𝑫(𝝍)+λv0TΩb(𝝃N)(𝝍)\displaystyle+\lambda_{e}\int_{0}^{T}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}_{N})(\nabla\cdot\boldsymbol{\psi})+2\mu_{v}\int_{0}^{T}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\xi}_{N}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{v}\int_{0}^{T}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\xi}_{N})(\nabla\cdot\boldsymbol{\psi})
α0TΩb𝒥b(τΔtηN)δpNb(τΔtηN)δ𝝍+c00TΩbtp¯Nrα0TΩb𝒥b(τΔtηN)δ𝝃Nb(τΔtηN)δr\displaystyle-\alpha\int_{0}^{T}\int_{\Omega_{b}}\mathcal{J}^{(\tau_{\Delta t}\eta_{N})^{\delta}}_{b}p_{N}\nabla^{(\tau_{\Delta t}\eta_{N})^{\delta}}_{b}\cdot\boldsymbol{\psi}+c_{0}\int_{0}^{T}\int_{\Omega_{b}}\partial_{t}\overline{p}_{N}\cdot r-\alpha\int_{0}^{T}\int_{\Omega_{b}}\mathcal{J}^{(\tau_{\Delta t}\eta_{N})^{\delta}}_{b}\boldsymbol{\xi}_{N}\cdot\nabla^{(\tau_{\Delta t}\eta_{N})^{\delta}}_{b}r
α0TΓ(ζN𝒆y𝒏(τΔtωN)δ)r+κ0TΩb𝒥b(τΔtηN)δb(τΔtηN)δpNb(τΔtηN)δr\displaystyle-\alpha\int_{0}^{T}\int_{\Gamma}(\zeta_{N}^{*}\boldsymbol{e}_{y}\cdot\boldsymbol{n}^{(\tau_{\Delta t}\omega_{N})^{\delta}})r+\kappa\int_{0}^{T}\int_{\Omega_{b}}\mathcal{J}^{(\tau_{\Delta t}\eta_{N})^{\delta}}_{b}\nabla^{(\tau_{\Delta t}\eta_{N})^{\delta}}_{b}p_{N}\cdot\nabla^{(\tau_{\Delta t}\eta_{N})^{\delta}}_{b}r
0TΓ[(𝒖NζN𝒆y)𝒏τΔtωN]r+0TΓΔωNΔφ=0.\displaystyle-\int_{0}^{T}\int_{\Gamma}[(\boldsymbol{u}_{N}-\zeta_{N}^{*}\boldsymbol{e}_{y})\cdot\boldsymbol{n}^{\tau_{\Delta t}\omega_{N}}]r+\int_{0}^{T}\int_{\Gamma}\Delta\omega_{N}\cdot\Delta\varphi=0.

For additional details about the limit passage in the semidiscrete formulation (61), we refer the reader to the discussion in Section 7.2 of [52] and Section 2.7.2 of [68].

Using the strong convergence results established above, combined with the previously established weak convergence results in Proposition 7.1, we can pass to the limit in all of the terms in the semidiscrete weak formulation except those involving time derivatives. However, we can handle these by a discrete integration by parts. For example, for the first integral, we can use a discrete integration by parts to obtain:

0TΩf(1+τΔtωNR)t𝒖¯N𝒗~N\displaystyle\int_{0}^{T}\int_{\Omega_{f}}\left(1+\frac{\tau_{\Delta t}\omega_{N}}{R}\right)\partial_{t}\overline{\boldsymbol{u}}_{N}\cdot\tilde{\boldsymbol{v}}_{N}
0TΩf(1+ωR)𝒖t𝒗~1R0TΩf(tω)𝒖𝒗~Ωf(1+ω0R)𝒖(0)𝒗~(0),\displaystyle\to-\int_{0}^{T}\int_{\Omega_{f}}\left(1+\frac{\omega}{R}\right)\boldsymbol{u}\cdot\partial_{t}\tilde{\boldsymbol{v}}-\frac{1}{R}\int_{0}^{T}\int_{\Omega_{f}}(\partial_{t}\omega)\boldsymbol{u}\cdot\tilde{\boldsymbol{v}}-\int_{\Omega_{f}}\left(1+\frac{\omega_{0}}{R}\right)\boldsymbol{u}(0)\cdot\tilde{\boldsymbol{v}}(0),

where 𝒗~N=𝒗ΦfτΔtωN\tilde{\boldsymbol{v}}_{N}=\boldsymbol{v}\circ\Phi^{\tau_{\Delta t}\omega_{N}}_{f} and 𝒗~=𝒗Φfω\tilde{\boldsymbol{v}}=\boldsymbol{v}\circ\Phi^{\omega}_{f} for 𝒗𝒳\boldsymbol{v}\in\mathcal{X}. See for example pg. 79-81 in [68].

The limiting weak formulation holds for all velocity test functions in the smooth test space, which can be extended to the general test space 𝒱testω\mathcal{V}^{\omega}_{\text{test}} defined in (39) by using a density argument. Therefore, we have shown that the approximate weak solutions converge, up to a subsequence, to a weak solution to the regularized problem, as stated in Theorem 5.1.

This completes the main result of this manuscript, stated in Theorem 5.1, providing existence of a weak solution to the nonlinearly coupled, regularized fluid-poroviscoelastic structure interaction problem, given in Definition 5.5.

We conclude this section by making the important observation that the weak solution that we have constructed to the regularized FPSI problem satisfies the desired energy estimate. This will be important for showing weak-classical consistency in the next section, and can be shown easily by using the discrete energy estimate for the approximate solutions.

Proposition 9.3.

(Energy estimate for the limiting solution to the regularized problem.) The weak solution (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega) constructed from the splitting scheme as the limit of approximate solutions satisfies the following energy estimate for almost every t[0,T]t\in[0,T]:

12Ωf(t)|𝒖|2+12ρbΩb|𝝃|2+12c0Ωb|p|2+μeΩb|𝑫(𝜼)|2\displaystyle\frac{1}{2}\int_{\Omega_{f}(t)}|\boldsymbol{u}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\xi}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta})|^{2}
+12λeΩb|𝜼|2+12ρpΓ|ζ|2+12Γ|Δω|2+2ν0tΩf(s)|𝑫(𝒖)|2\displaystyle+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta|^{2}+\frac{1}{2}\int_{\Gamma}|\Delta\omega|^{2}+2\nu\int_{0}^{t}\int_{\Omega_{f}(s)}|\boldsymbol{D}(\boldsymbol{u})|^{2}
+2μv0tΩb|𝑫(𝝃)|2+λv0tΩb|𝝃|2+κ0tΩbδ(s)|p|2+β0tΓ(s)|(ζ𝒆y𝒖)𝝉)|2E0,\displaystyle+2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\xi})|^{2}+\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\xi}|^{2}+\kappa\int_{0}^{t}\int_{\Omega^{\delta}_{b}(s)}|\nabla p|^{2}+\beta\int_{0}^{t}\int_{\Gamma(s)}|(\zeta\boldsymbol{e}_{y}-\boldsymbol{u})\cdot\boldsymbol{\tau})|^{2}\leq E_{0}, (101)

where E0E_{0} is the initial energy of the problem.

Proof.

The approximate solutions (𝒖N,𝜼N,pN,ωN)(\boldsymbol{u}_{N},\boldsymbol{\eta}_{N},p_{N},\omega_{N}) satisfy the following energy inequality:

12Ωf,N(t)|𝒖N|2+12ρbΩb|𝝃N|2+12c0Ωb|pN|2+μeΩb|𝑫(𝜼N)|2\displaystyle\frac{1}{2}\int_{\Omega_{f,N}(t)}|\boldsymbol{u}_{N}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\xi}_{N}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p_{N}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}_{N})|^{2}
+12λeΩb|𝜼N|2+12ρpΓ|ζN|2+12Γ|ΔωN|2+2ν0tΩf,N(s)|𝑫(𝒖N)|2\displaystyle+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}_{N}|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta_{N}|^{2}+\frac{1}{2}\int_{\Gamma}|\Delta\omega_{N}|^{2}+2\nu\int_{0}^{t}\int_{\Omega_{f,N}(s)}|\boldsymbol{D}(\boldsymbol{u}_{N})|^{2}
+2μv0tΩb|𝑫(𝝃N)|2+λv0tΩb|𝝃N|2+κ0tΩb,Nδ(s)|pN|2+β0tΓ(s)|(ζN𝒆y𝒖N)𝝉|2E0.\displaystyle+2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\xi}_{N})|^{2}+\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\xi}_{N}|^{2}+\kappa\int_{0}^{t}\int_{\Omega_{b,N}^{\delta}(s)}|\nabla p_{N}|^{2}+\beta\int_{0}^{t}\int_{\Gamma(s)}|(\zeta_{N}^{*}\boldsymbol{e}_{y}-\boldsymbol{u}_{N})\cdot\boldsymbol{\tau}|^{2}\leq E_{0}.

By using the weak and weak-star convergences of the approximate solutions, stated in Proposition 7.1 and the weak lower semicontinuity property of the norms, we can pass to the limit in the energy inequality, one recovers the energy inequality (9.3). ∎

10 Weak-classical consistency

We have now shown the existence of weak solutions to the regularized FPSI problem (40). However, it is not clear that the solutions to this regularized problem are physically relevant, since the regularized weak formulation is not equivalent to the original weak formulation without the regularization. However, we will demonstrate the following weak-classical consistency result: given a spatially and temporally smooth solution (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega) to the FPSI problem, then the weak solutions to the regularized problem with regularization parameter δ\delta, which we will denote by (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}), converge to the smooth solution as δ0\delta\to 0.

10.1 Notation

Since we will have to use spatial convolution of the solution to the regularized problem (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}), and spatial convolution of the smooth solution (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega), we introduce the following notation to avoid additional superscripts involving δ\delta.

  1. 1.

    Recall that (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}) denotes the weak solutions to the regularized problem (40);

  2. 2.

    We will use (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega) to denote a spatially and temporally smooth solution to the original FPSI problem (2), (3), (6), (7), (11)-(15);

  3. 3.

    We will use the superscript δ\delta notation

    𝜼δδ=(𝜼δ)δ:=δ2𝜼δσ(x/δ){\boldsymbol{\eta}}_{\delta}^{\delta}=(\boldsymbol{\eta}_{\delta})^{\delta}:=\delta^{-2}\boldsymbol{\eta}_{\delta}*\sigma(x/\delta)

    to denote the spatial convolution defined by (24) of the weak solution to the regularized problem with the smooth convolution δ\delta kernel;

  4. 4.

    Similarly, in the same spirit as in 3., we will use

    𝜼δ=(𝜼)δ=δ2𝜼σ(x/δ){\boldsymbol{\eta}}^{\delta}=(\boldsymbol{\eta})^{\delta}=\delta^{-2}\boldsymbol{\eta}*\sigma(x/\delta) (102)

    to denote the spatial convolution of the classical solution 𝜼\boldsymbol{\eta} with the convolution kernel;

  5. 5.

    We will use superscript δ\delta to denote the physical Biot domain under the regularized displacement:

    Ωb,δδ(t)=(𝑰+𝜼δδ(t))(Ωb).{\Omega}_{b,\delta}^{\delta}(t)=(\boldsymbol{I}+{\boldsymbol{\eta}}_{\delta}^{\delta}(t))(\Omega_{b}). (103)

Weak formulations reformulated. We note that even though the weak formulation (23) and the regularized weak formulation (40) are stated up until a fixed final time TT, we can reformulate the weak formulation for almost every time t[0,T]t\in[0,T] by using a cutoff function (see for example the proof of Lemma A.2 in the appendix where this is done explicitly).

Thus, the classical (temporally and spatially smooth) solution (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega) satisfies the following non-regularized weak formulation for almost all t[0,T]t\in[0,T], for all test functions (𝒗,φ,𝝍,r)𝒱test(\boldsymbol{v},\varphi,\boldsymbol{\psi},r)\in\mathcal{V}_{\text{test}} with the (moving domain) test space 𝒱test\mathcal{V}_{\text{test}} defined in (38):

0tΩf(s)𝒖t𝒗+120tΩf(s)[((𝒖)𝒖)𝒗((𝒖)𝒗)𝒖]+120tΓ1(s)(𝒖𝒏2𝝃1𝒏)𝒖𝒗\displaystyle-\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{u}\cdot\partial_{t}\boldsymbol{v}+\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f}(s)}[((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot\boldsymbol{v}-((\boldsymbol{u}\cdot\nabla)\boldsymbol{v})\cdot\boldsymbol{u}]+\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{1}(s)}(\boldsymbol{u}\cdot\boldsymbol{n}-2\boldsymbol{\xi}_{1}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\boldsymbol{v}
+2ν0tΩf(s)𝑫(𝒖):𝑫(𝒗)+0tΓ1(s)(12|𝒖|2p)(ψnvn)+β0tΓ1(s)(𝝃𝒖)𝒕(ψtvt)\displaystyle+2\nu\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{v})+\int_{0}^{t}\int_{\Gamma_{1}(s)}\left(\frac{1}{2}|\boldsymbol{u}|^{2}-p\right)(\psi_{n}-v_{n})+\beta\int_{0}^{t}\int_{\Gamma_{1}(s)}(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{t}(\psi_{t}-v_{t})
ρp0tΓtωtφ+0tΓΔωΔφρb0tΩbt𝜼1t𝝍+2μe0tΩb𝑫(𝜼1):𝑫(𝝍)\displaystyle-\rho_{p}\int_{0}^{t}\int_{\Gamma}\partial_{t}\omega\cdot\partial_{t}\varphi+\int_{0}^{t}\int_{\Gamma}\Delta\omega\cdot\Delta\varphi-\rho_{b}\int_{0}^{t}\int_{\Omega_{b}}\partial_{t}\boldsymbol{\eta}_{1}\cdot\partial_{t}\boldsymbol{\psi}+2\mu_{e}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}_{1}):\boldsymbol{D}(\boldsymbol{\psi})
+λe0tΩb(𝜼1)(𝝍)+2μv0tΩb𝑫(t𝜼):𝑫(𝝍)+λv0tΩb(t𝜼)(𝝍)\displaystyle+\lambda_{e}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}_{1})(\nabla\cdot\boldsymbol{\psi})+2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\partial_{t}\boldsymbol{\eta}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\partial_{t}\boldsymbol{\eta})(\nabla\cdot\boldsymbol{\psi})
α0tΩb(s)p𝝍c00tΩbptrα0tΩb(s)DDt𝜼rα0tΓ(s)(𝝃𝒏)r\displaystyle-\alpha\int_{0}^{t}\int_{\Omega_{b}(s)}p\nabla\cdot\boldsymbol{\psi}-c_{0}\int_{0}^{t}\int_{\Omega_{b}}p\partial_{t}r-{{\alpha\int_{0}^{t}\int_{\Omega_{b}(s)}\frac{D}{Dt}\boldsymbol{\eta}\cdot\nabla r-\alpha\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})r}}
+κ0tΩb(s)pr0tΓ1(s)((𝒖𝝃)𝒏)r\displaystyle+\kappa\int_{0}^{t}\int_{\Omega_{b}(s)}\nabla p\cdot\nabla r-{{\int_{0}^{t}\int_{\Gamma_{1}(s)}((\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n})r}}
=Ωf(t)𝒖(t)𝒗(t)ρpΓζ(t)𝝍(t)ρbΩb𝝃(t)𝝍(t)c0Ωbp(t)r(t)\displaystyle=-\int_{\Omega_{f}(t)}\boldsymbol{u}(t)\cdot\boldsymbol{v}(t)-\rho_{p}\int_{\Gamma}\zeta(t)\cdot\boldsymbol{\psi}(t)-\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}(t)\cdot\boldsymbol{\psi}(t)-c_{0}\int_{\Omega_{b}}p(t)\cdot r(t)
+Ωf(0)𝒖0𝒗(0)+ρpΓβ0𝝍(0)+ρbΩb𝝃0𝝍(0)+c0Ωbp0r(0).\displaystyle+\int_{\Omega_{f}(0)}\boldsymbol{u}_{0}\cdot\boldsymbol{v}(0)+\rho_{p}\int_{\Gamma}\beta_{0}\cdot\boldsymbol{\psi}(0)+\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}_{0}\cdot\boldsymbol{\psi}(0)+c_{0}\int_{\Omega_{b}}p_{0}\cdot r(0). (104)

Similarly, the solution to the regularized FPSI problem (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}) satisfies the following regularized weak formulation for every test function (𝒗,φ,𝝍,r)𝒱test(\boldsymbol{v},\varphi,\boldsymbol{\psi},r)\in\mathcal{V}_{\text{test}}, and for almost every t[0,Tδ]t\in[0,T_{\delta}] where the final time TδT_{\delta} potentially depends on δ\delta:

0tΩf,δ(s)𝒖δt𝒗+120tΩf,δ(s)[((𝒖δ)𝒖δ)𝒗((𝒖δ)𝒗)𝒖δ]\displaystyle-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\boldsymbol{u}_{\delta}\cdot\partial_{t}\boldsymbol{v}+\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta})\cdot\boldsymbol{v}-((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{v})\cdot\boldsymbol{u}_{\delta}]
+120tΓδ(s)(𝒖δ𝒏2𝝃δ𝒏)𝒖δ𝒗+2ν0tΩf,δ(s)𝑫(𝒖δ):𝑫(𝒗)\displaystyle+\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{u}_{\delta}\cdot\boldsymbol{n}-2\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n})\boldsymbol{u}_{\delta}\cdot\boldsymbol{v}+2\nu\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\boldsymbol{D}(\boldsymbol{u}_{\delta}):\boldsymbol{D}(\boldsymbol{v})
+0tΓδ(s)(12|𝒖δ|2pδ)(ψnvn)+β0tΓδ(s)(𝝃δ𝒖δ)𝒕(ψtvt)\displaystyle+\int_{0}^{t}\int_{\Gamma_{\delta}(s)}\left(\frac{1}{2}|\boldsymbol{u}_{\delta}|^{2}-p_{\delta}\right)(\psi_{n}-v_{n})+\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{t}(\psi_{t}-v_{t})
ρp0tΓtωδtφ+0tΓΔωδΔφρb0tΩbt𝜼δt𝝍+2μe0tΩb𝑫(𝜼δ):𝑫(𝝍)\displaystyle-\rho_{p}\int_{0}^{t}\int_{\Gamma}\partial_{t}\omega_{\delta}\cdot\partial_{t}\varphi+\int_{0}^{t}\int_{\Gamma}\Delta\omega_{\delta}\cdot\Delta\varphi-\rho_{b}\int_{0}^{t}\int_{\Omega_{b}}\partial_{t}\boldsymbol{\eta}_{\delta}\cdot\partial_{t}\boldsymbol{\psi}+2\mu_{e}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}_{\delta}):\boldsymbol{D}(\boldsymbol{\psi})
+λe0tΩb(𝜼δ)(𝝍)+2μv0tΩb𝑫(t𝜼δ):𝑫(𝝍)+λv0tΩb(t𝜼δ)(𝝍)\displaystyle+\lambda_{e}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}_{\delta})(\nabla\cdot\boldsymbol{\psi})+2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\partial_{t}\boldsymbol{\eta}_{\delta}):\boldsymbol{D}(\boldsymbol{\psi})+\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\partial_{t}\boldsymbol{\eta}_{\delta})(\nabla\cdot\boldsymbol{\psi})
α0tΩb,δδ(s)pδ𝝍c00tΩbpδtrα0tΩb,δδ(s)DδDt𝜼δrα0tΓδ(s)(𝝃δ𝒏)r\displaystyle-\alpha\int_{0}^{t}\int_{{\Omega}_{b,\delta}^{\delta}(s)}p_{\delta}\nabla\cdot\boldsymbol{\psi}-c_{0}\int_{0}^{t}\int_{\Omega_{b}}p_{\delta}\partial_{t}r-\alpha\int_{0}^{t}\int_{{\Omega}_{b,\delta}^{\delta}(s)}\frac{D^{\delta}}{Dt}\boldsymbol{\eta}_{\delta}\cdot\nabla r-\alpha\int_{0}^{t}\int_{{\Gamma}_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n})r
+κ0tΩb,δδ(s)pδr0tΓδ(s)((𝒖δ𝝃δ)𝒏)r\displaystyle+\kappa\int_{0}^{t}\int_{{\Omega}_{b,\delta}^{\delta}(s)}\nabla p_{\delta}\cdot\nabla r-\int_{0}^{t}\int_{\Gamma_{\delta}(s)}((\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n})r
=Ωf,δ(t)𝒖δ(t)𝒗(t)ρpΓζδ(t)φ(t)ρbΩb𝝃δ(t)𝝍(t)c0Ωbpδ(t)r(t)\displaystyle=-\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\boldsymbol{v}(t)-\rho_{p}\int_{\Gamma}\zeta_{\delta}(t)\cdot\varphi(t)-\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}_{\delta}(t)\cdot\boldsymbol{\psi}(t)-c_{0}\int_{\Omega_{b}}p_{\delta}(t)\cdot r(t)
+Ωf(0)𝒖0𝒗(0)+ρpΓβ0φ(0)+ρbΩb𝝃0𝝍(0)+c0Ωbp0r(0),\displaystyle+\int_{\Omega_{f}(0)}\boldsymbol{u}_{0}\cdot\boldsymbol{v}(0)+\rho_{p}\int_{\Gamma}\beta_{0}\cdot\varphi(0)+\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}_{0}\cdot\boldsymbol{\psi}(0)+c_{0}\int_{\Omega_{b}}p_{0}\cdot r(0), (105)

where DδDt\frac{D^{\delta}}{Dt} is the material derivative with respect to the regularized Biot displacement. We remark that while our existence proof in the previous sections holds for both a purely elastic and viscoelastic Biot medium, our weak-classical consistency result will hold in the specific case of a Biot poroviscoelastic medium so that the viscoelasticity parameters μv\mu_{v} and λv\lambda_{v} are strictly positive, and hence, the plate velocity ζδ𝒆y\zeta_{\delta}\boldsymbol{e}_{y} in the weak formulation is equivalently the trace of the Biot medium velocity 𝝃δL2(0,T;H1(Ωb))\boldsymbol{\xi}_{\delta}\in L^{2}(0,T;H^{1}(\Omega_{b})) along Γ\Gamma.

10.2 Statement of the result

In the remainder of the manuscript, we will prove the weak-classical consistency result. Before stating the result, we need to introduce some additional notation. Namely, to prove the weak-classical consistency, we will subtract the weak formulations for the two solutions 𝒖\boldsymbol{u} and 𝒖δ\boldsymbol{u}_{\delta} and test formally with the difference of the two solutions 𝒗=𝒖𝒖δ\boldsymbol{v}=\boldsymbol{u}-\boldsymbol{u}_{\delta}. However, the functions 𝒖\boldsymbol{u} and 𝒖δ\boldsymbol{u}_{\delta} are defined on different domains, and the test functions on the physical domains are required to be divergence-free, so we need to be able to transfer 𝒖δ\boldsymbol{u}_{\delta} to the physical domain for 𝒖\boldsymbol{u} in a way that preserves the divergence-free condition. To do this, we emphasize that we cannot use the usual ALE mapping Φfω:ΩfΩf(t)\Phi^{\omega}_{f}:\Omega_{f}\to\Omega_{f}(t) defined in (16). This is because given two structure displacements ω1\omega_{1} and ω2\omega_{2} which define two respective fluid domains Ωfω1\Omega^{\omega_{1}}_{f} and Ωfω2\Omega^{\omega_{2}}_{f} and a divergence-free function 𝒖1\boldsymbol{u}_{1} on Ωfω1\Omega^{\omega_{1}}_{f}, the function 𝒖𝚽^fω1(𝚽^fω2)1\boldsymbol{u}\circ\hat{\boldsymbol{\Phi}}^{\omega_{1}}_{f}\circ(\hat{\boldsymbol{\Phi}}^{\omega_{2}}_{f})^{-1} transformed via the ALE mapping is not necessarily divergence-free on Ωfω2\Omega^{\omega_{2}}_{f}. Therefore, we will have to use a different transformation to bring a divergence-free function defined on one fluid domain to a divergence-free function defined on another fluid domain.

For this purpose consider the two fluid domains

Ωf(t)={(x,y)2:0xL,Ryω(t,x)},\Omega_{f}(t)=\{(x,y)\in\mathbb{R}^{2}:0\leq x\leq L,-R\leq y\leq\omega(t,x)\},
Ωf,δ(t)={(x,y)2:0xL,Ryωδ(t,x)},\Omega_{f,\delta}(t)=\{(x,y)\in\mathbb{R}^{2}:0\leq x\leq L,-R\leq y\leq\omega_{\delta}(t,x)\},

that are associated to the plate displacements ω\omega and ωδ\omega_{\delta}.

We define a map between Ωf(t)\Omega_{f}(t) and Ωf,δ(t)\Omega_{f,\delta}(t), and a transformation that sends functions on one domain to functions on the other domain as follows. Let ψδ(t):Ωf,δ(t)Ωf(t)\psi_{\delta}(t):\Omega_{f,\delta}(t)\to\Omega_{f}(t) be the mapping defined by

ψδ(t,x,y)=(t,x,γδ(t,x)(R+y)R),whereγδ(t,x)=R+ω(t,x)R+ωδ(t,x).\psi_{\delta}(t,x,y)=(t,x,\gamma_{\delta}(t,x)(R+y)-R),\ {\rm{where}}\ \gamma_{\delta}(t,x)=\frac{R+\omega(t,x)}{R+\omega_{\delta}(t,x)}. (106)

This mapping, unfortunately, does not preserve the divergence free condition. However, if we calculate the gradient of the composite mapped function we get

(𝒖ψδ)=[(𝒖)ψδ]Jδ\nabla(\boldsymbol{u}\circ\psi_{\delta})=[(\nabla\boldsymbol{u})\circ\psi_{\delta}]J_{\delta} (107)

where

Jδ(t,x,y)=(10(R+y)xγδ(t,x)γδ(t,x)).J_{\delta}(t,x,y)=\begin{pmatrix}1&0\\ (R+y)\partial_{x}\gamma_{\delta}(t,x)&\gamma_{\delta}(t,x)\\ \end{pmatrix}. (108)

Similarly, for the regularized problem we define

J~δ=Jδψδ1=(10(R+y)γδ1xγδ(t,x)γδ(t,x)).\tilde{J}_{\delta}=J_{\delta}\circ\psi_{\delta}^{-1}=\begin{pmatrix}1&0\\ (R+y)\gamma_{\delta}^{-1}\partial_{x}\gamma_{\delta}(t,x)&\gamma_{\delta}(t,x)\\ \end{pmatrix}. (109)

These Jacobian matrices will now be used to define the transformations that map divergence free functions to divergence free functions.

Definition 10.1.

Part I: Given a divergence-free function 𝒖\boldsymbol{u} on Ωf(t)\Omega_{f}(t), the following transformation ^:𝒖𝒖^\widehat{\phantom{u}}:{\boldsymbol{u}}\mapsto\widehat{\boldsymbol{u}} maps 𝒖{\boldsymbol{u}} to a divergence free function 𝒖^\widehat{\boldsymbol{u}} on Ωf,δ(t)\Omega_{f,\delta}(t):

𝒖^=γδJδ1(𝒖ψδ).\widehat{\boldsymbol{u}}=\gamma_{\delta}J^{-1}_{\delta}\cdot(\boldsymbol{u}\circ\psi_{\delta}). (110)

Part II: Given a divergence-free function 𝒖δ\boldsymbol{u}_{\delta} on Ωf,δ(t)\Omega_{f,\delta}(t) the following transformation ˇ:𝒖δ𝒖ˇδ\widecheck{\phantom{u}}:\boldsymbol{u}_{\delta}\mapsto\widecheck{\boldsymbol{u}}_{\delta} maps 𝒖δ\boldsymbol{u}_{\delta} to a divergence free function 𝒖ˇδ\widecheck{\boldsymbol{u}}_{\delta} on Ωf(t)\Omega_{f}(t):

𝒖ˇδ=γδ1J~δ(𝒖δψδ1).\widecheck{\boldsymbol{u}}_{\delta}=\gamma_{\delta}^{-1}\tilde{J}_{\delta}\cdot(\boldsymbol{u}_{\delta}\circ\psi_{\delta}^{-1}). (111)
Remark 10.1.

Both transformations preserve the trace of functions along Γ\Gamma.

Note that even though the definition of 𝒖^\widehat{\boldsymbol{u}} depends on δ\delta, we will not explicitly notate this dependence, as δ\delta will be clear from the context. We now state the weak-classical consistency result.

Theorem 10.1.

(Weak-classical consistency) Let (𝜼0,𝝃0,ω0,ζ0,p0,𝒖0)(\boldsymbol{\eta}_{0},\boldsymbol{\xi}_{0},\omega_{0},\zeta_{0},p_{0},\boldsymbol{u}_{0}) be smooth initial data for the nonlinearly coupled FPSI problem (2), (3), (6), (7), (11)-(15). Suppose (𝜼,ω,p,𝒖)(\boldsymbol{\eta},\omega,p,\boldsymbol{u}) is a classical (temporally and spatially smooth) solution to this FPSI problem on the time interval [0,T][0,T]. Let (𝜼δ,ωδ,pδ,𝒖δ)(\boldsymbol{\eta}_{\delta},\omega_{\delta},p_{\delta},\boldsymbol{u}_{\delta}) denote the weak solution to the regularized FPSI problem (40) with regularity parameter δ\delta.

Then the following holds true:

  1. 1.

    (𝜼δ,ωδ,pδ,𝒖δ)(\boldsymbol{\eta}_{\delta},\omega_{\delta},p_{\delta},\boldsymbol{u}_{\delta}) is defined on the time interval [0,T][0,T] for all δ>0\delta>0, where the final time TT is independent of δ\delta;

  2. 2.

    The energy norm of the difference between the two solutions Eδ(t)E_{\delta}(t) converges to zero as δ0\delta\to 0, for all t[0,T]t\in[0,T], where

    Eδ(t)\displaystyle E_{\delta}(t) :=(𝒖^𝒖δ)(t)L2(Ωf,δ(t))2+0t𝑫(𝒖^𝒖δ)(s)L2(Ωf,δ(s))2𝑑s\displaystyle:=||(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})(t)||^{2}_{L^{2}(\Omega_{f,\delta}(t))}+\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})(s)||^{2}_{L^{2}(\Omega_{f,\delta}(s))}ds
    +(𝝃𝝃δ)(t)L2(Γ)2+(ωωδ)(t)H2(Γ)2+(𝝃𝝃δ)(t)L2(Ωb)2\displaystyle+||(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})(t)||_{L^{2}(\Gamma)}^{2}+||(\omega-\omega_{\delta})(t)||^{2}_{H^{2}(\Gamma)}+||(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})(t)||^{2}_{L^{2}(\Omega_{b})}
    +𝑫(𝜼𝜼δ)(t)L2(Ωb)2+((𝜼𝜼δ))(t)L2(Ωb)2+0t𝑫(𝝃𝝃δ)(s)L2(Ωb)2𝑑s\displaystyle+||\boldsymbol{D}(\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta})(t)||^{2}_{L^{2}(\Omega_{b})}+||(\nabla\cdot(\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta}))(t)||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\boldsymbol{D}(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})(s)||^{2}_{L^{2}(\Omega_{b})}ds
    +0t(𝝃𝝃δ)(s)L2(Ωb)2𝑑s+(ppδ)(t)L2(Ωb)2+0t(ppδ)(s)L2(Ωb,δδ(s))2𝑑s.\displaystyle+\int_{0}^{t}||\nabla\cdot(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})(s)||_{L^{2}(\Omega_{b})}^{2}ds+||(p-p_{\delta})(t)||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\nabla(p-p_{\delta})(s)||^{2}_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}ds. (112)

Preview of the main steps of the proof of weak-classical consistency. The proof is based on Gronwall’s inequality for Eδ(t)E_{\delta}(t). However, there are several obstacles to applying Gronwall’s inequality due to the fact that we are working on a moving domain problem. We summarize those main obstacles, and the main ideas behind their resolution here.

The main idea is to estimate the energy difference between (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega) and (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}), defined in (2) and obtain an estimate for Eδ(t)E_{\delta}(t) in terms of Eδ(0)E_{\delta}(0), the integral of Eδ(s)E_{\delta}(s) for times s[0,t]s\in[0,t], and other terms that have sufficiently strong convergence in δ\delta as δ0\delta\to 0:

Eδ(t)C(0t𝜼𝜼δL2(Ωb)2𝑑s+0tEδ(s)𝑑s)E_{\delta}(t)\leq C\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||^{2}_{L^{2}(\Omega_{b})}ds+\int_{0}^{t}E_{\delta}(s)ds\right) (113)

and then apply Gronwall’s inequality to obtain

Eδ(t)Cδ3eCt,E_{\delta}(t)\leq C\delta^{3}e^{Ct}, (114)

where CC is independent of δ\delta, and conclude that Eδ(t)0E_{\delta}(t)\to 0 as δ0\delta\to 0. We remark that the factor of δ3\delta^{3} appearing in the Gronwall estimate comes from an estimate of the convergence rate of the spatial convolution 𝜼δ\boldsymbol{\eta}^{\delta} to 𝜼\boldsymbol{\eta} in H1(Ωb)H^{1}(\Omega_{b}), which we establish in the upcoming Lemma 10.2.

To do this, we will test the weak formulations for 𝒖\boldsymbol{u} and 𝒖δ\boldsymbol{u}_{\delta} with appropriate test functions and use the energy inequality (9.3) from Proposition 9.3. More precisely, the main steps in the proof are:

  1. 1.

    Test the non-regularized weak formulation (10.1) for the classical solution (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega) with the “difference” of (𝒖,t𝜼,p,tω)(\boldsymbol{u},\partial_{t}\boldsymbol{\eta},p,\partial_{t}\omega) and (𝒖δ,t𝜼δ,pδ,tωδ)(\boldsymbol{u}_{\delta},\partial_{t}\boldsymbol{\eta}_{\delta},p_{\delta},\partial_{t}\omega_{\delta}), where the notion of the difference between these two solutions will be made precise in Section 10.3;

  2. 2.

    Test the regularized weak formulation (10.1) for (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}) with (𝒖,t𝜼,p,tω)(\boldsymbol{u},\partial_{t}\boldsymbol{\eta},p,\partial_{t}\omega);

  3. 3.

    Rewrite the energy inequality (9.3) for (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}) so that it parallels the terms in the weak formulation (10.1);

  4. 4.

    Combine the equations from Step 1, Step 2, and Step 3. This will give us an expression that we can analyze term by term in order to obtain estimate (113) for the energy difference Eδ(t)E_{\delta}(t). Details will be presented in Section 10.3;

  5. 5.

    Now, we have that the inequality (113) and the resulting Gronwall estimate (114) are proven locally in time, namely, on the interval [0,Tδ][0,T_{\delta}] along which certain boundedness assumptions on the Lagrangian map hold for the solution of the regularized problem. However, we want the estimate (114) to hold along the entire time interval [0,T][0,T], along which the classical solution is defined. Hence, we will use a bootstrap argument on boundedness assumptions of the Lagrangian map in order to propagate the estimate (113) to the entire time interval [0,T][0,T], see Section 10.4.

  6. 6.

    Apply Gronwall’s inequality to (113) holding on [0,T][0,T] to obtain the following bound for Eδ(t)E_{\delta}(t):

    Eδ(t)Cδ3eCt,E_{\delta}(t)\leq C\delta^{3}e^{Ct},

    where CC is independent of δ\delta, and conclude that Eδ(t)0E_{\delta}(t)\to 0 as δ0\delta\to 0.

Before we start with the proof of weak-classical consistency, we emphasize that there are two main mathematical difficulties that need to be addressed in the proof:

  1. 1.

    In step 1 above, we want to test (10.1) with the difference of (𝒖,t𝜼,p,tω)(\boldsymbol{u},\partial_{t}\boldsymbol{\eta},p,\partial_{t}\omega) and (𝒖δ,t𝜼δ,pδ,tωδ)(\boldsymbol{u}_{\delta},\partial_{t}\boldsymbol{\eta}_{\delta},p_{\delta},\partial_{t}\omega_{\delta}). This is formal because the test functions in 𝒱test\mathcal{V}_{\text{test}}, defined in (38), must be continuously differentiable in time, and furthermore, for the fluid velocities, the difference between 𝒖\boldsymbol{u} and 𝒖δ\boldsymbol{u}_{\delta} does not make sense, since these functions are defined on different fluid domains. Thus, we must carefully define which test functions we will use. This is addressed at the beginning of Section 10.3 below.

  2. 2.

    As mentioned in step 5 above, the regularized weak formulation involves integrals on the physical time-dependent Biot domain Ωb,δδ(t){\Omega}_{b,\delta}^{\delta}(t), which give an extra factor of det(𝑰+𝜼δδ)\det(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}) in the integrand from the Jacobian, when the integrals are transferred to the fixed reference Biot domain Ωb\Omega_{b}. This factor cannot be estimated in the finite energy space where 𝜼δ\boldsymbol{\eta}_{\delta} is only bounded uniformly in δ\delta in the function space L(0,T;H1(Ωb))L^{\infty}(0,T;H^{1}(\Omega_{b})). To obtain pointwise estimates of this term that hold on the time interval [0,T][0,T], where TT is independent of δ\delta, we need to use a bootstrap argument to get from the local pointwise estimates on [0,Tδ][0,T_{\delta}], where TδT_{\delta} depends on δ\delta, to the global, uniform estimates on [0,T][0,T]. This is addressed in Section 10.4 below.

10.3 Gronwall’s Inequality

We show that the following Gronwall’s inequality holds for almost all t[0,Tδ]t\in[0,T_{\delta}], where TδT_{\delta} depends on δ\delta. Later on we will use a bootstrap argument to show that the weak-classical consistency holds uniformly, on the entire interval [0,T][0,T] on which the classical solution exists.

Lemma 10.1.

Gronwall’s estimate. Suppose that there exist constants c>0c>0 and C>0C>0 which are independent of δ\delta such that the following estimates hold for almost all t[0,Tδ]t\in[0,T_{\delta}]:

det\displaystyle\det (𝑰+𝜼δδ)c>0,\displaystyle(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta})\geq c>0, (115)
0<c\displaystyle 0<c\leq |𝑰+𝜼δδ|C, pointwise in Ωb¯,\displaystyle|\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}|\leq C,\qquad\text{ pointwise in }\overline{\Omega_{b}}, (116)
|𝜼δδ|C, pointwise in Ωb¯,\displaystyle|\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}|\leq C,\qquad\text{ pointwise in }\overline{\Omega_{b}}, (117)

where the final time TδT_{\delta} potentially depends on δ\delta. Furthermore, let 𝜼\boldsymbol{\eta} and 𝜼δ{\boldsymbol{\eta}}^{\delta} denote the smooth solution and its regularization, defined on [0,T][0,T], and EδE_{\delta} be the energy norm difference (2). Then the following inequality hold:

Eδ(t)C(0t𝜼𝜼δL2(Ωb)2𝑑s+0tEδ(s)𝑑s),E_{\delta}(t)\leq C\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||^{2}_{L^{2}(\Omega_{b})}ds+\int_{0}^{t}E_{\delta}({{s}})ds\right), (118)

where Eδ(t)E_{\delta}(t) is defined by (2). Furthermore,

Eδ(t)Cδ3eCt.E_{\delta}(t)\leq C\delta^{3}e^{Ct}.

To prove Gronwall’s inequality, we want to test the non-regularized weak formulation formally with the difference between (𝒖,t𝜼,p,tω)(\boldsymbol{u},\partial_{t}\boldsymbol{\eta},p,\partial_{t}\omega) and (𝒖δ,t𝜼δ,pδ,tωδ)(\boldsymbol{u}_{\delta},\partial_{t}\boldsymbol{\eta}_{\delta},p_{\delta},\partial_{t}\omega_{\delta}). However, there are two reasons why this is not rigorously justified. First, t𝜼t𝜼δ\partial_{t}\boldsymbol{\eta}-\partial_{t}\boldsymbol{\eta}_{\delta} is not a continuously differentiable function in time as is required for the test functions, and hence, we will use a convolution in time and pass to the limit as the convolution parameter goes to zero. Second, the fluid velocities give an additional difficulty, as the fluid velocities are defined on time-dependent moving domains. Thus, we must transfer the fluid velocities between different time-dependent domains in order to make sense of the “difference” between 𝒖\boldsymbol{u} and 𝒖δ\boldsymbol{u}_{\delta} as a test function. Furthermore, the way in which we do this transformation and the way in which we perform the convolution in time must both respect the divergence-free nature of the fluid velocity on the time-dependent domain. We will address both of these difficulties as follows.

Construction of appropriate test functions (u,tη,p,tω)(uδ,tηδ,pδ,tωδ)(\boldsymbol{u},\partial_{t}\boldsymbol{\eta},p,\partial_{t}\omega)-(\boldsymbol{u}_{\delta},\partial_{t}\boldsymbol{\eta}_{\delta},p_{\delta},\partial_{t}\omega_{\delta}):

Difficulty 1: Lack of regularity in time. We address the first difficulty by defining a convolution in time. This will allow us to regularize t(𝜼𝜼δ)=𝝃𝝃δ\partial_{t}(\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta})=\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}, ppδp-p_{\delta}, and t(ωωδ)=ζζδ\partial_{t}(\omega-\omega_{\delta})=\zeta-\zeta_{\delta} so that these functions are continuously differentiable in time. Since the classical solution is already continuously differentiable in time, we only need to regularize the weak solutions to the regularized problem. Because these differences are all defined on fixed domains, we can use a standard convolution in time.

Convolution in time. Let j():j(\cdot):\mathbb{R}\to\mathbb{R} be a compactly supported even function with supp(j)[1,1]\text{supp}(j)\subset[-1,1] and j=1\displaystyle\int_{\mathbb{R}}j=1, and we define jν(t)=ν1j(ν1t)j_{\nu}(t)=\nu^{-1}j(\nu^{-1}t), where ν>0\nu>0 is the convolution parameter in time.

Consider ν>0\nu>0. Extend 𝝃δ\boldsymbol{\xi}_{\delta}, pδp_{\delta}, and ζδ\zeta_{\delta} to the larger interval [ν,T+ν][-\nu,T+\nu] by reflecting across t=0t=0 and t=Tt=T. For example, define:

𝝃δ(t)\displaystyle\boldsymbol{\xi}_{\delta}(t) =𝝃δ(t), for t[ν,0],\displaystyle=\boldsymbol{\xi}_{\delta}(-t),\ \text{ for }t\in[-\nu,0],
𝝃δ(t)\displaystyle\boldsymbol{\xi}_{\delta}(t) =𝝃δ(2Tt), for t[T,T+ν].\displaystyle=\boldsymbol{\xi}_{\delta}(2T-t),\ \text{ for }t\in[T,T+\nu].

Convolution in time is then defined by:

(𝝃δ)ν(t)=𝝃δ(t,)jν=𝝃δ(s)jν(ts)𝑑s,fort[0,T].(\boldsymbol{\xi}_{\delta})_{\nu}(t)=\boldsymbol{\xi}_{\delta}(t,\cdot)*j_{\nu}=\int_{\mathbb{R}}\boldsymbol{\xi}_{\delta}(s)j_{\nu}(t-s)ds,\ {\rm{for}}\ t\in[0,T].

The convolutions (pδ)ν(p_{\delta})_{\nu} and (ζδ)ν(\zeta_{\delta})_{\nu} are defined similarly. With these definitions we can now test with 𝝃(𝝃δ)ν\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}, p(pδ)νp-(p_{\delta})_{\nu}, and ζ(ζδ)ν\zeta-(\zeta_{\delta})_{\nu}.

Difficulty 2: Velocities are defined on moving domains. Because the fluid velocities are defined on moving time-dependent domains, we cannot directly apply a convolution in time. We must first be able to transform fluid velocities from one domain to another, while preserving the divergence-free condition, and then convolve in time. The transformation of fluid velocities from one domain to another, while preserving the divergence-free condition, will be performed using the following matrix, which was inspired by the transformation introduced in [41] (see also [63]):

K(s,t,x,y)=(R+ω(s,x)R+ω(t,x)0(R+y)x(R+ω(s,x)R+ω(t,x))1).K(s,t,x,y)=\begin{pmatrix}\frac{R+\omega(s,x)}{R+\omega(t,x)}&0\\ -(R+y)\partial_{x}\left(\frac{R+\omega(s,x)}{R+\omega(t,x)}\right)&1\\ \end{pmatrix}. (119)

This matrix has the following essential property: if 𝒖(x,y)\boldsymbol{u}(x,y) is a divergence-free function on the domain Ωf(s)\Omega_{f}(s) defined by the structure displacement ω(s,x)\omega(s,x), then the function

K(s,t,x,y)𝒖(x,R+ω(s,x)R+ω(t,x)(R+y)R)K(s,t,x,y)\boldsymbol{u}\left(x,\frac{R+\omega(s,x)}{R+\omega(t,x)}(R+y)-R\right)

is a divergence-free vector field on the domain Ωf(t)\Omega_{f}(t) defined by the structure displacement ω(t,x)\omega(t,x). Combined transformation of fluid velocities and convolution in time: We can now use this transformation to convolve in time, as follows. We extend 𝒖δ\boldsymbol{u}_{\delta} to [ν,T+ν][-\nu,T+\nu] by reflection, as above, and define, for t[0,T]t\in[0,T],

(𝒖δ)ν(t)=K2,δ(s,t,x,y)𝒖δ(s,x,R+ωδ(s,x)R+ωδ(t,x)(R+y)R)jν(ts)𝑑s.(\boldsymbol{u}_{\delta})_{\nu}(t)=\int_{\mathbb{R}}K_{2,\delta}(s,t,x,y)\boldsymbol{u}_{\delta}\left(s,x,\frac{R+\omega_{\delta}(s,x)}{R+\omega_{\delta}(t,x)}(R+y)-R\right)j_{\nu}(t-s)ds. (120)

For a divergence-free function 𝒗\boldsymbol{v}, extended as above in time to [ν,T+ν][-\nu,T+\nu], we can define 𝒗ν\boldsymbol{v}_{\nu} on Ωf(t)\Omega_{f}(t) analogously by

𝒗ν(t)=K1(s,t,x,y)𝒗(s,x,R+ω(s,x)R+ω(t,x)(R+y)R)jν(ts)𝑑s.\boldsymbol{v}_{\nu}(t)=\int_{\mathbb{R}}K_{1}(s,t,x,y)\boldsymbol{v}\left(s,x,\frac{R+\omega(s,x)}{R+\omega(t,x)}(R+y)-R\right)j_{\nu}(t-s)ds.

Here, K1(s,t,x,y)K_{1}(s,t,x,y) and K2,δ(s,t,x,y)K_{2,\delta}(s,t,x,y) are defined as K(s,t,x,y)K(s,t,x,y) with the choices of ω=ω\omega=\omega and ω=ωδ\omega=\omega_{\delta} respectively. An example of such a function 𝒗\boldsymbol{v} which will be convenient to consider on Ωf(t)\Omega_{f}(t) is the function 𝒖ˇδ\widecheck{\boldsymbol{u}}_{\delta} defined on Ωf(t)\Omega_{f}(t), which is the function 𝒖δ\boldsymbol{u}_{\delta} defined on Ωf,δ(t)\Omega_{f,\delta}(t) transferred in a divergence-free manner, as described above, onto the domain Ωf(t)\Omega_{f}(t). Specifically,

𝒖ˇδ(t,x,y)=(R+ωδ(t,x)R+ω(t,x)0(R+y)x(R+ωδ(t,x)R+ω(t,x))1)𝒖δ(t,x,R+ωδ(t,x)R+ω(t,x)(R+y)R).\widecheck{\boldsymbol{u}}_{\delta}(t,x,y)=\begin{pmatrix}\frac{R+\omega_{\delta}(t,x)}{R+\omega(t,x)}&0\\ -(R+y)\partial_{x}\left(\frac{R+\omega_{\delta}(t,x)}{R+\omega(t,x)}\right)&1\\ \end{pmatrix}\cdot{{\boldsymbol{u}_{\delta}}}\left({{t,x,\frac{R+\omega_{\delta}(t,x)}{R+\omega(t,x)}(R+y)-R}}\right).

We present the main properties of (𝒖δ)ν(\boldsymbol{u}_{\delta})_{\nu} in the proposition below, which are a specific case of Lemma 2.6 in [63].

Proposition 10.1.

Fix an arbitrary δ>0\delta>0. Given 𝒖δL2(0,T;H1(Ωb(t))\boldsymbol{u}_{\delta}\in L^{2}(0,T;H^{1}(\Omega_{b}(t)) and ω,ωδH02(Γ)\omega,\omega_{\delta}\in H_{0}^{2}(\Gamma), the following properties hold:

  • Divergence-free condition: div[(𝒖δ)ν]=0\text{div}\left[(\boldsymbol{u}_{\delta})_{\nu}\right]=0 and div[(𝒖ˇδ)ν]=0\text{div}[(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}]=0, ν>0\forall\nu>0 and t[0,T]\forall t\in[0,T];

  • Convergence properties:

    (𝒖δ)ν𝒖δ strongly in Lp(0,T;Lq(Ωf,δ(t))), for all p[1,),q[1,2),\displaystyle(\boldsymbol{u}_{\delta})_{\nu}\to\boldsymbol{u}_{\delta}\qquad\text{ strongly in }L^{p}(0,T;L^{q}(\Omega_{f,\delta}(t))),\qquad\text{ for all }p\in[1,\infty),q\in[1,2),
    (𝒖ˇδ)ν𝒖ˇδ strongly in Lp(0,T;Lq(Ωf,1(t))), for all p[1,),q[1,2),\displaystyle(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}\to\widecheck{\boldsymbol{u}}_{\delta}\qquad\text{ strongly in }L^{p}(0,T;L^{q}(\Omega_{f,1}(t))),\qquad\text{ for all }p\in[1,\infty),q\in[1,2),
    (𝒖δ)ν𝒖δ weakly in L2(0,T;W1,p(Ωf,δ(t))), for all p[1,2),\displaystyle(\boldsymbol{u}_{\delta})_{\nu}\rightharpoonup\boldsymbol{u}_{\delta}\qquad\text{ weakly in }L^{2}(0,T;W^{1,p}(\Omega_{f,\delta}(t))),\qquad\text{ for all }p\in[1,2),
    (𝒖ˇδ)ν𝒖ˇδ weakly in L2(0,T;W1,p(Ωf,1(t))), for all p[1,2).\displaystyle(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}\rightharpoonup\widecheck{\boldsymbol{u}}_{\delta}\qquad\text{ weakly in }L^{2}(0,T;W^{1,p}(\Omega_{f,1}(t))),\qquad\text{ for all }p\in[1,2).
Proof.

(Proof of Gronwall’s estimate.)

We begin by testing the weak formulation (10.1) for the classical solution (𝒖,𝜼,p,ω)(\boldsymbol{u},\boldsymbol{\eta},p,\omega) to the original non-regularized problem with

𝒗=𝒖(𝒖ˇδ)ν,φ=ζ(ζδ)ν,𝝍=𝝃(𝝃δ)ν,r=p(pδ)ν,\boldsymbol{v}=\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu},\quad\varphi=\zeta-(\zeta_{\delta})_{\nu},\quad\boldsymbol{\psi}=\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu},\quad r=p-(p_{\delta})_{\nu}, (121)

and then test the regularized weak formulations (10.1) for the weak solutions (𝒖δ,𝜼δ,pδ,ωδ)(\boldsymbol{u}_{\delta},\boldsymbol{\eta}_{\delta},p_{\delta},\omega_{\delta}) with

𝒗=𝒖^,φ=ζ,𝝍=𝝃,r=p.\boldsymbol{v}=\widehat{\boldsymbol{u}},\quad\varphi=\zeta,\quad\boldsymbol{\psi}=\boldsymbol{\xi},\quad r=p. (122)

Next, we rewrite the energy estimate in Proposition 9.3, which holds for the function 𝒖δ\boldsymbol{u}_{\delta}, in a more convenient form by adding extra terms that will cancel out, in order to have the energy inequality parallel the weak formulation term by term. In particular, we have that for almost every t[0,Tδ]t\in[0,T_{\delta}],

12Ωf,δ(t)|𝒖δ|2+120tΩf(s)[((𝒖δ)𝒖δ)𝒖δ((𝒖δ)𝒖δ]+120tΓδ(s)(𝒖δ𝒏2𝝃δ𝒏)𝒖δ𝒖δ+2ν0tΩf,δ(s)|𝑫(𝒖δ)|2+0tΓδ(s)(12|𝒖δ|2pδ)(𝝃δ𝒖δ)𝒏+β0tΓδ(s)|(𝝃δ𝒖δ)𝒕)|2+12ρpΓ|𝝃δ|2+12Γ|Δωδ|2+12ρbΩb|𝝃δ|2+μe0tΩb|𝑫(𝜼δ)(s)|2+12λeΩb|𝜼δ(s)|2+2μv0tΩb|𝑫(𝝃δ)|2+λv0tΩb|𝝃δ|2α0tΩb,δδ(s)pδ𝝃δ+12c00tΩb|pδ(s)|2α0tΩb,δδ(s)DδDt𝜼δpδα0tΓδδ(s)(𝝃δ𝒏)pδ+κ0tΩb,δδ(s)|pδ|20tΓδ(s)((𝒖δ𝝃δ)𝒏)pδ12Ωf(0)|𝒖0|2+12ρpΓ|𝝃0|2+12Γ|Δω0|2+12ρbΩb|𝝃0|2+μeΩb|𝑫(𝜼0)|2+12λeΩb|𝜼0|2+12c0Ωb|p0|2.\frac{1}{2}\int_{\Omega_{f,\delta}(t)}|\boldsymbol{u}_{\delta}|^{2}+\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f}(s)}[((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta})\cdot\boldsymbol{u}_{\delta}-((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta}]+\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{u}_{\delta}\cdot\boldsymbol{n}-2\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n})\boldsymbol{u}_{\delta}\cdot\boldsymbol{u}_{\delta}\\ +2\nu\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}|\boldsymbol{D}(\boldsymbol{u}_{\delta})|^{2}+\int_{0}^{t}\int_{\Gamma_{\delta}(s)}\left(\frac{1}{2}|\boldsymbol{u}_{\delta}|^{2}-p_{\delta}\right)(\boldsymbol{\xi}_{\delta}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{n}+\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}|(\boldsymbol{\xi}_{\delta}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{t})|^{2}\\ +\frac{1}{2}\rho_{p}\int_{\Gamma}|\boldsymbol{\xi}_{\delta}|^{2}+\frac{1}{2}\int_{\Gamma}|\Delta\omega_{\delta}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\xi}_{\delta}|^{2}+\mu_{e}\int_{0}^{t}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}_{\delta})(s)|^{2}\\ +\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}_{\delta}(s)|^{2}+2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\xi}_{\delta})|^{2}+\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\xi}_{\delta}|^{2}\\ -{\alpha}\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}p_{\delta}\nabla\cdot\boldsymbol{\xi}_{\delta}+\frac{1}{2}c_{0}\int_{0}^{t}\int_{\Omega_{b}}|p_{\delta}(s)|^{2}-{\alpha}\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}\frac{{D}^{\delta}}{Dt}\boldsymbol{\eta}_{\delta}\cdot\nabla p_{\delta}-{\alpha}\int_{0}^{t}\int_{{\Gamma}^{\delta}_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n})p_{\delta}\\ +\kappa\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}|\nabla p_{\delta}|^{2}-\int_{0}^{t}\int_{\Gamma_{\delta}(s)}((\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n})p_{\delta}\leq\frac{1}{2}\int_{\Omega_{f}(0)}|\boldsymbol{u}_{0}|^{2}+\frac{1}{2}\rho_{p}\int_{\Gamma}|\boldsymbol{\xi}_{0}|^{2}\\ +\frac{1}{2}\int_{\Gamma}|\Delta\omega_{0}|^{2}+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\xi}_{0}|^{2}+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}_{0})|^{2}+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}_{0}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p_{0}|^{2}. (123)

Finally, we combine the weak formulation for 𝒖\boldsymbol{u} tested with (121), subtract the regularized weak formulation for 𝒖δ\boldsymbol{u}_{\delta} tested with (122), and add the energy estimate (123) for 𝒖δ\boldsymbol{u}_{\delta} to obtain an expression of the form

i=118Ti0,\sum_{i=1}^{18}T_{i}\leq 0, (124)

where the terms TiT_{i} are given below. We have to estimate each term, and the combined estimate will give the Gronwall’s inequality (118). To make this section more concise, we summarize the final estimates here, and present details of the derivation of these terms and the estimates in Appendix A.3.

As a notational note, in many of the estimates on the terms TiT_{i} that follow, we will use Cauchy’s inequality with ϵ\epsilon often: |ab|ϵ|a|2+C(ϵ)|b|2\displaystyle|ab|\leq\epsilon|a|^{2}+C(\epsilon)|b|^{2} where ϵ>0\epsilon>0 is any parameter and C(ϵ)C(\epsilon) is a constant that depends on the final choice of ϵ>0\epsilon>0. In the inequalities that appear, ϵ>0\epsilon>0 will hence be a parameter appearing on dissipative terms that will, at the conclusion of the estimates, be chosen small enough so that the dissipative terms from the estimates on TiT_{i} with ϵ\epsilon can be absorbed by the dissipative terms in (2) to give the final inequality (113).

Term T1. Term T1T_{1} is defined as follows:

T1=\displaystyle T_{1}= 0tΩf(s)𝒖t[𝒖(𝒖ˇδ)ν]120tΓ(s)(𝝃𝒏)𝒖[𝒖(𝒖ˇδ)ν]+Ωf(t)𝒖(t)[𝒖(𝒖ˇδ)ν](t)\displaystyle-\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{u}\cdot\partial_{t}\left[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}\right]-\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}]+\int_{\Omega_{f}(t)}\boldsymbol{u}(t)\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}](t) (125)
Ωf(0)𝒖(0)[𝒖(𝒖ˇδ)ν](0)0tΩf,δ(s)𝒖δt𝒖^120tΓδ(s)(𝝃δ𝒏δ)𝒖δ𝒖^\displaystyle-\int_{\Omega_{f}(0)}\boldsymbol{u}(0)\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}](0)-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\boldsymbol{u}_{\delta}\cdot\partial_{t}\widehat{\boldsymbol{u}}-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta})\boldsymbol{u}_{\delta}\cdot\widehat{\boldsymbol{u}}
+Ωf,δ(t)𝒖δ(t)𝒖^(t)Ωf(0)𝒖δ(0)𝒖^(0)+12Ωf,δ(t)|𝒖δ(t)|212Ωf,δ(0)|𝒖0|2.\displaystyle+\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\widehat{\boldsymbol{u}}(t)-\int_{\Omega_{f}(0)}\boldsymbol{u}_{\delta}(0)\cdot\widehat{\boldsymbol{u}}(0)+\frac{1}{2}\int_{\Omega_{f,\delta}(t)}|\boldsymbol{u}_{\delta}(t)|^{2}-\frac{1}{2}\int_{\Omega_{f,\delta}(0)}|\boldsymbol{u}_{0}|^{2}. (126)

This term is estimated so that after taking the limit as ν0\nu\to 0, the contribution of this term becomes

T1=12Ωf,δ(t)|(𝒖^𝒖δ)(t)|2+R1,T_{1}=\frac{1}{2}\int_{\Omega_{f,\delta}(t)}|(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})(t)|^{2}+{{R_{1}}},

where

|R1|\displaystyle{{|R_{1}|}}\leq ϵ0t𝒖^𝒖δH1(Ωf,δ(s))2\displaystyle\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{H^{1}(\Omega_{f,\delta}(s))}
+C(ϵ)(0tωωδH2(Γ)2+0ttωtωδL2(Γ)2+0t𝒖^𝒖δL2(Ωf,δ(s))2).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\partial_{t}\omega-\partial_{t}\omega_{\delta}||_{L^{2}(\Gamma)}^{2}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Term T2. Term T2T_{2} is defined as follows:

T2=\displaystyle T_{2}= 120tΩf(s)((𝒖)𝒖)[𝒖(𝒖ˇδ)ν]120tΩf(s)(𝒖)[𝒖(𝒖ˇδ)ν]𝒖\displaystyle\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f}(s)}((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}]-\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f}(s)}(\boldsymbol{u}\cdot\nabla)[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}]\cdot\boldsymbol{u}
120tΩf,δ(s)((𝒖δ)𝒖δ)(𝒖^𝒖δ)+120tΩf,δ(s)((𝒖δ)(𝒖^𝒖δ))𝒖δ.\displaystyle-\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta})\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})+\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}((\boldsymbol{u}_{\delta}\cdot\nabla)(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}))\cdot\boldsymbol{u}_{\delta}. (127)

After taking the limit ν0\nu\to 0, term T2T_{2} can be estimated as follows:

|T2|ϵ0t(𝒖^𝒖δ)L2(Ωf,δ(s))2+C(ϵ)(0tωωδH2(Γ)2+0t𝒖^𝒖δL2(Ωf,δ(s))2).|T_{2}|\leq\epsilon\int_{0}^{t}||\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Term T3. Term T3T_{3} is defined as follows:

T3=\displaystyle T_{3}= 120tΓ(s)(𝒖𝒏𝝃𝒏)𝒖[𝒖(𝒖ˇδ)ν]120tΓδ(s)(𝒖δ𝒏δ𝝃δ𝒏δ)𝒖δ𝒖^\displaystyle\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{u}\cdot\boldsymbol{n}-\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}]-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{u}_{\delta}\cdot\boldsymbol{n}_{\delta}-\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta})\boldsymbol{u}_{\delta}\cdot\widehat{\boldsymbol{u}}
+120tΓ(s)|𝒖|2(𝝃𝒏𝒖𝒏)120tΓ(s)|𝒖|2[(𝝃δ)ν𝒏(𝒖ˇδ)ν𝒏]120tΓδ(s)|𝒖δ|2(𝝃𝒏δ𝒖^𝒏δ).\displaystyle+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}|\boldsymbol{u}|^{2}(\boldsymbol{\xi}\cdot\boldsymbol{n}-\boldsymbol{u}\cdot\boldsymbol{n})-\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}|\boldsymbol{u}|^{2}[(\boldsymbol{\xi}_{\delta})_{\nu}\cdot\boldsymbol{n}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}\cdot\boldsymbol{n}]-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}|\boldsymbol{u}_{\delta}|^{2}(\boldsymbol{\xi}\cdot\boldsymbol{n}_{\delta}-\widehat{\boldsymbol{u}}\cdot\boldsymbol{n}_{\delta}). (128)

After taking the limit ν0\nu\to 0, term T3T3 can be estimated as follows:

|T3|\displaystyle|T_{3}|\leq ϵ0t𝒖^𝒖δH1(Ωf,δ(s))2\displaystyle\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{H^{1}(\Omega_{f,\delta}(s))}^{2}
+C(ϵ)(0tωωδH2(Γ)2+0t𝝃𝝃δL2(Γ)2+0t𝒖^𝒖δL2(Ωf,δ(s))2).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Term T4. Term T4T_{4} is defined as follows:

T4=\displaystyle T_{4}= 2ν0tΩf(s)𝑫(𝒖):𝑫(𝒖(𝒖ˇδ)ν)2ν0tΩf,δ(s)𝑫(𝒖δ):𝑫(𝒖^𝒖δ).\displaystyle 2\nu\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu})-2\nu\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\boldsymbol{D}(\boldsymbol{u}_{\delta}):\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}). (129)

After taking the limit ν0\nu\to 0, term T4T4 can be estimated as follows:

T4=2ν0tΩf,δ(s)|𝑫(𝒖^𝒖δ)|2+R4,T_{4}=2\nu\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}|\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})|^{2}+{{R_{4}}},

where

|R4|ϵ0t𝑫(𝒖^𝒖δ)L2(Ωf,δ(s))2+C(ϵ)(0tωωδH2(Γ)2+0t𝒖^𝒖δL2(Ωf,δ(s))2).{{|R_{4}|}}\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||^{2}_{L^{2}(\Omega_{f,\delta}(s))}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Term T5. Term T5T_{5} is defined as follows:

T5=β0tΓ(s)(𝝃𝒖)𝝉(s)[(𝝃(𝝃δ)ν)𝝉(s)(𝒖(𝒖ˇδ)ν)𝝉(s)]β0tΓδ(s)(𝝃δ𝒖δ)𝝉δ(s)[(𝝃𝝃δ)𝝉δ(s)(𝒖^𝒖δ)𝝉δ(s)],T_{5}=\beta\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}(s)[(\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu})\cdot\boldsymbol{\tau}(s)-(\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu})\cdot\boldsymbol{\tau}(s)]\\ -\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)-(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)],

where 𝝉(s)\boldsymbol{\tau}(s) is the unit tangent vector to Γ(s)\Gamma(s) and 𝝉δ(s)\boldsymbol{\tau}_{\delta}(s) is the unit tangent vector to Γδ(s)\Gamma_{\delta}(s).

After taking the limit ν0\nu\to 0, term T5T_{5} can be estimated as follows:

T5=β0tΓδ(s)|(𝝃𝝃δ)𝝉δ(s)(𝒖^𝒖δ)𝝉δ(s)|2+R5,T_{5}=\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}|(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)-(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)|^{2}+R_{5},

where

|R5|ϵ0t𝑫(𝒖^𝒖δ)L2(Ωf,δ(s))+C(ϵ)(0tωωδH2(Γ)2+0t𝝃𝝃δL2(Γ)2).|R_{5}|\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||_{L^{2}(\Gamma)}^{2}\right).

Terms T6-T8. Terms T6T_{6}-T8T_{8} are defined as follows:

T6=\displaystyle T_{6}= ρp0tΓζt[ζ(ζδ)ν]+ρpΓζ(s)[ζ(t)(ζδ)ν(t)]ρpΓζ(0)[ζ(0)(ζδ)ν(0)]\displaystyle-\rho_{p}\int_{0}^{t}\int_{\Gamma}\zeta\cdot\partial_{t}\left[\zeta-(\zeta_{\delta})_{\nu}\right]+\rho_{p}\int_{\Gamma}\zeta(s)\cdot\left[\zeta(t)-(\zeta_{\delta})_{\nu}(t)\right]-\rho_{p}\int_{\Gamma}\zeta(0)\cdot\left[\zeta(0)-(\zeta_{\delta})_{\nu}(0)\right]
+ρp0tΓζδtζρpΓζδ(t)ζ(t)+ρpΓζδ(0)ζ(0)+12ρpΓ|ζδ(t)|212ρpΓ|ζ0|2.\displaystyle+\rho_{p}\int_{0}^{t}\int_{\Gamma}\zeta_{\delta}\cdot\partial_{t}\zeta-\rho_{p}\int_{\Gamma}\zeta_{\delta}(t)\cdot\zeta(t)+\rho_{p}\int_{\Gamma}\zeta_{\delta}(0)\cdot\zeta(0)+\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta_{\delta}(t)|^{2}-\frac{1}{2}\rho_{p}\int_{\Gamma}|\zeta_{0}|^{2}. (130)
T7=\displaystyle T_{7}= 0tΓΔωΔ[ζ(ζδ)ν]0tΓΔωδΔζ+12Γ|Δωδ(t)|212Γ|Δω0|2.\displaystyle\int_{0}^{t}\int_{\Gamma}\Delta\omega\cdot\Delta\left[\zeta-(\zeta_{\delta})_{\nu}\right]-\int_{0}^{t}\int_{\Gamma}\Delta\omega_{\delta}\cdot\Delta\zeta+\frac{1}{2}\int_{\Gamma}|\Delta\omega_{\delta}(t)|^{2}-\frac{1}{2}\int_{\Gamma}|\Delta\omega_{0}|^{2}. (131)
T8=\displaystyle T_{8}= ρb0tΩbt𝜼t[𝝃(𝝃δ)ν]+ρbΩb𝝃(t)[𝝃(s)(𝝃δ)ν(s)]ρbΩb𝝃(0)[𝝃(0)(𝝃δ)ν(0)]\displaystyle-\rho_{b}\int_{0}^{t}\int_{\Omega_{b}}\partial_{t}\boldsymbol{\eta}\cdot\partial_{t}\left[\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}\right]+\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}(t)\cdot\left[\boldsymbol{\xi}(s)-(\boldsymbol{\xi}_{\delta})_{\nu}(s)\right]-\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}(0)\cdot\left[\boldsymbol{\xi}(0)-(\boldsymbol{\xi}_{\delta})_{\nu}(0)\right]
+ρb0tΩbt𝜼δt𝝃ρbΩb𝝃δ(t)𝝃(t)+ρbΩb𝝃δ(0)𝝃(0)+12ρbΩb|𝝃δ(t)|212ρbΩb|𝝃0|2.\displaystyle+\rho_{b}\int_{0}^{t}\int_{\Omega_{b}}\partial_{t}\boldsymbol{\eta}_{\delta}\cdot\partial_{t}\boldsymbol{\xi}-\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}_{\delta}(t)\cdot\boldsymbol{\xi}(t)+\rho_{b}\int_{\Omega_{b}}\boldsymbol{\xi}_{\delta}(0)\cdot\boldsymbol{\xi}(0)+\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\xi}_{\delta}(t)|^{2}-\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|\boldsymbol{\xi}_{0}|^{2}. (132)

After taking the limit ν0\nu\to 0, the terms T6T_{6}-T8T_{8} become:

T6=12ρpΓ|(ζζδ)(t)|2,T7=12Γ|Δ(ωωδ)(t)|2,T8=12ρbΩb|(𝝃𝝃δ)(t)|2.T_{6}=\frac{1}{2}\rho_{p}\int_{\Gamma}|(\zeta-\zeta_{\delta})(t)|^{2},\ \ \ T_{7}=\frac{1}{2}\int_{\Gamma}|\Delta(\omega-\omega_{\delta})(t)|^{2},\ \ \ T_{8}=\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})(t)|^{2}.

Terms T9-T12. Terms T9T_{9}-T12T_{12} are defined as follows:

T9=\displaystyle T_{9}= 2μe0tΩb𝑫(𝜼):𝑫[𝝃(𝝃δ)ν]2μe0tΩb𝑫(𝜼δ):𝑫(𝝃)+μeΩb|𝑫(𝜼δ)(t)|2μeΩb|𝑫(𝜼0)|2.\displaystyle 2\mu_{e}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}):\boldsymbol{D}\left[\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}\right]-2\mu_{e}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\eta}_{\delta}):\boldsymbol{D}(\boldsymbol{\xi})+\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}_{\delta})(t)|^{2}-\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}_{0})|^{2}.
T10=\displaystyle T_{10}= λe0tΩb(𝜼)([𝝃(𝝃δ)ν])λe0tΩb(𝜼δ)(𝝃)+12λeΩb|𝜼δ(t)|212λeΩb|𝜼0|2.\displaystyle\lambda_{e}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta})\left(\nabla\cdot\left[\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}\right]\right)-\lambda_{e}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\eta}_{\delta})(\nabla\cdot\boldsymbol{\xi})+\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}_{\delta}(t)|^{2}-\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\eta}_{0}|^{2}.
T11=\displaystyle T_{11}= 2μv0tΩb𝑫(𝝃):𝑫[𝝃(𝝃δ)ν]2μv0tΩb𝑫(𝝃δ):𝑫(𝝃)+2μv0tΩb|𝑫(𝝃δ)|2.\displaystyle 2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\xi}):\boldsymbol{D}\left[\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}\right]-2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{D}(\boldsymbol{\xi}_{\delta}):\boldsymbol{D}(\boldsymbol{\xi})+2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\xi}_{\delta})|^{2}.
T12=\displaystyle T_{12}= λv0tΩb(𝝃)([𝝃(𝝃δ)ν])λv0tΩb(𝝃δ)(𝝃)+λv0tΩb|𝝃δ|2.\displaystyle\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\xi})\left(\nabla\cdot\left[\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}\right]\right)-\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}(\nabla\cdot\boldsymbol{\xi}_{\delta})(\nabla\cdot\boldsymbol{\xi})+\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}|\nabla\cdot\boldsymbol{\xi}_{\delta}|^{2}. (133)

Because 𝝃δL2(0,T;H1(Ωb))\boldsymbol{\xi}_{\delta}\in L^{2}(0,T;H^{1}(\Omega_{b})) where Ωb\Omega_{b} is a fixed domain, we have that (𝝃δ)ν𝝃δ(\boldsymbol{\xi}_{\delta})_{\nu}\to\boldsymbol{\xi}_{\delta} strongly in L2(0,T;H1(Ωb))L^{2}(0,T;H^{1}(\Omega_{b})). Hence, as ν0\nu\to 0, we have that Terms 9-12 converge to the following:

T9=μeΩb|𝑫(𝜼𝜼δ)(t)|2,T10=12λeΩb|(𝜼𝜼δ)(t)|2,T_{9}=\mu_{e}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta})(t)|^{2},\qquad T_{10}=\frac{1}{2}\lambda_{e}\int_{\Omega_{b}}|\nabla\cdot(\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta})(t)|^{2},
T11=2μv0tΩb|𝑫(𝝃𝝃δ)|2,T12=λv0tΩb|(𝝃𝝃δ)|2.T_{11}=2\mu_{v}\int_{0}^{t}\int_{\Omega_{b}}|\boldsymbol{D}(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})|^{2},\qquad T_{12}=\lambda_{v}\int_{0}^{t}\int_{\Omega_{b}}|\nabla\cdot(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})|^{2}.

Term T13. Term T13T_{13} is defined as follows:

T13=\displaystyle T_{13}= α0tΩb(s)p([𝝃(𝝃δ)ν])+α0tΩb,δδ(s)pδ((𝝃𝝃δ)).\displaystyle-{\alpha}\int_{0}^{t}\int_{\Omega_{b}(s)}p\left(\nabla\cdot\left[\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}\right]\right)+{\alpha}\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}p_{\delta}\left(\nabla\cdot(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\right).

After taking the limit ν0\nu\to 0, term T13T_{13} can be estimated as follows:

|T13|\displaystyle|T_{13}|\leq C(ϵ)0t𝜼𝜼δL2(Ωb)2+ϵ0t(𝝃𝝃δ)L2(Ωb)2\displaystyle C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}+\epsilon\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||_{L^{2}(\Omega_{b})}^{2}
+C(ϵ)(0t𝜼𝜼δL2(Ωb)2+0tωωδH2(Γ)2+0tppδL2(Ωb)2).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||p-p_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right).

Term T14. Term T14T_{14} is defined as follows:

T14=\displaystyle T_{14}= c00tΩbpt[p(pδ)ν]+c0Ωbp(t)[p(t)(pδ)ν(t)]c0Ωbp0[p(0)(pδ)ν(0)]\displaystyle-c_{0}\int_{0}^{t}\int_{\Omega_{b}}p\cdot\partial_{t}\left[p-(p_{\delta})_{\nu}\right]+c_{0}\int_{\Omega_{b}}p(t)\cdot\left[p(t)-(p_{\delta})_{\nu}(t)\right]-c_{0}\int_{\Omega_{b}}p_{0}\cdot\left[p(0)-(p_{\delta})_{\nu}(0)\right]
+c00tΩbpδtpc0Ωbpδ(t)p(t)+c0Ωb|p0|2+12c0Ωb|pδ(t)|212c0Ωb|p0|2.\displaystyle+c_{0}\int_{0}^{t}\int_{\Omega_{b}}p_{\delta}\cdot\partial_{t}p-c_{0}\int_{\Omega_{b}}p_{\delta}(t)\cdot p(t)+c_{0}\int_{\Omega_{b}}|p_{0}|^{2}+\frac{1}{2}c_{0}\int_{\Omega_{b}}|p_{\delta}(t)|^{2}-\frac{1}{2}c_{0}\int_{\Omega_{b}}|p_{0}|^{2}. (134)

This term can be handled in the same way as Terms 6-8. In the limit as ν0\nu\to 0, the contribution from this term is

T14=12c0Ωb|(ppδ)(t)|2.T_{14}=\frac{1}{2}c_{0}\int_{\Omega_{b}}|(p-p_{\delta})(t)|^{2}.

Term T15. Term T15T_{15} is defined as follows:

T15=\displaystyle T_{15}= α0tΩb(s)𝝃[p(pδ)ν]+α0tΩb,δδ(s)𝝃δ(ppδ).\displaystyle-{\alpha}\int_{0}^{t}\int_{\Omega_{b}(s)}\boldsymbol{\xi}\cdot\nabla\left[p-(p_{\delta})_{\nu}\right]+{\alpha}\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}\boldsymbol{\xi}_{\delta}\cdot\nabla(p-p_{\delta}). (135)

After taking the limit ν0\nu\to 0, term T15T_{15} can be estimated as follows:

|T15|\displaystyle|T_{15}|\leq ϵ0t(ppδ)L2(Ωb,δδ(s))2+C(ϵ)0t𝜼𝜼δL2(Ωb)2\displaystyle\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||_{L^{2}(\Omega^{\delta}_{b,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}
+C(ϵ)(0t𝜼𝜼δL2(Ωb)2+0tωωδH2(Γ)2+0tt𝜼t𝜼δL2(Ωb)2).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\partial_{t}\boldsymbol{\eta}-\partial_{t}\boldsymbol{\eta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right).

Term T16. Term T16T_{16} is defined as follows:

T16=\displaystyle T_{16}= α0tΓ(s)(𝝃𝒏)[p(pδ)ν]+α0tΓδδ(s)(𝝃δ𝒏)(ppδ).\displaystyle-{\alpha}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\left[p-(p_{\delta})_{\nu}\right]+{\alpha}\int_{0}^{t}\int_{{\Gamma}^{\delta}_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n})(p-p_{\delta}). (136)

After passing to the limit as ν0\nu\to 0, this term can be estimated as follows:

|T16|ϵ(0t||𝝃𝝃δ||L2(Ωb)2+0t||ppδ||L2(Ωb,δδ(s))2)+C(ϵ)(0t||ppδ||L2(Ωb)2+0t||𝜼𝜼δ||L2(Ωb)2+0t||𝜼𝜼δ||L2(Ωb)2+0t||ωωδ||H2(Γ)2+0t||𝝃𝝃δ||L2(Ωb)2).|T_{16}|\leq\epsilon\left(\int_{0}^{t}||\nabla\boldsymbol{\xi}-\nabla\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\nabla p-\nabla p_{\delta}||_{L^{2}(\Omega^{\delta}_{b,\delta}(s))}^{2}\right)+C(\epsilon)\left(\int_{0}^{t}||p-p_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right.\\ \left.+\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}^{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right).

Term T17. Term T17T_{17} is defined as follows:

T17=\displaystyle T_{17}= κ0tΩb(s)p[p(pδ)ν]κ0tΩb,δδ(s)pδ(ppδ).\displaystyle\kappa\int_{0}^{t}\int_{\Omega_{b}(s)}\nabla p\cdot\nabla\left[p-(p_{\delta})_{\nu}\right]-\kappa\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}\nabla p_{\delta}\cdot\nabla(p-p_{\delta}). (137)

This term can be estimated as follows:

T17κ0tΩb,δδ(s)|(ppδ)|2+R17,T_{17}\leq\kappa\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}|\nabla(p-p_{\delta})|^{2}+{{R_{17}}},

where the remainder is bounded by

|R17|\displaystyle{{|R_{17}|}}\leq ϵ0t||(ppδ)||2L2(Ωδb,δ(s))\displaystyle\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||^{2}_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}
+C(ϵ)(0t||𝜼𝜼δ||L2(Ωb)2+0t||𝜼𝜼δ||L2(Ωb)2+0t||ωωδ||H2(Γ)2).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}\right).

Term T18. Term T18T_{18} is defined as follows:

T18\displaystyle T_{18} =0tΓ(s)p(𝒖𝝃)𝒏0tΓ(s)p[(𝒖δ)ν(𝝃δ)ν]𝒏0tΓδ(s)pδ(𝒖𝝃)𝒏\displaystyle=\int_{0}^{t}\int_{\Gamma(s)}p(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n}-\int_{0}^{t}\int_{\Gamma(s)}p[(\boldsymbol{u}_{\delta})_{\nu}-(\boldsymbol{\xi}_{\delta})_{\nu}]\cdot\boldsymbol{n}-\int_{0}^{t}\int_{\Gamma_{\delta}(s)}p_{\delta}(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n}
+0tΓδ(s)pδ(𝒖δ𝝃δ)𝒏0tΓ(s)((𝒖𝝃)𝒏)[p(pδ)ν]+0tΓδ(s)((𝒖δ𝝃δ)𝒏)(ppδ).\displaystyle+\int_{0}^{t}\int_{\Gamma_{\delta}(s)}p_{\delta}(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n}-\int_{0}^{t}\int_{\Gamma(s)}((\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n})[p-(p_{\delta})_{\nu}]+\int_{0}^{t}\int_{\Gamma_{\delta}(s)}((\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n})(p-p_{\delta}).

This term can be estimated as follows:

|T18|\displaystyle|T_{18}|\leq ϵ(0t||𝑫(𝒖^𝒖δ)||2L2(Ωf,δ(s))+0t||(𝝃𝝃δ)||L2(Ωb)+0t||(ppδ)||2L2(Ωδb,δ(s)))\displaystyle\epsilon\left(\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||^{2}_{L^{2}(\Omega_{f,\delta}(s))}+\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\nabla(p-p_{\delta})||^{2}_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}\right)
+C(ϵ)0t||ωωδ||2H2(Γ).\displaystyle+C(\epsilon)\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}.

The combined estimates for the terms T1T_{1}-T18T_{18} give the estimate:

Eδ(t)ϵ(0t||𝒖^𝒖δ||2H1(Ωf,δ(s))+0t||𝑫(𝒖^𝒖δ)||2L2(Ωf,δ(s))+0t||(𝝃𝝃δ)||2L2(Ωb)+0t||(ppδ)||2L2(Ωb,δδ(s)))+C(ϵ)(0t||ωωδ||H2(Γ)2+0t||𝝃𝝃δ||2L2(Γ)+0t||𝒖^𝒖δ||2L2(Ωf,δ(s))+0t||𝜼𝜼δ||2L2(Ωb)+0t||(𝜼𝜼δ)||L2(Ωb)2+0t||ppδ||L2(Ωb)2+0t||𝝃𝝃δ||2L2(Ωb)),E_{\delta}(t)\leq\epsilon\left(\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{H^{1}(\Omega_{f,\delta}(s))}+\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||^{2}_{L^{2}(\Omega_{f,\delta}(s))}\right.\\ \left.+\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\nabla(p-p_{\delta})||^{2}_{L^{2}(\Omega_{b,\delta}^{\delta}(s))}\right)\\ +C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{L^{2}(\Omega_{f,\delta}(s))}+\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}^{\delta}||^{2}_{L^{2}(\Omega_{b})}\right.\\ \left.+\int_{0}^{t}||\nabla(\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta})||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||p-p_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right),

where ϵ>0\epsilon>0 remains to be chosen, and C(ϵ)C(\epsilon) is a constant depending only on ϵ\epsilon, that is independent of tt and δ\delta. By using Korn’s inequality, Poincaré’s inequality, and by choosing ϵ\epsilon sufficiently small to absorb the terms on the right hand side into the dissipation terms in Eδ(t)E_{\delta}(t) defined by (2), we then obtain the final inequality (118). This finishes the proof of the Gronwall estimate presented in Lemma 10.1. ∎

All that is left to show to complete the proof of weak-classical consistency stated in Theorem 10.1, is to argue that Gronwall’s inequality (118) holds for all t[0,T]t\in[0,T] where TT is independent of δ\delta. This will also imply the first statement in the theorem, which states that (𝜼δ,ωδ,pδ,𝒖δ)(\boldsymbol{\eta}_{\delta},\omega_{\delta},p_{\delta},\boldsymbol{u}_{\delta}) is uniformly defined on the time interval [0,T][0,T] for all δ>0\delta>0. In order to do this we use a bootstrap argument presented in the next subsection.

10.4 Bootstrap argument

To obtain the desired Gronwall estimate as stated in Lemma 10.1, we need the following uniform bounds (115), (116), and (117) on the factor det(𝑰+𝜼δδ)\det(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}), which appears in the regularized weak formulation (40) defined on the fixed reference domain Ωb\Omega_{b}:

det(𝑰+𝜼δδ)\displaystyle\det(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}) c>0,\displaystyle\geq c>0,
0<c|𝑰+𝜼δδ|\displaystyle 0<c\leq|\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}| C, pointwise in Ωb¯,\displaystyle\leq C,\qquad\text{ pointwise in }\overline{\Omega_{b}},
|𝜼δδ|\displaystyle|\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}| C, pointwise in Ωb¯,\displaystyle\leq C,\qquad\text{ pointwise in }\overline{\Omega_{b}},

which need to hold for all t[0,T]t\in[0,T] where T>0T>0 is independent of δ\delta. Notice that we only have uniform boundedness of 𝜼δ\boldsymbol{\eta}_{\delta} with respect to δ\delta in L(0,T;H1(Ωb))L^{\infty}(0,T;H^{1}(\Omega_{b})), which implies that det(𝑰+𝜼δδ)\det(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}) is uniformly bounded with respect to δ\delta only in L(0,T;L1(Ωb))L^{\infty}(0,T;L^{1}(\Omega_{b})), which is insufficient for estimating any integrands with this factor.

To get around this difficulty we use the following strategy. Recall that by the way the weak solution to the regularized problem was constructed using the splitting scheme, we have that there exists a sufficiently small constant cc (uniform in δ\delta) such that

det(𝑰+𝜼δδ)c>0,\det(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta})\geq c>0, (138)

for all t[0,Tδ]t\in[0,T_{\delta}] where Tδ>0T_{\delta}>0 may depend on δ\delta. This estimate holds at least locally, although not locally uniformly, for each δ>0\delta>0. In fact, similarly, the following three estimates (115), (116), and (117) from Lemma 10.1 hold locally, for t[0,Tδ]t\in[0,T_{\delta}], where TδT_{\delta} may depend on δ\delta, with positive constants cc and CC that are independent of δ\delta:

det\displaystyle\det (𝑰+𝜼δδ)c>0,\displaystyle(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta})\geq c>0,
0<c\displaystyle 0<c\leq |𝑰+𝜼δδ|C, pointwise in Ωb¯,\displaystyle|\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}|\leq C,\qquad\text{ pointwise in }\overline{\Omega_{b}},
|𝜼δδ|C, pointwise in Ωb¯.\displaystyle|\nabla{\boldsymbol{\eta}}_{\delta}^{\delta}|\leq C,\qquad\text{ pointwise in }\overline{\Omega_{b}}.

These estimates imply that for sufficiently small c>0c>0, the following inequality also holds locally, for all t[0,Tδ]t\in[0,T_{\delta}]:

0<C1|(𝑰+𝜼δδ)1|c1.0<C^{-1}\leq|(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}_{\delta}^{\delta})^{-1}|\leq c^{-1}. (139)

Let [0,T][0,T] denote the time interval on which the classical solution 𝜼\boldsymbol{\eta} exists. Then, we can choose c>0c>0 and C>0C>0 so that the inequalities (138)-(139) also hold for the classical solution for all t[0,T]t\in[0,T].

We will now show that the time interval on which estimates (138)-(139) hold for the weak solution of the regularized problem 𝜼δδ{\boldsymbol{\eta}}_{\delta}^{\delta} can, in fact, be extended to the entire interval [0,T][0,T], namely, that estimates (138)-(139) hold globally, uniformly in δ\delta, where TT is independent of δ\delta. We will do this by using a classical bootstrap argument the steps of which we present below. This bootstrap argument will propagate the desired estimates which hold a priori locally on [0,Tδ][0,T_{\delta}] (for a time TδT_{\delta} possibly depending on δ\delta) to an entire uniform time interval [0,T][0,T].

The global uniform estimates will follow if we can show that 𝜼\nabla\boldsymbol{\eta} and 𝜼δδ\nabla{\boldsymbol{\eta}}_{\delta}^{\delta} are pointwise uniformly “close”, i.e.,

|(𝜼𝜼δδ)(t,x)|0pointwiseuniformlyin[0,T]×Ωbasδ0.|(\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}_{\delta}^{\delta})(t,x)|\to 0\ {\rm{pointwise\ uniformly\ in}}\ [0,T]\times\Omega_{b}\ {\rm{as}}\ \delta\to 0. (140)

To obtain this estimate we start with the main proof of Gronwall’s inequality under the assumptions that (115), (116), and (117) are locally valid for t[0,Tδ]t\in[0,T_{\delta}]:

Eδ(t)C10t||(𝜼δ𝜼)(s)||2H1(Ωb)ds+C20tEδ(s)ds,E_{\delta}(t)\leq C_{1}\int_{0}^{t}||({\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta})(s)||^{2}_{H^{1}(\Omega_{b})}ds+C_{2}\int_{0}^{t}E_{\delta}(s)ds, (141)

where the constants C1C_{1} and C2C_{2} are independent of δ\delta. Then, by Lemma 10.2 below, we obtain that the first term on the right hand-side above can be estimated as follows:

||𝜼δ𝜼||H1(Ωb)Cδ3/2, for all t[0,T],||{\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta}||_{H^{1}(\Omega_{b})}\leq C\delta^{3/2},\qquad\text{ for all }t\in[0,T], (142)

since the classical solution 𝜼\boldsymbol{\eta} is spatially smooth, and 𝜼δ{\boldsymbol{\eta}}^{\delta} is the convolution of 𝜼\boldsymbol{\eta} with the smooth δ\delta kernel, defined in (102). With this essential observation, the Gronwall estimate based on (141) gives

Eδ(t)C1(0t||(𝜼δ𝜼)(s)||2H1(Ωb)ds)eC2tC1(0T||(𝜼δ𝜼)(s)||2H1(Ωb)ds)eC2tCδ3eC2t.E_{\delta}(t)\leq C_{1}\left(\int_{0}^{t}||({\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta})(s)||^{2}_{H^{1}(\Omega_{b})}ds\right)e^{C_{2}t}\leq C_{1}\left(\int_{0}^{T}||({\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta})(s)||^{2}_{H^{1}(\Omega_{b})}ds\right)e^{C_{2}t}\leq C\delta^{3}e^{C_{2}t}.

By the definition of Eδ(t)E_{\delta}(t) and an application of Poincare’s and Korn’s inequalities on Ωb\Omega_{b}, see Proposition 6.1, this implies that the following terms in the definition of Eδ(t)E_{\delta}(t)

||(𝜼𝜼δ)(t)||H1(Ωb)Cδ3/2,and||(ωωδ)(t)||H2(Γ)Cδ3/20asδ0||(\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta})(t)||_{H^{1}(\Omega_{b})}\leq C\delta^{3/2},\ {\rm{and}}\ ||(\omega-\omega_{\delta})(t)||_{H^{2}(\Gamma)}\leq C\delta^{3/2}\to 0\ {\rm{as}}\ \delta\to 0

converge to zero as δ0\delta\to 0 at a rate of δ3/2\delta^{3/2}, as long as the assumptions (115), (116), and (117) hold. Therefore, by Hölder’s inequality, for sufficiently small δ>0\delta>0, we can prove that the following estimate holds:

|(𝜼δ𝜼δδ)(t,x)|\displaystyle|(\nabla{\boldsymbol{\eta}}^{\delta}-\nabla{\boldsymbol{\eta}}_{\delta}^{\delta})(t,x)| =|Ω~b(𝜼𝜼δ)(t,y)σδ(xy)dy|Cδ3/2δ10,\displaystyle=\left|\int_{\tilde{\Omega}_{b}}(\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta})(t,y)\sigma_{\delta}(x-y)dy\right|\leq C\delta^{3/2}\cdot\delta^{-1}\to 0, (143)
pointwise uniformly in [0,T]×Ωb as δ0,\displaystyle\text{pointwise uniformly in $[0,T]\times\Omega_{b}$ as $\delta\to 0$},

where CC is independent of δ\delta. More precisely, notice that the convolution integral in (143) is defined on the domain Ω~b{\tilde{\Omega}_{b}}, which is triple the size of the domain Ωb\Omega_{b}. Furthermore, we recall that the convolution and the larger domain Ω~b\tilde{\Omega}_{b} are defined using odd extensions as in Definition 5.1. Thus, by the definition of the odd extensions of 𝜼\boldsymbol{\eta} and 𝜼δ\boldsymbol{\eta}_{\delta} to the larger domain Ω~b\tilde{\Omega}_{b}, we get

||𝜼𝜼δ||H1(Ω~b)C(||𝜼𝜼δ||H1(Ωb)+||ωωδ||H1(Γ)).||\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta}||_{H^{1}(\tilde{\Omega}_{b})}\leq C\left(||\boldsymbol{\eta}-\boldsymbol{\eta}_{\delta}||_{H^{1}(\Omega_{b})}+||\omega-\omega_{\delta}||_{H^{1}(\Gamma)}\right).

In addition, since we have extended the functions 𝜼\boldsymbol{\eta} and 𝜼δ\boldsymbol{\eta}_{\delta} to the larger domain Ω~b\tilde{\Omega}_{b}, the estimate (143) holds for all δ\delta such that {(x,y)2:dist((x,y),Ωb)δ}Ω~b\{(x,y)\in\mathbb{R}^{2}:\text{dist}((x,y),\Omega_{b})\leq\delta\}\subset\tilde{\Omega}_{b}. Thus, |(𝜼δ𝜼δδ)(t,x)|0|(\nabla{\boldsymbol{\eta}}^{\delta}-\nabla{\boldsymbol{\eta}}_{\delta}^{\delta})(t,x)|\to 0 pointwise uniformly in [0,T][0,T] as δ0\delta\to 0.

To obtain (140) it suffices to show that |(𝜼𝜼δ)(t,x)|0|(\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta})(t,x)|\to 0 pointwise uniformly in [0,T][0,T] as δ0\delta\to 0. This follows from Lemma 10.2. Namely, Lemma 10.2 implies

|(𝜼𝜼δ)(t,x)|Cδ0pointwise uniformly in [0,T]×Ωb as δ0.|(\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta})(t,x)|\leq C\delta\to 0\ \text{pointwise uniformly in $[0,T]\times\Omega_{b}$ as $\delta\to 0$}. (144)

So combining (144) with (143), we get (140).

To conclude the bootstrap argument, we combine the fact that det(𝑰+𝜼)c>0\text{det}(\boldsymbol{I}+\nabla\boldsymbol{\eta})\geq c>0 on a time interval [0,T][0,T] with the estimate (140) (which states that 𝜼δδ\nabla\boldsymbol{\eta}^{\delta}_{\delta} and 𝜼\nabla\boldsymbol{\eta} are uniformly close on the full time interval [0,T][0,T] as δ0\delta\to 0), to deduce that det(𝑰+𝜼δδ)c>0\text{det}(\boldsymbol{I}+\nabla\boldsymbol{\eta}^{\delta}_{\delta})\geq c>0 also uniformly on [0,T][0,T], for all sufficiently small δ\delta. Similarly, for all sufficiently small δ\delta, the assumptions (116) and (117) will also hold up to the final time T>0T>0, as we can also propagate the estimates (116) and (117) similarly by combining estimate (140) with the fact that these estimates hold for the classical solution 𝜼\boldsymbol{\eta} on some time interval [0,T][0,T]. This closes the bootstrap argument, and so we obtain that the estimate (115), and similarly the estimates (116) and (117), hold uniformly up to the final time T>0T>0 uniformly in δ\delta.

We end this section by proving the following lemma, which establishes convergence of the spatial convolution of the classical solution 𝜼\boldsymbol{\eta} in H1(Ωb)H^{1}(\Omega_{b}), which is needed for the argument described above, in the estimates (142) and (144).

Lemma 10.2.

Let 𝜼L(0,T;Vd)\boldsymbol{\eta}\in L^{\infty}(0,T;V_{d}) be an arbitrary but fixed smooth function in time and space on [0,T]×Ωb¯[0,T]\times\overline{\Omega_{b}}, where VdV_{d} is defined in (32). Then, there exists a constant CC independent of δ>0\delta>0, depending only on 𝜼\boldsymbol{\eta}, such that

maxt[0,T]||𝜼δ𝜼||H1(Ωb)Cδ3/2,and|𝜼δ𝜼|CδxΩ¯bandt[0,T].\max_{t\in[0,T]}||{\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta}||_{H^{1}(\Omega_{b})}\leq C\delta^{3/2},\ {\rm{and}}\ |\nabla{\boldsymbol{\eta}}^{\delta}-\nabla\boldsymbol{\eta}|\leq C\delta\ \quad\forall x\in\overline{\Omega}_{b}\ {\rm{and}}\ \forall t\in[0,T].
Remark 10.2.

More generally, if ff is a smooth function on 2\mathbb{R}^{2} with sufficient decay at infinity, such as a Schwartz function, then the argument below shows that the function f~\tilde{f} defined by

f~=fσδ on 2\tilde{f}=f*\sigma_{\delta}\qquad\text{ on }\mathbb{R}^{2}

would satisfy ||f~f||H1(Ωb)Cδ2||\tilde{f}-f||_{H^{1}(\Omega_{b})}\leq C\delta^{2} for a constant CC. However, because we are working on a bounded domain Ωb\Omega_{b}, we must use an odd extension to define the spatial convolution of 𝜼\boldsymbol{\eta}. Since the odd extension of 𝜼\boldsymbol{\eta} to the larger domain Ω~b\tilde{\Omega}_{b} is not necessarily smooth on Ω~b\tilde{\Omega}_{b} even if 𝜼\boldsymbol{\eta} is a smooth function on Ωb¯\overline{\Omega_{b}}, we incur a loss in our estimate due to potential irregularities of the odd extension due to the behavior of the initial function 𝜼\boldsymbol{\eta} near the boundary Ωb\partial\Omega_{b}, which gives rise to the convergence rate δ3/2\delta^{3/2} instead of the optimal rate of convergence δ2\delta^{2}.

Proof.

Separate the domain Ωb=(0,L)×(0,R)\Omega_{b}=(0,L)\times(0,R) into two parts:

Ωb,1=(δ,Lδ)×(δ,Rδ),Ωb,2=ΩbΩb,1.\Omega_{b,1}=(\delta,L-\delta)\times(\delta,R-\delta),\qquad\Omega_{b,2}=\Omega_{b}\setminus\Omega_{b,1}.

For 𝒙Ωb,1\boldsymbol{x}\in\Omega_{b,1}, we note that because the convolution kernel σδ\sigma_{\delta} is radially symmetric,

(𝜼δ𝜼)(𝒙)=Ωb(12𝜼(𝒙+𝒙)𝜼(𝒙)+12𝜼(𝒙𝒙))σδ(𝒙)d𝒙,({\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta})(\boldsymbol{x})=\int_{\Omega_{b}}\left(\frac{1}{2}\boldsymbol{\eta}(\boldsymbol{x}+\boldsymbol{x}^{\prime})-\boldsymbol{\eta}(\boldsymbol{x})+\frac{1}{2}\boldsymbol{\eta}(\boldsymbol{x}-\boldsymbol{x}^{\prime})\right)\sigma_{\delta}(\boldsymbol{x}^{\prime})d\boldsymbol{x}^{\prime},
(𝜼δ𝜼)(𝒙)=Ωb(12𝜼(𝒙+𝒙)𝜼(𝒙)+12𝜼(𝒙𝒙))σδ(𝒙)d𝒙.(\nabla{\boldsymbol{\eta}}^{\delta}-\nabla\boldsymbol{\eta})(\boldsymbol{x})=\int_{\Omega_{b}}\left(\frac{1}{2}\nabla\boldsymbol{\eta}(\boldsymbol{x}+\boldsymbol{x}^{\prime})-\nabla\boldsymbol{\eta}(\boldsymbol{x})+\frac{1}{2}\nabla\boldsymbol{\eta}(\boldsymbol{x}-\boldsymbol{x}^{\prime})\right)\sigma_{\delta}(\boldsymbol{x}^{\prime})d\boldsymbol{x}^{\prime}.

For 𝒙Ωb,1\boldsymbol{x}\in\Omega_{b,1}, these points are at least δ\delta away from the boundary. Therefore, we have the following estimate for the discretized second derivative:

|12𝜼(𝒙+𝒙)𝜼(x)+12𝜼(𝒙𝒙)|Cδ2 for |𝒙|δ,\left|\frac{1}{2}\boldsymbol{\eta}(\boldsymbol{x}+\boldsymbol{x}^{\prime})-\boldsymbol{\eta}(x)+\frac{1}{2}\boldsymbol{\eta}(\boldsymbol{x}-\boldsymbol{x}^{\prime})\right|\leq C\delta^{2}\quad\text{ for }|\boldsymbol{x}^{\prime}|\leq\delta,

and similarly for 𝜼\nabla\boldsymbol{\eta}, by using the fact that 𝜼\boldsymbol{\eta} is spatially smooth in Ωb¯\overline{\Omega_{b}}. Therefore,

|(𝜼δ𝜼)(𝒙)|Cδ2,|(𝜼δ𝜼)(𝒙)|Cδ2, for 𝒙Ωb,1,|({\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta})(\boldsymbol{x})|\leq C\delta^{2},\quad|(\nabla{\boldsymbol{\eta}}^{\delta}-\nabla\boldsymbol{\eta})(\boldsymbol{x})|\leq C\delta^{2},\qquad\text{ for }\boldsymbol{x}\in\Omega_{b,1}, (145)

for a constant CC depending only on 𝜼\boldsymbol{\eta}.

For 𝒙Ωb,2\boldsymbol{x}\in\Omega_{b,2} we cannot use the same estimate, since after extending 𝜼\boldsymbol{\eta} to the larger domain Ω~b\tilde{\Omega}_{b}, the extended function on Ω~b\tilde{\Omega}_{b} does not necessarily have a continuous second derivative, as a result of the properties of odd extension, and in fact, there may be discontinuities of the second derivative along the boundary Ωb\partial\Omega_{b}. However, 𝜼\nabla\boldsymbol{\eta} on the larger domain Ω~b\tilde{\Omega}_{b} is still Lipschitz continuous. Thus, we instead use the equations:

(𝜼δ𝜼)(𝒙)=Ωb(𝜼(𝒙+𝒙)𝜼(𝒙))σδ(𝒙)d𝒙,({\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta})(\boldsymbol{x})=\int_{\Omega_{b}}(\boldsymbol{\eta}(\boldsymbol{x}+\boldsymbol{x}^{\prime})-\boldsymbol{\eta}(\boldsymbol{x}))\sigma_{\delta}(\boldsymbol{x}^{\prime})d\boldsymbol{x}^{\prime},
(𝜼δ𝜼)(𝒙)=Ωb(𝜼(𝒙+𝒙)𝜼(𝒙))σδ(𝒙)d𝒙.(\nabla{\boldsymbol{\eta}}^{\delta}-\nabla\boldsymbol{\eta})(\boldsymbol{x})=\int_{\Omega_{b}}(\nabla\boldsymbol{\eta}(\boldsymbol{x}+\boldsymbol{x}^{\prime})-\nabla\boldsymbol{\eta}(\boldsymbol{x}))\sigma_{\delta}(\boldsymbol{x}^{\prime})d\boldsymbol{x}^{\prime}.

Since 𝒙Ωb,2\boldsymbol{x}\in\Omega_{b,2}, even if |𝒙|δ|\boldsymbol{x}^{\prime}|\leq\delta, we may have that 𝒙+𝒙\boldsymbol{x}+\boldsymbol{x}^{\prime} is outside of Ωb\Omega_{b}. However, due to the Lipschitz continuity of 𝜼\nabla\boldsymbol{\eta} on the larger domain Ω~b\tilde{\Omega}_{b}, we still have the estimates

|𝜼(𝒙+𝒙)𝜼(𝒙)|Cδ,|𝜼(𝒙+𝒙)𝜼(𝒙)|Cδ, for 𝒙Ωb,2,|𝒙|δ,|\boldsymbol{\eta}(\boldsymbol{x}+\boldsymbol{x}^{\prime})-\boldsymbol{\eta}(\boldsymbol{x})|\leq C\delta,\quad|\nabla\boldsymbol{\eta}(\boldsymbol{x}+\boldsymbol{x}^{\prime})-\nabla\boldsymbol{\eta}(\boldsymbol{x})|\leq C\delta,\qquad\text{ for }\boldsymbol{x}\in\Omega_{b,2},|\boldsymbol{x^{\prime}}|\leq\delta,

which give

|(𝜼δ𝜼)(𝒙)|Cδ,|(𝜼δ𝜼)(𝒙)|Cδ, for 𝒙Ωb,2.|({\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta})(\boldsymbol{x})|\leq C\delta,\quad|(\nabla{\boldsymbol{\eta}}^{\delta}-\nabla\boldsymbol{\eta})(\boldsymbol{x})|\leq C\delta,\qquad\text{ for }\boldsymbol{x}\in\Omega_{b,2}. (146)

The area of Ωb,2\Omega_{b,2} is bounded by (2R+2L)δ(2R+2L)\delta, so by (145) and (146), we have ||𝜼δ𝜼||H1(Ωb)Cδ3/2||{\boldsymbol{\eta}}^{\delta}-\boldsymbol{\eta}||_{H^{1}(\Omega_{b})}\leq C\delta^{3/2} for a spatially smooth function 𝜼\boldsymbol{\eta} on Ωb¯\overline{\Omega_{b}}, where CC depends only on the norms of up to the second spatial derivative of 𝜼\boldsymbol{\eta} on Ωb¯\overline{\Omega_{b}}. The generalization of this result to a function 𝜼\boldsymbol{\eta} that also depends on time and is spatially smooth in both space and time follows analogously.

This completes the proof of the weak-classical consistency results. This proof effectively shows that the weak solutions that we have constructed to the regularized FPSI problem converge (in the energy norm on a uniform time interval) as the regularization parameter goes to zero to a classical solution of the original (non-regularized) FPSI problem when such a classical solution to the original FPSI problem exists.

11 Conclusions

In this manuscript we proved the existence of a weak solution to a fluid-structure interaction problem between the flow of an incompressible, viscous fluid and a multi-layered poroelastic/poroviscoelastic structure consisting of the Biot equations of poro(visco)elasticity and a thin, reticular interface with mass and elastic energy, which is transparent to fluid flow. The fluid and multilayered structure are nonlinearly coupled, giving rise to significant difficulties in the existence proof, associated with the geometric nonlinearity of the coupled problem. The existence proof is constructive, and it consists of two major steps. In the fist step we proved the existence of a weak solution to a regularized problem in the class of finite energy solutions. In the second step we showed that the solution of this regularized problem converges to a classical solution to the original, nonregularized probroblem as the regularization parameter tends to zero, as long as the original problem possesses a classical solution. While the proof of the existence of a weak solution to the regularized problem only requires that the Biot structure is poroelastic, additional regularity of the Biot poroelatic medium is required to prove the weak-classical consistency-the Biot structure is assumed to be poroviscoelastic. This weak-classical consistency result also shows that the solution we constructed is unique in the sense of weak-classical uniqueness.

We make a few comments about extensions of these results on fluid-poroelastic structure interaction to the case of three spatial dimensions, as the model problem discussed in this manuscript involves two spatial dimensions. For the existence proof, the constructive existence proof outlined for the two-dimensional FPSI problem carries out to the case of fluid-poroelastic structure interaction between a fluid modeled by the Navier-Stokes in three spatial dimensions and a three-dimensional Biot poroviscoelastic medium, separated by a two-dimensional reticular plate. In the course of such an analysis to a three-dimensional problem, one would encounter several new difficulties, which we briefly discuss here. First, the odd extension used to define the convolution in Definition 5.1 would have to be modified, but a similar odd extension could be used in three dimensions too. More importantly, the plate displacement in the finite energy space is in H2(Γ)H^{2}(\Gamma), which for a two-dimensional plate interface separating a 3D fluid and 3D Biot medium, would produce an interface that is only α\alpha-Hölder continuous for 0<α<1/20<\alpha<1/2. Thus, we would be working with time-dependent fluid domains Ωf(t)\Omega_{f}(t), which are not uniformly Lipschitz, which is a geometric requirement for many classical results such as the trace theorem. However, such problems have already been addressed in the fluid-structure interaction literature, for example in [51], and techniques exist for the analysis of FSI systems in three spatial dimensions, see for example [53] and [55]. Hence, the proof of constructive existence of a weak solution to a regularized 3D FPSI problem is expected to carry through similarly without significant challenges to give an analogue of Theorem 5.1 for an analogous 3D FPSI model.

However, we emphasize that it is still an open question to show weak-classical consistency for the 3D FPSI problem. Although one can still show existence of weak solutions to the regularized problem, using the current analysis, an analogue of Theorem 10.1 is at the moment unattainable. The issue is the rate of convergence in Lemma 10.2 of the convolution of the odd extension 𝜼δ\boldsymbol{\eta}^{\delta} to the original displacement 𝜼\boldsymbol{\eta} when 𝜼\boldsymbol{\eta} is a smooth function in time and space, which is on the order of δ3/2\delta^{3/2}. In two-dimensions, the usual convolution kernel σδ\sigma_{\delta} has an L2L^{2} norm on the order of δ1\delta^{-1}, so we get a crucial convergence to zero in the estimate (143) as δ0\delta\to 0 stating that the gradients of 𝜼δ\boldsymbol{\eta}_{\delta} and 𝜼δδ\boldsymbol{\eta}^{\delta}_{\delta} converge pointwise uniformly in the limit as δ0\delta\to 0, which is the estimate that allows us to close our bootstrap argument. In three dimensions, we would lose this convergence to zero since in three dimensions, σδ\sigma_{\delta} has an L2L^{2} norm on the order of δ3/2\delta^{-3/2} while the convolution estimate would still give a convergence rate of δ3/2\delta^{3/2} for the L2L^{2} norm of (𝜼𝜼δ)(\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}). Hence, to establish a corresponding weak-classical consistency result for the three dimensional problem, one would either have to improve the convergence rate of the convolution 𝜼δ\boldsymbol{\eta}^{\delta} to the actual function 𝜼\boldsymbol{\eta} in Lemma 10.2, or find an alternative regularization/extension procedure in place of that in Definition 5.1 that exhibits a better convergence rate than δ3/2\delta^{3/2} as in Lemma 10.2.

A final interesting extension of this work is to consider the singular limit as the thin interface thickness converges to zero, and analyze the resulting FPSI problem between the Navier-Stokes equations for an incompressible, viscous fluid and the Biot equations, nonlinearly coupled over the moving interface, without a reticular plate separating the two. Preliminary results indicate that this will be possible under certain assumptions, including viscoelasticity of the Biot medium. In this case, we believe that one could obtain an analogous result for existence of a weak solution, first to a regularized problem, either through an adaptation of the methods presented in this manuscript for the case with a reticular plate, or as a singular limit of weak solutions to the regularized problem with a plate as the plate thickness goes to zero. Then, under the assumption of the existence of a classical solution to the FPSI problem with direct Biot-fluid contact (and no plate), one could pursue a similar weak-classical consistency result showing convergence of weak solutions (to the regularized problem) to the classical solution as the regularization parameter goes to zero. Consequently, we would have a weak-classical consistency result of the same type as in the current paper for the FPSI model without a reticular plate.

Acknowledgements

We would like to thank the anonymous referees for their careful reading of our manuscript and for providing insightful comments that improved the quality of this paper. Partial support for this research was provided by the NSF MSPRF fellowship DMS-2303177 (Jeffrey Kuan), the NSF grants DMS-2247000 and DMS-2011319 (Sunčica Čanić and Jeffrey Kuan), and the Croatian Science Foundation under project number IP-2022-10-2962 (Boris Muha).

Appendix A Appendix

A.1 Weak continuity of solutions to the regularized FPSI problem

In this appendix, we show a result related to weak continuity of solutions to the regularized FPSI problem, namely, we will show that as ν0\nu\to 0:

Ωf,δ(0)𝒖^(0)(𝒖δ)ν(0)Ωf,δ(0)|𝒖0|2, and Ωf,δ(t)𝒖^(t)(𝒖δ)ν(t)Ωf,δ(t)𝒖^(t)𝒖δ(t),\int_{\Omega_{f,\delta}(0)}\widehat{\boldsymbol{u}}(0)\cdot(\boldsymbol{u}_{\delta})_{\nu}(0)\to\int_{\Omega_{f,\delta}(0)}|\boldsymbol{u}_{0}|^{2},\quad\text{ and }\quad\int_{\Omega_{f,\delta}(t)}\widehat{\boldsymbol{u}}(t)\cdot(\boldsymbol{u}_{\delta})_{\nu}(t)\to\int_{\Omega_{f,\delta}(t)}\widehat{\boldsymbol{u}}(t)\cdot\boldsymbol{u}_{\delta}(t),

for almost all points 0<tT0<t\leq T.

This result will be used in Section A.3 to estimate the first term T1T_{1} in (124) in the Gronwall’s estimate. We will show weak continuity through the following series of lemmas.

Lemma A.1.

Let ωL(0,T;H02(Γ))W1,(0,T;L2(Γ))\omega\in L^{\infty}(0,T;H_{0}^{2}(\Gamma))\cap W^{1,\infty}(0,T;L^{2}(\Gamma)) with

mint[0,T],x[0,L]R+ω(t,x)>0,\min_{t\in[0,T],x\in[0,L]}R+\omega(t,x)>0,

define the moving fluid domain Ωfω(t)\Omega_{f}^{\omega}(t). Then, given 𝒖L2(0,T;H1(Ωfω(t)))L(0,T;L2(Ωfω(t)))\boldsymbol{u}\in L^{2}(0,T;H^{1}(\Omega_{f}^{\omega}(t)))\cap L^{\infty}(0,T;L^{2}(\Omega_{f}^{\omega}(t))) where Ωfω(t)={(x,y)2:0xL,Ryω(t,x)}\Omega_{f}^{\omega}(t)=\{(x,y)\in\mathbb{R}^{2}:0\leq x\leq L,-R\leq y\leq\omega(t,x)\}, we have that

||𝒖ν(t,x,y)𝒖(t,x,y)||L2(Ωfω(t))0 as ν0,||\boldsymbol{u}_{\nu}(t,x,y)-\boldsymbol{u}(t,x,y)||_{L^{2}(\Omega_{f}^{\omega}(t))}\to 0\qquad\text{ as }\nu\to 0,

for almost all t[0,T]t\in[0,T].

Proof.

Recall that in the case of real-valued functions, one shows convergence of the convolution to the function itself almost everywhere by using the Lebesgue differentiation theorem [32]. To apply the theorem in this context, we need to apply it to a function taking values in a fixed Banach space rather than a time-dependent Banach space.

As a result, we consider the following function,

𝒗(t,x,y)=K(t,0,x,y)𝒖(t,x,R+ω(t,x)R+ω(0,x)(R+y)R),\boldsymbol{v}(t,x,y)=K(t,0,x,y)\boldsymbol{u}\left(t,x,\frac{R+\omega(t,x)}{R+\omega(0,x)}(R+y)-R\right),

where we have pulled the fluid velocity back to the fixed initial domain Ωωf(0)\Omega^{\omega}_{f}(0). We recall the definition of K(s,t,z,r)K(s,t,z,r) from (119) and its inverse:

K(s,t,x,y)=(R+ω(s,x)R+ω(t,x)0(R+y)x(R+ω(s,x)R+ω(t,x))1),K1(s,t,x,y)=(R+ω(t,x)R+ω(s,x)0(R+y)R+ω(t,x)R+ω(s,x)x(R+ω(s,x)R+ω(t,x))1).K(s,t,x,y)=\begin{pmatrix}\frac{R+\omega(s,x)}{R+\omega(t,x)}&0\\ -(R+y)\partial_{x}\left(\frac{R+\omega(s,x)}{R+\omega(t,x)}\right)&1\\ \end{pmatrix},\quad K^{-1}(s,t,x,y)=\begin{pmatrix}\frac{R+\omega(t,x)}{R+\omega(s,x)}&0\\ (R+y)\frac{R+\omega(t,x)}{R+\omega(s,x)}\partial_{x}\left(\frac{R+\omega(s,x)}{R+\omega(t,x)}\right)&1\\ \end{pmatrix}.

By the uniform boundedness of R+ω(t,x)R+\omega(t,x) and |xω(t,x)||\partial_{x}\omega(t,x)|, and mint[0,T],x[0,L]R+ω(t,x)>0\min_{t\in[0,T],x\in[0,L]}R+\omega(t,x)>0, it is immediate to see that 𝒗(t,z,r)\boldsymbol{v}(t,z,r) is in L(0,T;L2(Ωωf(0)))L^{\infty}(0,T;L^{2}(\Omega^{\omega}_{f}(0))), where we emphasize that L2(Ωωf(0))L^{2}(\Omega^{\omega}_{f}(0)) is a fixed function space that no longer depends on time.

By Lebesgue’s differentiation theorem, almost every t[0,T]t\in[0,T] is a Lebesgue point satisfying

limν012νtνt+ν||𝒗(t,)𝒗(s,)||L2(Ωωf(0))ds0.\lim_{\nu\to 0}\frac{1}{2\nu}\int_{t-\nu}^{t+\nu}||\boldsymbol{v}(t,\cdot)-\boldsymbol{v}(s,\cdot)||_{L^{2}(\Omega^{\omega}_{f}(0))}ds\to 0. (147)

Recall that by definition (120),

𝒖ν(t,x,y)=K(s,t,x,y)𝒖(s,x,R+ω(s,x)R+ω(t,x)(R+y)R)jν(ts)ds.\boldsymbol{u}_{\nu}(t,x,y)=\int_{\mathbb{R}}K(s,t,x,y)\boldsymbol{u}\left(s,x,\frac{R+\omega(s,x)}{R+\omega(t,x)}(R+y)-R\right)j_{\nu}(t-s)ds.

Thus, we compute

𝒖ν(t,x,y)𝒖(t,x,y)=\displaystyle\boldsymbol{u}_{\nu}(t,x,y)-\boldsymbol{u}(t,x,y)= (K(s,t,x,y)𝒖(s,x,R+ω(s,x)R+ω(t,x)(R+y)R)𝒖(t,x,y))jν(ts)ds\displaystyle\int_{\mathbb{R}}\left(K(s,t,x,y)\boldsymbol{u}\left(s,x,\frac{R+\omega(s,x)}{R+\omega(t,x)}(R+y)-R\right)-\boldsymbol{u}(t,x,y)\right)\cdot j_{\nu}(t-s)ds
:=\displaystyle:= I1+I2,\displaystyle I_{1}+I_{2},

where

I1=\displaystyle I_{1}=\int_{\mathbb{R}} K1(t,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)\displaystyle K^{-1}\left(t,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\cdot
(𝒗(s,x,R+ω(0,x)R+ω(t,x)(R+y)R)𝒗(t,x,R+ω(0,x)R+ω(t,x)(R+y)R))jν(ts)ds,\displaystyle\left(\boldsymbol{v}\left(s,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)-\boldsymbol{v}\left(t,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right)j_{\nu}(t-s)ds,
I2=\displaystyle I_{2}=\int_{\mathbb{R}} (K(s,t,x,y)K1(s,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)K1(t,0,x,R+ω(0,x)R+ω(t,x)(R+y)R))\displaystyle\left(K(s,t,x,y)K^{-1}\left(s,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)-K^{-1}\left(t,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right)\cdot
𝒗(s,x,R+ω(0,x)R+ω(t,x)(R+y)R)jν(ts)ds.\displaystyle\boldsymbol{v}\left(s,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)j_{\nu}(t-s){{ds}}.

We estimate each of these terms as follows. For I1I_{1}, we compute that

K1(t,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)=(R+ω(0,x)R+ω(t,x)0(R+y)(R+ω(0,x)R+ω(t,x))2x(R+ω(t,x)R+ω(0,x))1),K^{-1}\left(t,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)=\begin{pmatrix}\frac{R+\omega(0,x)}{R+\omega(t,x)}&0\\ (R+y)\left(\frac{R+\omega(0,x)}{R+\omega(t,x)}\right)^{2}\partial_{x}\left(\frac{R+\omega(t,x)}{R+\omega(0,x)}\right)&1\\ \end{pmatrix},

which we note is uniformly bounded on [0,T][0,T]. Hence, using the fact that |jν(ts)|1ν|j_{\nu}(t-s)|\leq\frac{1}{\nu}, we get

||I1||L2(Ωωf(t))\displaystyle\small||I_{1}||_{L^{2}(\Omega^{\omega}_{f}(t))} C1νtνt+ν||𝒗(s,x,R+ω(0,x)R+ω(t,x)(R+y)R)𝒗(t,x,R+ω(0,x)R+ω(t,x)(R+y)R)||L2(Ωωf(t))ds\displaystyle\leq C\cdot\frac{1}{\nu}\int_{t-\nu}^{t+\nu}\left|\left|\boldsymbol{v}\left(s,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)-\boldsymbol{v}\left(t,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right|\right|_{L^{2}(\Omega^{\omega}_{f}(t))}ds
C1νtνt+ν(R+ω(t,x)R+ω(0,x))1/2||𝒗(s,x,y)𝒗(t,x,y)||L2(Ωωf(0))ds0,\displaystyle\leq C\cdot\frac{1}{\nu}\int_{t-\nu}^{t+\nu}\left(\frac{R+\omega(t,x)}{R+\omega(0,x)}\right)^{1/2}||\boldsymbol{v}(s,x,y)-\boldsymbol{v}(t,x,y)||_{L^{2}(\Omega^{\omega}_{f}(0))}ds\to 0,

as ν0\nu\to 0 if tt is a Lebesgue point, by (147) and the uniform boundedness of R+ω(t,x)R+ω(0,x)\frac{R+\omega(t,x)}{R+\omega(0,x)} on [0,T][0,T].

To estimate I2I_{2}, we can use the continuity in time of ω\omega and xω\partial_{x}\omega to calculate that

|K(s,t,x,y)K1(s,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)K1(t,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)|0,\left|K(s,t,x,y)K^{-1}\left(s,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)-K^{-1}\left(t,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right|\to 0,

uniformly in (x,y)(x,y) as sts\to t. Now, we estimate

||I2||L2(Ωωf(t))\displaystyle\small||I_{2}||_{L^{2}(\Omega^{\omega}_{f}(t))} maxx,yΩωf(t)|K(s,t,x,y)K1(s,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)\displaystyle\leq\int_{\mathbb{R}}\max_{x,y\in\Omega^{\omega}_{f}(t)}\left|K(s,t,x,y)K^{-1}\left(s,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right.
K1(t,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)|\displaystyle\left.\quad-K^{-1}\left(t,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right|
||𝒗(s,x,R+ω(0,x)R+ω(t,x)(R+y)R)||L2(Ωωf(t))jν(ts)ds\displaystyle\quad\cdot\left|\left|\boldsymbol{v}\left(s,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right|\right|_{L^{2}(\Omega^{\omega}_{f}(t))}\cdot j_{\nu}(t-s)ds
maxx,yΩωf(t)|K(s,t,x,y)K1(s,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)\displaystyle\leq\int_{\mathbb{R}}\max_{x,y\in\Omega^{\omega}_{f}(t)}\left|K(s,t,x,y)K^{-1}\left(s,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right.
K1(t,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)|\displaystyle\quad\left.-K^{-1}\left(t,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right|
(R+ω(t,x)R+ω(0,x))1/2||𝒗(s,x,y)||L2(Ωωf(0))jν(ts)ds\displaystyle\quad\cdot\left(\frac{R+\omega(t,x)}{R+\omega(0,x)}\right)^{1/2}\cdot\left|\left|\boldsymbol{v}\left(s,x,y\right)\right|\right|_{L^{2}(\Omega^{\omega}_{f}(0))}\cdot j_{\nu}(t-s)ds
Cmaxx,yΩωf(t)|K(s,t,x,y)K1(s,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)\displaystyle\leq C\int_{\mathbb{R}}\max_{x,y\in\Omega^{\omega}_{f}(t)}\left|K(s,t,x,y)K^{-1}\left(s,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right.
K1(t,0,x,R+ω(0,x)R+ω(t,x)(R+y)R)|jν(ts)ds,\displaystyle\left.\quad-K^{-1}\left(t,0,x,\frac{R+\omega(0,x)}{R+\omega(t,x)}(R+y)-R\right)\right|\cdot j_{\nu}(t-s)ds,

where we used the fact that 𝒗L(0,T;L2(Ωωf(0)))\boldsymbol{v}\in L^{\infty}(0,T;L^{2}(\Omega^{\omega}_{f}(0))). Thus, we conclude that ||I2||L2(Ωωf(t))0||I_{2}||_{L^{2}(\Omega^{\omega}_{f}(t))}\to 0 as ν0\nu\to 0. This completes the proof. ∎

We also have a weak continuity lemma, which states that the value of 𝒖δ\boldsymbol{u}_{\delta} tested against any function in the fluid function space has a continuity property as t0t\to 0.

Lemma A.2.

Consider an arbitrary 𝒒C1(0,T;Vf,δ(t))\boldsymbol{q}\in C^{1}(0,T;V_{f,\delta}(t)) and the weak solution 𝒖δ\boldsymbol{u}_{\delta} to the regularized problem for arbitrary δ\delta, where Vf,δ(t)V_{f,\delta}(t) is defined by the displacement ωδ\omega_{\delta} and (27). There exists a measure zero subset SS of [0,T][0,T] (depending on δ\delta) such that

limt0,t[0,T]ScΩf,δ(t)𝒖δ(t)𝒒(t)=Ωf,δ(0)𝒖0𝒒(0).\lim_{t\to 0,t\in[0,T]\cap S^{c}}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\boldsymbol{q}(t)=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\boldsymbol{q}(0).
Proof.

Consider the following function for each τ[0,T]\tau\in[0,T] and α>0\alpha>0, given by

Jτ,ν(t)=10tjν(sτ)ds,J_{\tau,\nu}(t)=1-\int_{0}^{t}j_{\nu}(s-\tau)ds, (148)

and note that Jτ,ν(t)=jν(tτ)J_{\tau,\nu}^{\prime}(t)=-j_{\nu}(t-\tau). We want to test the regularized weak formulation for 𝒖δ\boldsymbol{u}_{\delta} with the test function Jτ,ν(t)𝒒J_{\tau,\nu}(t)\boldsymbol{q} for certain admissible choices of τ\tau. To see which τ\tau we want to choose, we define the function

𝒘(t,x,y)=R+ωδ(t)R+ωδ(0)𝒖δ(t,x,R+ωδ(t)R+ωδ(0)(R+y)R)𝒒(t,x,R+ωδ(t)R+ωδ(0)(R+y)R).\boldsymbol{w}(t,x,y)=\frac{R+\omega_{\delta}(t)}{R+\omega_{\delta}(0)}\cdot\boldsymbol{u}_{\delta}\left(t,x,\frac{R+\omega_{\delta}(t)}{R+\omega_{\delta}(0)}(R+y)-R\right)\cdot\boldsymbol{q}\left(t,x,\frac{R+\omega_{\delta}(t)}{R+\omega_{\delta}(0)}(R+y)-R\right).

We claim that 𝒘L(0,T;L1(Ωf,δ(0)))\boldsymbol{w}\in L^{\infty}(0,T;L^{1}(\Omega_{f,\delta}(0))). To see this, we compute by a change of variables that

||𝒘(t,x,y)||L1(Ωf,δ(0))=Ωf,δ(t)|𝒖δ(t,x,y)𝒒(t,x,y)|,||\boldsymbol{w}(t,x,y)||_{L^{1}(\Omega_{f,\delta}(0))}=\int_{\Omega_{f,\delta}(t)}|\boldsymbol{u}_{\delta}(t,x,y)\cdot\boldsymbol{q}(t,x,y)|,

and we then use the fact that 𝒖δ,𝒒L(0,T;L2(Ωf,δ(t)))\boldsymbol{u}_{\delta},\boldsymbol{q}\in L^{\infty}(0,T;L^{2}(\Omega_{f,\delta}(t))).

Hence, by the Lebesgue differentiation theorem, there exists a measurable subset S[0,T]S\subset[0,T] of measure zero such that every point in [0,T]Sc[0,T]\cap S^{c} is a Lebesgue point of 𝒘\boldsymbol{w}, in the sense that

limν012ντντ+ν||𝒘(τ,)𝒘(s,)||L1(Ωf,δ(0))ds0.\lim_{\nu\to 0}\frac{1}{2\nu}\int_{\tau-\nu}^{\tau+\nu}||\boldsymbol{w}(\tau,\cdot)-\boldsymbol{w}(s,\cdot)||_{L^{1}(\Omega_{f,\delta}(0))}ds\to 0. (149)

for every τ[0,T]Sc\tau\in[0,T]\cap S^{c}. These are the τ\tau for which we will consider the test function Jτ,ν(t)𝒒J_{\tau,\nu}(t)\boldsymbol{q}. For the test functions for the Biot medium and the plate, we will take these test functions to be zero. Hence, in the regularized weak formulation (10.1), we will test with (𝒗,φ,𝝍,r)=(Jτ,ν(t)𝒒,0,0,0)(\boldsymbol{v},\varphi,\boldsymbol{\psi},r)=(J_{\tau,\nu}(t)\boldsymbol{q},0,0,0).

Hence, we obtain the following equality:

0TΩf,δ(t)𝒖δt(Jτ,ν(t)𝒒)+120TΩf,δ(t)[((𝒖δ)𝒖δ)(Jτ,ν(t)𝒒)((𝒖δ)(Jτ,ν(t)𝒒))𝒖δ]\displaystyle-\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot\partial_{t}(J_{\tau,\nu}(t)\boldsymbol{q})+\frac{1}{2}\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}[((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta})\cdot(J_{\tau,\nu}(t)\boldsymbol{q})-((\boldsymbol{u}_{\delta}\cdot\nabla)(J_{\tau,\nu}(t)\boldsymbol{q}))\cdot\boldsymbol{u}_{\delta}]
+120TΓδ(t)(𝒖δ𝒏2𝝃δ𝒏)𝒖δ(Jτ,ν(t)𝒒)+2ν0TΩf,δ(t)𝑫(𝒖δ):𝑫(Jτ,ν(t)𝒒)\displaystyle+\frac{1}{2}\int_{0}^{T}\int_{\Gamma_{\delta}(t)}(\boldsymbol{u}_{\delta}\cdot\boldsymbol{n}-2\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n})\boldsymbol{u}_{\delta}\cdot(J_{\tau,\nu}(t)\boldsymbol{q})+2\nu\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{D}(\boldsymbol{u}_{\delta}):\boldsymbol{D}(J_{\tau,\nu}(t)\boldsymbol{q})
0TΓδ(t)(12|𝒖δ|2pδ)Jτ,ν(t)qnβ0TΓδ(t)[(ξδ)τ(uδ)τ]Jτ,ν(t)qτ=Ωf,δ(0)𝒖0Jτ,ν(0)𝒒(0).\displaystyle-\int_{0}^{T}\int_{\Gamma_{\delta}(t)}\left(\frac{1}{2}|\boldsymbol{u}_{\delta}|^{2}-p_{\delta}\right)J_{\tau,\nu}(t)q_{n}-\beta\int_{0}^{T}\int_{\Gamma_{\delta}(t)}[(\xi_{\delta})_{\tau}-(u_{\delta})_{\tau}]\cdot J_{\tau,\nu}(t)q_{\tau}=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot J_{\tau,\nu}(0)\boldsymbol{q}(0).

Consider τ(0,T)Sc\tau\in(0,T)\cap S^{c}. We want to pass to the limit as ν0\nu\to 0, and then pass to the limit as τ0\tau\to 0, in order to obtain the desired result.

First, we pass to the limit as ν0\nu\to 0. We handle the convergences as follows.

First term: We will show that because τ\tau is a Lebesgue point of 𝒘\boldsymbol{w},

0TΩf,δ(t)𝒖δt(Jτ,ν(t)𝒒)Ωf,δ(τ)𝒖δ(τ)𝒒(τ)0tΩf,δ(t)𝒖δt𝒒, as ν0.-\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot\partial_{t}(J_{\tau,\nu}(t)\boldsymbol{q})\to\int_{\Omega_{f,\delta}(\tau)}\boldsymbol{u}_{\delta}(\tau)\boldsymbol{q}(\tau)-\int_{0}^{t}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot\partial_{t}\boldsymbol{q},\qquad\text{ as }\nu\to 0.

We compute that

0TΩf,δ(t)𝒖δt(Jτ,ν(t)𝒒)=0TΩf,δ(t)𝒖δjν(tτ)𝒒0TΩf,δ(t)𝒖δJτ,ν(t)t𝒒.-\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot\partial_{t}(J_{\tau,\nu}(t)\boldsymbol{q})=\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot j_{\nu}(t-\tau)\boldsymbol{q}-\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot J_{\tau,\nu}(t)\partial_{t}\boldsymbol{q}.

It is easy to see that

0TΩf,δ(t)𝒖δJτ,ν(t)t𝒒0tΩf,δ(t)𝒖δt𝒒.\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}J_{\tau,\nu}(t)\partial_{t}\boldsymbol{q}\to\int_{0}^{t}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\partial_{t}\boldsymbol{q}.

So it remains to show that

0TΩf,δ(t)𝒖δjν(tτ)𝒒Ωf,δ(τ)𝒖δ(τ)𝒒(τ), as ν0.\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot j_{\nu}(t-\tau)\boldsymbol{q}\to\int_{\Omega_{f,\delta}(\tau)}\boldsymbol{u}_{\delta}(\tau)\boldsymbol{q}(\tau),\qquad\text{ as }\nu\to 0.

By a change of variables, we compute that

0TΩf,δ(t)𝒖δjν(tτ)𝒒=0TΩf,δ(τ)R+ωδ(t)R+ωδ(τ)𝒖δ(t,x,R+ωδ(t)R+ωδ(τ)(R+y)R)jν(tτ)𝒒(t,x,R+ωδ(t)R+ωδ(τ)(R+y)R)=0TΩf,δ(τ)R+ωδ(0)R+ωδ(τ)𝒘(t,x,R+ωδ(0)R+ωδ(τ)(R+y)R)jν(tτ)=0TΩf,δ(0)𝒘(t,x,y)jν(tτ).\int_{0}^{T}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot j_{\nu}(t-\tau)\boldsymbol{q}\\ =\int_{0}^{T}\int_{\Omega_{f,\delta}(\tau)}\frac{R+\omega_{\delta}(t)}{R+\omega_{\delta}(\tau)}\cdot\boldsymbol{u}_{\delta}\left(t,x,\frac{R+\omega_{\delta}(t)}{R+\omega_{\delta}(\tau)}(R+y)-R\right)\cdot j_{\nu}(t-\tau)\boldsymbol{q}\left(t,x,\frac{R+\omega_{\delta}(t)}{R+\omega_{\delta}(\tau)}(R+y)-R\right)\\ =\int_{0}^{T}\int_{\Omega_{f,\delta}(\tau)}\frac{R+\omega_{\delta}(0)}{R+\omega_{\delta}(\tau)}\boldsymbol{w}\left(t,x,\frac{R+\omega_{\delta}(0)}{R+\omega_{\delta}(\tau)}(R+y)-R\right)\cdot j_{\nu}(t-\tau)=\int_{0}^{T}\int_{\Omega_{f,\delta}(0)}\boldsymbol{w}(t,x,y)\cdot j_{\nu}(t-\tau).

By (149), we have that

0TΩf,δ(0)𝒘(t,x,y)jν(tτ)Ωf,δ(0)𝒘(τ,x,y)=Ωf,δ(τ)𝒖δ(τ)𝒒(τ),\int_{0}^{T}\int_{\Omega_{f,\delta}(0)}\boldsymbol{w}(t,x,y)\cdot j_{\nu}(t-\tau)\to\int_{\Omega_{f,\delta}(0)}\boldsymbol{w}(\tau,x,y)=\int_{\Omega_{f,\delta}(\tau)}\boldsymbol{u}_{\delta}(\tau)\cdot\boldsymbol{q}(\tau),

which establishes the desired convergence.

Final term: It is immediate to see that for all sufficiently small ν>0\nu>0,

Ωf,δ(0)𝒖0Jτ,ν(0)𝒒(0)=Ωf,δ(0)𝒖0𝒒(0).\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot J_{\tau,\nu}(0)\boldsymbol{q}(0)=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\boldsymbol{q}(0).

We can now easily take ν0\nu\to 0 in the remaining terms to obtain that for any τ(0,T)Sc\tau\in(0,T)\cap S^{c},

Ωf,δ(τ)𝒖δ(τ)𝒒(τ)0tΩf,δ(t)𝒖δt𝒒+120tΩf,δ(t)[((𝒖δ)𝒖δ)𝒒((𝒖δ)𝒒)𝒖δ]+120tΓδ(t)(𝒖δ𝒏2𝝃δ𝒏)𝒖δ𝒒+2ν0tΩf,δ(t)𝑫(𝒖δ):𝑫(𝒒)0tΓδ(t)(12|𝒖δ|2pδ)qnβ0tΓδ(t)[(ξδ)τ(uδ)τ]qτ=Ωf,δ(0)𝒖0𝒒(0).\int_{\Omega_{f,\delta}(\tau)}\boldsymbol{u}_{\delta}(\tau)\cdot\boldsymbol{q}(\tau)-\int_{0}^{t}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}\cdot\partial_{t}\boldsymbol{q}+\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(t)}[((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta})\cdot\boldsymbol{q}-((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{q})\cdot\boldsymbol{u}_{\delta}]\\ +\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(t)}(\boldsymbol{u}_{\delta}\cdot\boldsymbol{n}-2\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n})\boldsymbol{u}_{\delta}\cdot\boldsymbol{q}+2\nu\int_{0}^{t}\int_{\Omega_{f,\delta}(t)}\boldsymbol{D}(\boldsymbol{u}_{\delta}):\boldsymbol{D}(\boldsymbol{q})\\ -\int_{0}^{t}\int_{\Gamma_{\delta}(t)}\left(\frac{1}{2}|\boldsymbol{u}_{\delta}|^{2}-p_{\delta}\right)q_{n}-\beta\int_{0}^{t}\int_{\Gamma_{\delta}(t)}[(\xi_{\delta})_{\tau}-(u_{\delta})_{\tau}]\cdot q_{\tau}=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\boldsymbol{q}(0).

Passing to the limit as τ0\tau\to 0 with τ(0,T)Sc\tau\in(0,T)\cap S^{c} gives the desired result.

Lemma A.3.

Let 𝒖0\boldsymbol{u}_{0} be divergence free and smooth on Ωf(0)¯\overline{\Omega_{f}(0)}. Define

𝒒~(t,x,y)=Kδ(0,t,x,y)𝒖0(x,R+ωδ(0,x)R+ωδ(t,x)(R+y)R),\tilde{\boldsymbol{q}}(t,x,y)=K_{\delta}(0,t,x,y)\boldsymbol{u}_{0}\left(x,\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(t,x)}(R+y)-R\right), (150)

where KδK_{\delta} is given by (119). Then, there exists a sequence of functions 𝒒~mC1c(0,T;Vf,δ(t))\tilde{\boldsymbol{q}}_{m}\in C^{1}_{c}(0,T;V_{f,\delta}(t)), with Vf,δ(t)V_{f,\delta}(t) determined by the plate displacement ωδ\omega_{\delta} via the definition (27), such that

max0tT||𝒒~𝒒~m||L2(Ωf,δ(t))0, as m.\max_{0\leq t\leq T}||\tilde{\boldsymbol{q}}-\tilde{\boldsymbol{q}}_{m}||_{L^{2}(\Omega_{f,\delta}(t))}\to 0,\qquad\text{ as }m\to\infty.
Proof.

There exists a rectangular two-dimensional maximal domain ΩM\Omega_{M} of the form [0,L]×[R,Rmax][0,L]\times[-R,R_{max}] for some positive constant RmaxR_{max} that contains all of the domains Ωf,δ(t)\Omega_{f,\delta}(t) for t[0,T]t\in[0,T]. We will extend 𝒒~\tilde{\boldsymbol{q}} to the maximal spacetime domain [0,T]×ΩM[0,T]\times\Omega_{M} by extending vertically in the radial direction by the trace of 𝒒~\tilde{\boldsymbol{q}} along Γδ(t)\Gamma_{\delta}(t). In particular, we define

𝒒~(t,x,y)=K(0,t,x,ωδ(t,x))𝒖0(x,ωδ(0,x)), for (t,x,y)([0,T]×ΩM)([0,T]×Ωf,δ(t)).\small\tilde{\boldsymbol{q}}(t,x,y)=K(0,t,x,\omega_{\delta}(t,x))\boldsymbol{u}_{0}\left(x,\omega_{\delta}(0,x)\right),\qquad\text{ for }(t,x,y)\in([0,T]\times\Omega_{M})-([0,T]\times\Omega_{f,\delta}(t)). (151)

Note that this extension preserves the divergence free property.

We have the following two claims about the extended function, considered as a function on the fixed maximal domain ΩM\Omega_{M}. First, we claim that 𝒒~L(0,T;H1(ΩM))\tilde{\boldsymbol{q}}\in L^{\infty}(0,T;H^{1}(\Omega_{M})). Second, we claim that 𝒒~C(0,T;L2(ΩM))\tilde{\boldsymbol{q}}\in C(0,T;L^{2}(\Omega_{M})). To see that 𝒒~L(0,T;H1(ΩM))\tilde{\boldsymbol{q}}\in L^{\infty}(0,T;H^{1}(\Omega_{M})), we note that ωδ\omega_{\delta} and xωδ\partial_{x}\omega_{\delta} are bounded uniformly pointwise, and furthermore 𝒖0\boldsymbol{u}_{0} and its first spatial derivatives are bounded by assumption. In addition, x2ωδL(0,T;L2(Γ))\partial_{x}^{2}\omega_{\delta}\in L^{\infty}(0,T;L^{2}(\Gamma)), which allows us to conclude that 𝒒~L(0,T;H1(ΩM))\tilde{\boldsymbol{q}}\in L^{\infty}(0,T;H^{1}(\Omega_{M})).

Next, we want to verify that 𝒒~C(0,T;L2(ΩM))\tilde{\boldsymbol{q}}\in C(0,T;L^{2}(\Omega_{M})). Consider any t[0,T]t\in[0,T] and consider any s[0,T]s\in[0,T] with sts\neq t. We define the following regions:

A(s,t)=ΩMf(Ωf,δ(s)Ωf,δ(t))c,\displaystyle A(s,t)=\Omega^{M}_{f}\cap(\Omega_{f,\delta}(s)\cup\Omega_{f,\delta}(t))^{c},\ B(s,t)=[Ωf,δ(s)(Ωf,δ(t))c][(Ωf,δ(s))cΩf,δ(t)],\displaystyle B(s,t)=[\Omega_{f,\delta}(s)\cap(\Omega_{f,\delta}(t))^{c}]\cup[(\Omega_{f,\delta}(s))^{c}\cap\Omega_{f,\delta}(t)],
C(s,t)=Ωf,δ(s)Ωf,δ(t).\displaystyle C(s,t)=\Omega_{f,\delta}(s)\cap\Omega_{f,\delta}(t).

Consider ϵ>0\epsilon>0. We want to find h>0h>0 such that

||𝒒~(t,)𝒒~(s,)||2L2(ΩM)ϵ, for all s(th,t+h)[0,T].||\tilde{\boldsymbol{q}}(t,\cdot)-\tilde{\boldsymbol{q}}(s,\cdot)||^{2}_{L^{2}(\Omega_{M})}\leq\epsilon,\qquad\text{ for all }s\in(t-h,t+h)\cap[0,T]. (152)

We compute that

||𝒒~(t,)𝒒~(s,)||L2(ΩM)2\displaystyle||\tilde{\boldsymbol{q}}(t,\cdot)-\tilde{\boldsymbol{q}}(s,\cdot)||_{L^{2}(\Omega_{M})}^{2} =A(s,t)|𝒒~(t,x,y)𝒒~(s,x,y)|2+B(s,t)|𝒒~(t,x,y)𝒒~(s,x,y)|2\displaystyle=\int_{A(s,t)}|\tilde{\boldsymbol{q}}(t,x,y)-\tilde{\boldsymbol{q}}(s,x,y)|^{2}+\int_{B(s,t)}|\tilde{\boldsymbol{q}}(t,x,y)-\tilde{\boldsymbol{q}}(s,x,y)|^{2}
+C(s,t)|𝒒~(t,x,y)𝒒~(s,x,y)|2=IA+IB+IC.\displaystyle+\int_{C(s,t)}|\tilde{\boldsymbol{q}}(t,x,y)-\tilde{\boldsymbol{q}}(s,x,y)|^{2}=I_{A}+I_{B}+I_{C}. (153)

We estimate each of the terms IAI_{A}, IBI_{B}, and ICI_{C} separately.

For IAI_{A}, we recall that we are extending by the trace as in (151) on A(s,t)A(s,t), so we have that

IA=A(s,t)|Kδ(0,t,x,ωδ(t,x))Kδ(0,s,x,ωδ(s,x))|2|𝒖0(x,ωδ(0,x))|2.I_{A}=\int_{A(s,t)}|K_{\delta}(0,t,x,\omega_{\delta}(t,x))-K_{\delta}(0,s,x,\omega_{\delta}(s,x))|^{2}\cdot|\boldsymbol{u}_{0}(x,\omega_{\delta}(0,x))|^{2}.

We have that |𝒖0(x,ωδ(0,x))|M1|\boldsymbol{u}_{0}(x,\omega_{\delta}(0,x))|\leq M_{1} for some constant M1M_{1} by the fact that 𝒖0\boldsymbol{u}_{0} is continuous on Ωf(0)¯\overline{\Omega_{f}(0)}. By continuity, we can choose h>0h>0 sufficiently small so that

|Kδ(0,t,x,ωδ(t,x))Kδ(0,s,x,ωδ(s,x))|2<ϵ3M12(R+Rmax)L, for all s(th,t+h)[0,T].|K_{\delta}(0,t,x,\omega_{\delta}(t,x))-K_{\delta}(0,s,x,\omega_{\delta}(s,x))|^{2}<\frac{\epsilon}{3M_{1}^{2}(R+R_{max})L},\qquad\text{ for all }s\in(t-h,t+h)\cap[0,T].

Thus, for all s(th,t+h)[0,T]s\in(t-h,t+h)\cap[0,T],

IA|A(s,t)|ϵ3(R+Rmax)Lϵ3.I_{A}\leq|A(s,t)|\cdot\frac{\epsilon}{3(R+R_{max})L}\leq\frac{\epsilon}{3}.

For IBI_{B}, we will use the fact that ωδ\omega_{\delta} does not change much in time over small time intervals, by continuity. We note that there exists a uniform constant M2M_{2} such that |𝒒~|M2|\tilde{\boldsymbol{q}}|\leq M_{2} on [0,T]×ΩM[0,T]\times\Omega_{M}. Hence,

IB=B(s,t)|𝒒~(t,z,r)𝒒~(s,z,r)|2|B(s,t)|4M22=4M220L|ωδ(t,x)ωδ(s,x)|dx.I_{B}=\int_{B(s,t)}|\tilde{\boldsymbol{q}}(t,z,r)-\tilde{\boldsymbol{q}}(s,z,r)|^{2}\leq|B(s,t)|\cdot 4M_{2}^{2}=4M_{2}^{2}\int_{0}^{L}|\omega_{\delta}(t,x)-\omega_{\delta}(s,x)|dx.

Because ωδL(0,T;H20(Γ))W1,(0,T;L2(Γ))\omega_{\delta}\in L^{\infty}(0,T;H^{2}_{0}(\Gamma))\cap W^{1,\infty}(0,T;L^{2}(\Gamma)), there exists h>0h>0 sufficiently small such that

|ωδ(t,x)ωδ(s,x)|ϵ12M22L, for all x[0,L] and s(th,t+h)[0,T].|\omega_{\delta}(t,x)-\omega_{\delta}(s,x)|\leq\frac{\epsilon}{12M_{2}^{2}L},\qquad\text{ for all }x\in[0,L]\text{ and }s\in(t-h,t+h)\cap[0,T].

This allows us to conclude that IBϵ3I_{B}\leq\frac{\epsilon}{3}, for all s(th,t+h)[0,T]s\in(t-h,t+h)\cap[0,T].

For ICI_{C}, we refer to the definition of 𝒒~\tilde{\boldsymbol{q}} in (150) and note that Kδ(0,t,x,y)K_{\delta}(0,t,x,y) is continuous in time uniformly in (x,y)[0,L]×[R,Rmax](x,y)\in[0,L]\times[-R,R_{max}], 𝒖0\boldsymbol{u}_{0} is uniformly continuous as a function on Ωf(0)¯\overline{\Omega_{f}(0)}, and ωδ(t,x)\omega_{\delta}(t,x) is continuous in time uniformly in x[0,L]x\in[0,L]. Hence, there exists h>0h>0 sufficiently small such that

|𝒒~(t,x,y)𝒒~(s,x,y)|2ϵ3(R+Rmax)L, for all (x,y)C(s,t) and s(th,t+h)[0,T],|\tilde{\boldsymbol{q}}(t,x,y)-\tilde{\boldsymbol{q}}(s,x,y)|^{2}\leq\frac{\epsilon}{3(R+R_{max})L},\qquad\text{ for all }(x,y)\in C(s,t)\text{ and }s\in(t-h,t+h)\cap[0,T],

which gives the desired result that ICϵ3I_{C}\leq\frac{\epsilon}{3} for all s(th,t+h)[0,T]s\in(t-h,t+h)\cap[0,T]. Thus, by using (A.1), we have established (152).

Since 𝒒~L(0,T;H1(ΩM))C(0,T;L2(ΩM))\tilde{\boldsymbol{q}}\in L^{\infty}(0,T;H^{1}(\Omega_{M}))\cap C(0,T;L^{2}(\Omega_{M})), we can extend 𝒒~\tilde{\boldsymbol{q}} to a continuous function on all of \mathbb{R} as follows. We can find an increasing sequence TmT_{m} with TmTT_{m}\to T as mm\to\infty, such that 𝒒~(Tm)H1(ΩM)\tilde{\boldsymbol{q}}(T_{m})\in H^{1}(\Omega_{M}) for all mm. Define an extension 𝒒^m\hat{\boldsymbol{q}}_{m} for each mm to all of \mathbb{R} by 𝒒^m=𝒒~\hat{\boldsymbol{q}}_{m}=\tilde{\boldsymbol{q}} if t[0,Tm]t\in[0,T_{m}],

𝒒^m=𝒒~(0), if t<0,𝒒^m=𝒒~(Tm), if t>Tm.\hat{\boldsymbol{q}}_{m}=\tilde{\boldsymbol{q}}(0),\qquad\text{ if }t<0,\qquad\hat{\boldsymbol{q}}_{m}=\tilde{\boldsymbol{q}}(T_{m}),\qquad\text{ if }t>T_{m}.

Define

𝒒~m=𝒒^mj1/m,\tilde{\boldsymbol{q}}_{m}=\hat{\boldsymbol{q}}_{m}*j_{1/m},

where the convolution is a convolution in time with jνj_{\nu} for α=1/m\alpha=1/m. Because 𝒒^mL(0,T;H1(ΩM))C(0,T;L2(ΩM))\hat{\boldsymbol{q}}_{m}\in L^{\infty}(0,T;H^{1}(\Omega_{M}))\cap C(0,T;L^{2}(\Omega_{M})) with 𝒒^m\hat{\boldsymbol{q}}_{m} being divergence free for every t[0,T]t\in[0,T], we have that 𝒒~m\tilde{\boldsymbol{q}}_{m} restricted to t[0,T]{t}×Ωf,δ(t)\bigcup_{t\in[0,T]}\{t\}\times\Omega_{f,\delta}(t) gives a function in C1([0,T);Vf,δ(t))C^{1}([0,T);V_{f,\delta}(t)), where Vf,δ(t)V_{f,\delta}(t) is the space defined in (27) with the plate displacement ωδ\omega_{\delta}. The fact that

max0tT||𝒒~𝒒~m||L2(Ωf,δ(t))0, as m,\max_{0\leq t\leq T}||\tilde{\boldsymbol{q}}-\tilde{\boldsymbol{q}}_{m}||_{L^{2}(\Omega_{f,\delta}(t))}\to 0,\qquad\text{ as }m\to\infty,

follows from the uniform continuity of 𝒒~\tilde{\boldsymbol{q}} on [0,T][0,T] as a function taking values in L2(ΩM)L^{2}(\Omega_{M}), convergence properties of convolutions, and the fact that 𝒒~C(0,T;L2(ΩM))\tilde{\boldsymbol{q}}\in C(0,T;L^{2}(\Omega_{M})) which gives the convergence

maxt[Tm,T]||𝒒~(T)𝒒~(t)||L2(ΩM)0, as m.\max_{t\in[T_{m},T]}||\tilde{\boldsymbol{q}}(T)-\tilde{\boldsymbol{q}}(t)||_{L^{2}(\Omega_{M})}\to 0,\qquad\text{ as }m\to\infty.

Lemma A.4.

For the function 𝒒~\tilde{\boldsymbol{q}} defined in (150), there exists a measure zero subset SS of [0,T][0,T] such that

limt0,t[0,T]ScΩf,δ(t)𝒖δ(t)𝒒~(t)=Ωf,δ(0)𝒖0𝒒~(0).\lim_{t\to 0,t\in[0,T]\cap S^{c}}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}(t)=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}(0).
Proof.

Note that because t𝒒~\partial_{t}\tilde{\boldsymbol{q}} is not necessarily in H1(Ωf,δ(t))H^{1}(\Omega_{f,\delta}(t)), 𝒒~\tilde{\boldsymbol{q}} is not a valid test function. Thus, we use the sequence 𝒒~mC1(0,T;𝒱f,δ(t))\tilde{\boldsymbol{q}}_{m}\in C^{1}(0,T;\mathcal{V}_{f,\delta}(t)) from Lemma A.3, which satisfies

max0tT||𝒒~𝒒~m||L2(Ωf,δ(t))0, as m.\max_{0\leq t\leq T}||\tilde{\boldsymbol{q}}-\tilde{\boldsymbol{q}}_{m}||_{L^{2}(\Omega_{f,\delta}(t))}\to 0,\qquad\text{ as }m\to\infty.

We can then apply Lemma A.2 to each of the test functions 𝒒~m\tilde{\boldsymbol{q}}_{m}, to deduce that there exists a measure zero subset SmS_{m} of [0,T][0,T] such that

limt0,t[0,T]SmcΩf,δ(t)𝒖δ(t)𝒒~m(t)=Ωf,δ(0)𝒖0𝒒~m(0).\lim_{t\to 0,t\in[0,T]\cap S_{m}^{c}}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}_{m}(t)=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}_{m}(0).

In addition, by uniform boundedness, 𝒖δL(0,T;L2(Ωf,δ(t)))\boldsymbol{u}_{\delta}\in L^{\infty}(0,T;L^{2}(\Omega_{f,\delta}(t))), and hence, there exists a measure zero subset S0S_{0} of [0,T][0,T], and a positive constant CC such that ||𝒖0||L2(Ωf,δ(0))C||\boldsymbol{u}_{0}||_{L^{2}(\Omega_{f,\delta}(0))}\leq C, and

||𝒖δ(t)||L2(Ωf,δ(t))C, for all tS0c.||\boldsymbol{u}_{\delta}(t)||_{L^{2}(\Omega_{f,\delta}(t))}\leq C,\qquad\text{ for all }t\in S_{0}^{c}. (154)

Define S=S0m1SmS=S_{0}\cup\bigcup_{m\geq 1}S_{m}, which is also a measure zero subset of [0,T][0,T]. Then, for each mm,

limt0,t[0,T]ScΩf,δ(t)𝒖δ(t)𝒒~m(t)=Ωf,δ(0)𝒖0𝒒~m(0).\lim_{t\to 0,t\in[0,T]\cap S^{c}}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}_{m}(t)=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}_{m}(0). (155)

By passing to the limit in mm, we claim that in addition,

limt0,t[0,T]ScΩf,δ(t)𝒖δ(t)𝒒~(t)=Ωf,δ(0)𝒖0𝒒~(0).\lim_{t\to 0,t\in[0,T]\cap S^{c}}\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}(t)=\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}(0).

To see this, consider ϵ>0\epsilon>0. We claim that there exists h>0h>0 sufficiently small such that for all t(0,h)Sct\in(0,h)\cap S^{c},

|Ωf,δ(t)𝒖δ(t)𝒒~(t)Ωf,δ(0)𝒖0𝒒~(0)|<ϵ.\left|\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}(t)-\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}(0)\right|<\epsilon.

We can choose MM sufficiently large such that max0tT||𝒒~𝒒~M||L2(Ωf,δ(t))<ϵ3C\displaystyle\max_{0\leq t\leq T}||\tilde{\boldsymbol{q}}-\tilde{\boldsymbol{q}}_{M}||_{L^{2}(\Omega_{f,\delta}(t))}<\frac{\epsilon}{3C}, where CC is defined by (154). Therefore, for all t[0,T]Sct\in[0,T]\cap S^{c},

|Ωf,δ(t)𝒖δ(t)𝒒~(t)Ωf,δ(t)𝒖δ(t)𝒒~M(t)|<ϵ3.\left|\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}(t)-\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}_{M}(t)\right|<\frac{\epsilon}{3}.

In addition,

|Ωf,δ(0)𝒖0𝒒~(0)Ωf,δ(0)𝒖0𝒒~M(0)|<ϵ3.\left|\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}(0)-\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}_{M}(0)\right|<\frac{\epsilon}{3}.

By applying (155) with m=Mm=M, we can choose h>0h>0 sufficiently small such that for all t(0,h)Sct\in(0,h)\cap S^{c},

|Ωf,δ(t)𝒖δ(t)𝒒~M(t)Ωf,δ(0)𝒖0𝒒~M(0)|<ϵ3.\left|\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}_{M}(t)-\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}_{M}(0)\right|<\frac{\epsilon}{3}.

Thus, by applying the triangle inequality, we have that for all t(0,h)Sct\in(0,h)\cap S^{c},

|Ωf,δ(t)𝒖δ(t)𝒒~(t)Ωf,δ(0)𝒖0𝒒~(0)|<ϵ,\left|\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\tilde{\boldsymbol{q}}(t)-\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}(0)\right|<\epsilon,

which establishes the desired result.

We can now prove the final result of this appendix. We recall the definition of 𝒖^\widehat{\boldsymbol{u}} from (110).

Lemma A.5.

In the limit as ν0\nu\to 0 we have the following convergence results:

Ωf,δ(0)𝒖^(0)(𝒖δ)ν(0)Ωf,δ(0)|𝒖0|2, and Ωf,δ(t)𝒖^(t)(𝒖δ)ν(t)Ωf,δ(t)𝒖^(t)𝒖δ(t),\int_{\Omega_{f,\delta}(0)}\widehat{\boldsymbol{u}}(0)\cdot(\boldsymbol{u}_{\delta})_{\nu}(0)\to\int_{\Omega_{f,\delta}(0)}|\boldsymbol{u}_{0}|^{2},\quad\text{ and }\quad\int_{\Omega_{f,\delta}(t)}\widehat{\boldsymbol{u}}(t)\cdot(\boldsymbol{u}_{\delta})_{\nu}(t)\to\int_{\Omega_{f,\delta}(t)}\widehat{\boldsymbol{u}}(t)\cdot\boldsymbol{u}_{\delta}(t),

for almost all points t(0,T]t\in(0,T].

Proof.

The second convergence for almost all points t(0,T]t\in(0,T] follows directly from Lemma A.1 and the fact that 𝒖^L(0,T;L2(Ωf,δ(t)))\widehat{\boldsymbol{u}}\in L^{\infty}(0,T;L^{2}(\Omega_{f,\delta}(t))).

So we just need to verify the convergence at t=0t=0. To do this, we note that 𝒖^(0)=𝒖0\widehat{\boldsymbol{u}}(0)=\boldsymbol{u}_{0}. Hence,

Ωf,δ(0)\displaystyle\int_{\Omega_{f,\delta}(0)} 𝒖^(0)(𝒖δ)ν(0)\displaystyle\widehat{\boldsymbol{u}}(0)\cdot(\boldsymbol{u}_{\delta})_{\nu}(0)
=Ωω0(Kδ(s,0,x,y)𝒖δ(s,x,R+ωδ(s,x)R+ωδ(0,x)(R+y)R)jδ(ts)ds)𝒖0(x,y)dxdy\displaystyle=\int_{\Omega^{\omega_{0}}}\left(\int_{\mathbb{R}}K_{\delta}(s,0,x,y)\boldsymbol{u}_{\delta}\left(s,x,\frac{R+\omega_{\delta}(s,x)}{R+\omega_{\delta}(0,x)}(R+y)-R\right)j_{\delta}(t-s)ds\right)\boldsymbol{u}_{0}(x,y)dxdy
=(Ωω0Kδ(s,0,x,y)𝒖δ(s,x,R+ωδ(s,x)R+ωδ(0,x)(R+y)R)𝒖0(x,y)dxdy)jν(ts)ds\displaystyle=\int_{\mathbb{R}}\left(\int_{\Omega^{\omega_{0}}}K_{\delta}(s,0,x,y)\boldsymbol{u}_{\delta}\left(s,x,\frac{R+\omega_{\delta}(s,x)}{R+\omega_{\delta}(0,x)}(R+y)-R\right)\cdot\boldsymbol{u}_{0}(x,y)dxdy\right)j_{\nu}(t-s)ds
=(Ωf,δ(s)𝒖δ(s,x,y)R+ωδ(0,x)R+ωδ(s,x)Kδt(s,0,x,R+ωδ(0,x)R+ωδ(s,x)(R+y)R)\displaystyle=\int_{\mathbb{R}}\left(\int_{\Omega_{f,\delta}(s)}\boldsymbol{u}_{\delta}(s,x,y)\cdot\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}K_{\delta}^{t}\left(s,0,x,\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}(R+y)-R\right)\right.
𝒖0(x,R+ωδ(0,x)R+ωδ(s,x)(R+y)R)dxdy)jν(ts)ds.\displaystyle\quad\left.\cdot\boldsymbol{u}_{0}\left(x,\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}(R+y)-R\right)dxdy\right)j_{\nu}(t-s)ds.

We compute

R+ωδ(0,x)R+ωδ(s,x)Kδt(s,0,x,R+ωδ(0,x)R+ωδ(s,x)(R+y)R)=(1(R+y)(R+ωδ(0,x)R+ωδ(s,x))0R+ωδ(0,x)R+ωδ(s,x))\displaystyle\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}\cdot K_{\delta}^{t}\left(s,0,x,\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}(R+y)-R\right)=\begin{pmatrix}1&(R+y)\nabla\left(\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}\right)\\ 0&\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}\\ \end{pmatrix}
=(R+ωδ(0,x)R+ωδ(s,x)0(R+y)(R+ωδ(0,x)R+ωδ(s,x))1)+(1R+ωδ(0,x)R+ωδ(s,x)(R+y)(R+ωδ(0,x)R+ωδ(s,x))(R+y)(R+ωδ(0,x)R+ωδ(s,x))R+ωδ(0,x)R+ωδ(s,x)1)\displaystyle=\begin{pmatrix}\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}&0\\ -(R+y)\nabla\left(\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}\right)&1\\ \end{pmatrix}+\begin{pmatrix}1-\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}&(R+y)\nabla\left(\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}\right)\\ (R+y)\nabla\left(\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}\right)&\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}-1\\ \end{pmatrix}
:=Kδ(0,s,x,y)+Rδ(0,s,x,y).\displaystyle:=K_{\delta}(0,s,x,y)+R_{\delta}(0,s,x,y).

Hence,

Ωf,δ(0)𝒖^(0)(𝒖δ)ν(0)\displaystyle\int_{\Omega_{f,\delta}(0)}\widehat{\boldsymbol{u}}(0)\cdot(\boldsymbol{u}_{\delta})_{\nu}(0)
=(Ωf,δ(s)𝒖δ(s,x,y)Kδ(0,s,x,y)𝒖0(x,R+ωδ(0,x)R+ωδ(s,x)(R+y)R)dxdy)jν(ts)ds\displaystyle=\int_{\mathbb{R}}\left(\int_{\Omega_{f,\delta}(s)}\boldsymbol{u}_{\delta}(s,x,y)\cdot K_{\delta}(0,s,x,y)\boldsymbol{u}_{0}\left(x,\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}(R+y)-R\right)dxdy\right)j_{\nu}(t-s)ds
+(Ωf,δ(s)𝒖δ(s,x,y)Rδ(0,s,x,y)𝒖0(x,R+ωδ(0,x)R+ωδ(s,x)(R+y)R)dxdy)jν(ts)ds=IK,δ+IR,δ.\displaystyle+\int_{\mathbb{R}}\left(\int_{\Omega_{f,\delta}(s)}\boldsymbol{u}_{\delta}(s,x,y)\cdot R_{\delta}(0,s,x,y)\boldsymbol{u}_{0}\left(x,\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}(R+y)-R\right)dxdy\right)j_{\nu}(t-s)ds=I_{K,\delta}+I_{R,\delta}.

Note that

IK,δ=(Ωf,δ(s)𝒖δ(s,x,y)𝒒~(s,x,y)dxdy)jν(ts)dsI_{K,\delta}=\int_{\mathbb{R}}\left(\int_{\Omega_{f,\delta}(s)}\boldsymbol{u}_{\delta}(s,x,y)\cdot\tilde{\boldsymbol{q}}(s,x,y)dxdy\right)j_{\nu}(t-s)ds

where 𝒒~\tilde{\boldsymbol{q}} is defined by (150). Since 𝒖δ(s)=𝒖δ(s)\boldsymbol{u}_{\delta}(s)=\boldsymbol{u}_{\delta}(-s) so that ωδ(s)=ωδ(s)\omega_{\delta}(s)=\omega_{\delta}(-s) for s0s\leq 0 (see the extension procedure), we conclude by Lemma A.4 that

IK,δΩf,δ(0)𝒖0𝒒~(0)=Ωf,δ(0)|𝒖0|2, as ν0.I_{K,\delta}\to\int_{\Omega_{f,\delta}(0)}\boldsymbol{u}_{0}\cdot\tilde{\boldsymbol{q}}(0)=\int_{\Omega_{f,\delta}(0)}|\boldsymbol{u}_{0}|^{2},\qquad\text{ as }\nu\to 0.

So it suffices to show that IR,δ0I_{R,\delta}\to 0 as ν0\nu\to 0. This follows from the fact that |Rδ|0|R_{\delta}|\to 0 uniformly as s0s\to 0. In particular,

Ωf,δ(s)|𝒖δ(s,x,y)𝒖0(x,R+ωδ(0,x)R+ωδ(s,x)(R+y)R)|dxdyC, for almost all s[0,T],\int_{\Omega_{f,\delta}(s)}\left|\boldsymbol{u}_{\delta}(s,x,y)\cdot\boldsymbol{u}_{0}\left(x,\frac{R+\omega_{\delta}(0,x)}{R+\omega_{\delta}(s,x)}(R+y)-R\right)\right|dxdy\leq C,\qquad\text{ for almost all }s\in[0,T],

by the boundedness of 𝒖δL(0,T;L2(Ωf,δ(t))\boldsymbol{u}_{\delta}\in L^{\infty}(0,T;L^{2}(\Omega_{f,\delta}(t)) and the fact that 𝒖0\boldsymbol{u}_{0} is uniformly bounded. In addition, by the continuity properties of ωδ\omega_{\delta} in time, we have that

max(x,y)Ωf,δ(s)¯|Rδ(0,s,x,y)|0, as s0,\max_{(x,y)\in\overline{\Omega_{f,\delta}(s)}}|R_{\delta}(0,s,x,y)|\to 0,\qquad\text{ as }s\to 0,

which implies that IR,δ0I_{R,\delta}\to 0 as ν0\nu\to 0. This completes the proof.

We will use this result in the next section to estimate the first term T1T_{1}, see (124) in the Gronwall’s estimate.

A.2 Gronwall’s terms estimates

In this appendix we provide details of the derivation of the terms appearing in (124) and the calculations providing the desired estimates of the terms in (124) used to prove Gronwall’s estimate in Section 10.3.

Term T1. To derive term T1T_{1}, defined in (126), we first multiply the weak formulation (10.1) for 𝒖\boldsymbol{u} with the test function 𝒗=𝒖(𝒖ˇδ)ν\boldsymbol{v}=\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu} to obtain the terms:

T1,1\displaystyle T_{1,1} =0tΩf(s)𝒖t[𝒖(𝒖ˇδ)ν]120tΓ(s)(𝝃𝒏)𝒖[𝒖(𝒖ˇδ)ν]\displaystyle=-\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{u}\cdot\partial_{t}\left[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}\right]-\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}]
+Ωf(t)𝒖(t)[𝒖(𝒖ˇδ)ν](t)Ωf(0)𝒖(0)[𝒖(𝒖ˇδ)ν](0),\displaystyle+\int_{\Omega_{f}(t)}\boldsymbol{u}(t)\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}](t)-\int_{\Omega_{f}(0)}\boldsymbol{u}(0)\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}](0),

where Ωf(0)\Omega_{f}(0) is the fluid domain corresponding to the initial structure displacement ω0\omega_{0}. We note that 𝒖\boldsymbol{u} is smooth in time and (𝒖ˇδ)ν(\widecheck{\boldsymbol{u}}_{\delta})_{\nu} is differentiable in time as a result of the time convolution. Thus, by the Reynold’s transport theorem,

T1,1=0tΩf(s)t𝒖[𝒖(𝒖ˇδ)ν]+120tΓ(s)(𝝃𝒏)𝒖[𝒖(𝒖ˇδ)ν].T_{1,1}=\int_{0}^{t}\int_{\Omega_{f}(s)}\partial_{t}\boldsymbol{u}\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}]+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot[\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu}].

Because 𝒖\boldsymbol{u} is smooth and by the weak convergence properties of (𝒖ˇδ)ν(\widecheck{\boldsymbol{u}}_{\delta})_{\nu} in Proposition 10.1,

T1,1=0tΩf(s)t𝒖[𝒖𝒖ˇδ]+120tΓ(s)(𝝃𝒏)𝒖[𝒖𝒖ˇδ]+K1,1,ν,T_{1,1}=\int_{0}^{t}\int_{\Omega_{f}(s)}\partial_{t}\boldsymbol{u}\cdot[\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta}]+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot[\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta}]+K_{1,1,\nu},

where K1,1,ν0K_{1,1,\nu}\to 0 as ν0\nu\to 0. Using estimates as found in [63], we can transfer the first integral from Ω1(s)\Omega_{1}(s) to Ωf,δ(s)\Omega_{f,\delta}(s) at the cost of an additional term, so that

T1,1=0tΩf,δ(s)t𝒖^(𝒖^𝒖δ)+120tΓ(s)(𝝃𝒏)𝒖(𝒖𝒖ˇδ)+R~1+K1,1,ν,T_{1,1}=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\partial_{t}\widehat{\boldsymbol{u}}\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})+{{\tilde{R}_{1}}}+K_{1,1,\nu},

where

|R~1|ϵ0t||𝒖^𝒖δ||2H1(Ωf,δ(s))+C(ϵ)(0t||ωωδ||2H2(Γ)+0t||tωtωδ||L2(Γ)2+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).{{|\tilde{R}_{1}|}}\leq\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{H^{1}(\Omega_{f,\delta}(s))}\\ +C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\partial_{t}\omega-\partial_{t}\omega_{\delta}||_{L^{2}(\Gamma)}^{2}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Thus, by using Proposition 10.1 again,

T1,1=0tΩf,δ(s)t𝒖^(𝒖^(𝒖δ)ν)+120tΓ(s)(𝝃𝒏)𝒖(𝒖(𝒖ˇδ)ν)+R~1+K1,1,ν,T_{1,1}=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\partial_{t}\widehat{\boldsymbol{u}}\cdot(\widehat{\boldsymbol{u}}-(\boldsymbol{u}_{\delta})_{\nu})+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot(\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu})+{{\tilde{R}_{1}+K_{1,1,\nu}}}, (156)

where K1,1,ν0K_{1,1,\nu}\to 0 as ν0\nu\to 0.

Next, we test the regularized weak formulation for 𝒖δ\boldsymbol{u}_{\delta} with 𝒖^\widehat{\boldsymbol{u}} and obtain the following terms:

T1,2=0tΩf,δ(s)𝒖δt𝒖^120tΓδ(s)(𝝃δ𝒏δ)𝒖δ𝒖^+Ωf,δ(t)𝒖δ(t)𝒖^(t)Ωf(0)𝒖δ(0)𝒖^(0).\displaystyle T_{1,2}=-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\boldsymbol{u}_{\delta}\cdot\partial_{t}\widehat{\boldsymbol{u}}-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta})\boldsymbol{u}_{\delta}\cdot\widehat{\boldsymbol{u}}+\int_{\Omega_{f,\delta}(t)}\boldsymbol{u}_{\delta}(t)\cdot\widehat{\boldsymbol{u}}(t)-\int_{\Omega_{f}(0)}\boldsymbol{u}_{\delta}(0)\cdot\widehat{\boldsymbol{u}}(0).

We want to integrate by parts in time, but 𝒖δ\boldsymbol{u}_{\delta} is not necessarily smooth in time. Thus, we replace 𝒖δ\boldsymbol{u}_{\delta} by its time regularization (𝒖δ)ν(\boldsymbol{u}_{\delta})_{\nu} at the cost of a term K1,2,νK_{1,2,\nu} which goes to zero as ν0\nu\to 0 by Proposition 10.1. Combining this with the Reynold’s transport theorem, we get:

T1,2=0tΩf,δ(s)t[(𝒖δ)ν]𝒖^+120tΓδ(s)(𝝃δ𝒏δ)(𝒖δ)ν𝒖^+K1,2,ν,T_{1,2}=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\partial_{t}\left[(\boldsymbol{u}_{\delta})_{\nu}\right]\cdot\widehat{\boldsymbol{u}}+\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta})(\boldsymbol{u}_{\delta})_{\nu}\cdot\widehat{\boldsymbol{u}}+K_{1,2,\nu}, (157)

where K1,2,ν0K_{1,2,\nu}\to 0 as ν0\nu\to 0.

Now, from the energy inequality, we obtain the terms

T1,3=12Ωf,δ(t)|𝒖δ(t)|212Ωf,δ(0)|𝒖δ(0)|2.T_{1,3}=\frac{1}{2}\int_{\Omega_{f,\delta}(t)}|\boldsymbol{u}_{\delta}(t)|^{2}-\frac{1}{2}\int_{\Omega_{f,\delta}(0)}|\boldsymbol{u}_{\delta}(0)|^{2}. (158)

Using the Reynold’s transport theorem, the total contribution T1=T1,1T1,2+T1,3T_{1}=T_{1,1}-T_{1,2}+T_{1,3} is

T1\displaystyle T_{1} =12Ωf,δ(t)|𝒖^(t)|212Ωf,δ(0)|𝒖^(0)|2Ωf,δ(t)(𝒖^(𝒖δ)ν)(t)+Ωf,δ(0)(𝒖^(𝒖δ)ν)(0)\displaystyle=\frac{1}{2}\int_{\Omega_{f,\delta}(t)}|\widehat{\boldsymbol{u}}(t)|^{2}-\frac{1}{2}\int_{\Omega_{f,\delta}(0)}|\widehat{\boldsymbol{u}}(0)|^{2}-\int_{\Omega_{f,\delta}(t)}(\widehat{\boldsymbol{u}}\cdot(\boldsymbol{u}_{\delta})_{\nu})(t)+\int_{\Omega_{f,\delta}(0)}(\widehat{\boldsymbol{u}}\cdot(\boldsymbol{u}_{\delta})_{\nu})(0)
+12Ωf,δ(t)|𝒖δ(t)|212Ωf,δ(0)|𝒖δ(0)|2120tΓδ(s)(𝝃δ𝒏δ)𝒖^(𝒖^(𝒖δ)ν)\displaystyle+\frac{1}{2}\int_{\Omega_{f,\delta}(t)}|\boldsymbol{u}_{\delta}(t)|^{2}-\frac{1}{2}\int_{\Omega_{f,\delta}(0)}|\boldsymbol{u}_{\delta}(0)|^{2}-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta})\widehat{\boldsymbol{u}}\cdot(\widehat{\boldsymbol{u}}-(\boldsymbol{u}_{\delta})_{\nu})
+120tΓ(s)(𝝃𝒏)𝒖(𝒖(𝒖ˇδ)ν)+R~1+K1,1,ν+K1,2,ν.\displaystyle+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot(\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu})+{{\tilde{R}_{1}+K_{1,1,\nu}}}+K_{1,2,\nu}.

By Proposition 10.1, (𝒖δ)ν(\boldsymbol{u}_{\delta})_{\nu} and (𝒖ˇδ)ν(\widecheck{\boldsymbol{u}}_{\delta})_{\nu} converge weakly to 𝒖δ\boldsymbol{u}_{\delta} and 𝒖ˇδ\widecheck{\boldsymbol{u}}_{\delta} respectively, weakly in L2(0,T,W1,p(Ωf,δ(t)))L^{2}(0,T,W^{1,p}(\Omega_{f,\delta}(t))) and L2(0,T,W1,p(Ωf,1(t)))L^{2}(0,T,W^{1,p}(\Omega_{f,1}(t))) for all p[1,2)p\in[1,2). Furthermore, by Lemma A.5 proved in the appendix above, we have that

Ωf,δ(0)(𝒖^(𝒖δ)ν)(0)Ωf,δ(0)(𝒖^𝒖δ)(0),Ωf,δ(t)(𝒖^(𝒖δ)ν)(t)Ωf,δ(t)(𝒖^𝒖δ)(t).\int_{\Omega_{f,\delta}(0)}(\widehat{\boldsymbol{u}}\cdot(\boldsymbol{u}_{\delta})_{\nu})(0)\to\int_{\Omega_{f,\delta}(0)}(\widehat{\boldsymbol{u}}\cdot\boldsymbol{u}_{\delta})(0),\quad\int_{\Omega_{f,\delta}(t)}(\widehat{\boldsymbol{u}}\cdot(\boldsymbol{u}_{\delta})_{\nu})(t)\to\int_{\Omega_{f,\delta}(t)}(\widehat{\boldsymbol{u}}\cdot\boldsymbol{u}_{\delta})(t). (159)

Thus, taking the limit as ν0\nu\to 0, the contribution of this term is now

T1\displaystyle T_{1} =12Ωf,δ(s)|(𝒖^𝒖δ)(t)|212Ωf,δ(0)|(𝒖^𝒖δ)(0)|2\displaystyle=\frac{1}{2}\int_{\Omega_{f,\delta}(s)}|(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})(t)|^{2}-\frac{1}{2}\int_{\Omega_{f,\delta}(0)}|(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})(0)|^{2}
120tΓδ(s)(𝝃δ𝒏δ)𝒖^(𝒖^𝒖δ)+120tΓ(s)(𝝃𝒏)𝒖(𝒖𝒖ˇδ)+R~1.\displaystyle-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta})\widehat{\boldsymbol{u}}\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})+{{\tilde{R}_{1}}}.

Since 𝒖^(0)=𝒖δ(0)=𝒖0\widehat{\boldsymbol{u}}(0)=\boldsymbol{u}_{\delta}(0)=\boldsymbol{u}_{0}, we obtain after some standard estimates that

T1=12Ωf,δ(t)|(𝒖^𝒖δ)(t)|2+R1,T_{1}=\frac{1}{2}\int_{\Omega_{f,\delta}(t)}|(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})(t)|^{2}+{{R_{1}}},

where

|R1|\displaystyle{{|R_{1}|}} ϵ0t||𝒖^𝒖δ||2H1(Ωf,δ(s))\displaystyle\leq\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{H^{1}(\Omega_{f,\delta}(s))}
+C(ϵ)(0t||ωωδ||2H2(Γ)+0t||tωtωδ||L2(Γ)2+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\partial_{t}\omega-\partial_{t}\omega_{\delta}||_{L^{2}(\Gamma)}^{2}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

This completes the calculations associated with term T1T_{1}.

Term T2. To estimate term T2T_{2}, defined in (127) above, we notice that since (𝒖ˇδ)ν(\widecheck{\boldsymbol{u}}_{\delta})_{\nu} converges weakly to 𝒖ˇδ\widecheck{\boldsymbol{u}}_{\delta} in L2(0,T;W1,p(Ωf,δ(t)))L^{2}(0,T;W^{1,p}(\Omega_{f,\delta}(t))) for p[1,2)p\in[1,2) by Proposition 10.1, and because 𝒖\boldsymbol{u} is smooth, as ν0\nu\to 0, we have that T2T_{2} converges to

T2\displaystyle T_{2} :=120tΩf(s)((𝒖)𝒖)(𝒖𝒖ˇδ)120tΩf(s)((𝒖)(𝒖𝒖ˇδ))𝒖\displaystyle:=\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f}(s)}((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})-\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f}(s)}((\boldsymbol{u}\cdot\nabla)(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta}))\cdot\boldsymbol{u}
120tΩf,δ(s)((𝒖δ)𝒖δ)(𝒖^𝒖δ)+120tΩf,δ(s)((𝒖δ)(𝒖^𝒖δ))𝒖δ.\displaystyle-\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta})\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})+\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}((\boldsymbol{u}_{\delta}\cdot\nabla)(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}))\cdot\boldsymbol{u}_{\delta}.

We note that the quantity 120tΩf,δ(s)((𝒖δ)𝒖δ)𝒖δ,\displaystyle\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}((\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta})\cdot\boldsymbol{u}_{\delta}, is well-defined because 𝒖δL(0,T;L2(Ωf,δ(t)))L2(0,T;H1(Ωf,δ(t)))\boldsymbol{u}_{\delta}\in L^{\infty}(0,T;L^{2}(\Omega_{f,\delta}(t)))\cap L^{2}(0,T;H^{1}(\Omega_{f,\delta}(t))), which by interpolation is in L4(0,T;H1/2(Ωf,δ(t)))L^{4}(0,T;H^{1/2}(\Omega_{f,\delta}(t))), and hence by Sobolev inequalities embeds into L4(0,T;L4(Ωf,δ(t)))L^{4}(0,T;L^{4}(\Omega_{f,\delta}(t))).

We want to transfer the integrals

0tΩf(s)((𝒖)𝒖)(𝒖𝒖ˇδ),0tΩf(s)((𝒖)(𝒖𝒖ˇδ))𝒖,\int_{0}^{t}\int_{\Omega_{f}(s)}((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta}),\qquad\int_{0}^{t}\int_{\Omega_{f}(s)}((\boldsymbol{u}\cdot\nabla)(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta}))\cdot\boldsymbol{u}, (160)

to integrals on Ωf,δ(s)\Omega_{f,\delta}(s) by using the map ψδ(s,):Ωf,δ(s)Ωf(s)\psi_{\delta}(s,\cdot):\Omega_{f,\delta}(s)\to\Omega_{f}(s) defined by (106). We use

𝒖^=γδJδ1(𝒖ψδ),𝒖^𝒖δ=γδJδ1((𝒖𝒖ˇδ)ψδ),\widehat{\boldsymbol{u}}=\gamma_{\delta}J_{\delta}^{-1}\cdot(\boldsymbol{u}\circ\psi_{\delta}),\qquad\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}=\gamma_{\delta}J_{\delta}^{-1}\cdot((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}),

where we recall the definitions of the appropriate terms from (106), (108), (110), and (111).

Following arguments found in [63], we obtain the following estimates. We have, using (107), that

0tΩf(s)((𝒖)𝒖)(𝒖𝒖ˇδ)=0tΩf,δ(s)γδ[((𝒖ψδ))Jδ1(𝒖ψδ)](𝒖𝒖ˇδ)ψδ\displaystyle\int_{0}^{t}\int_{\Omega_{f}(s)}((\boldsymbol{u}\cdot\nabla)\boldsymbol{u})\cdot(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\gamma_{\delta}[(\nabla(\boldsymbol{u}\circ\psi_{\delta}))J_{\delta}^{-1}(\boldsymbol{u}\circ\psi_{\delta})]\cdot(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}
=0tΩf,δ(s)[((𝒖ψδ))𝒖^][γδ1Jδ(𝒖^𝒖δ)]\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[\gamma_{\delta}^{-1}J_{\delta}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})]
=0tΩf,δ(s)[((𝒖ψδ))𝒖^](𝒖^𝒖δ)0tΩf,δ(s)[((𝒖ψδ))𝒖^][(Iγδ1Jδ)(𝒖^𝒖δ)]\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[(I-\gamma_{\delta}^{-1}J_{\delta})(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})]
=0tΩf,δ(s)((𝒖^)𝒖^)(𝒖^𝒖δ)+0tΩf,δ(s)(((IγδJδ1)(𝒖ψδ))𝒖^)(𝒖^𝒖δ)\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}((\nabla\widehat{\boldsymbol{u}})\widehat{\boldsymbol{u}})\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})+\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla((I-\gamma_{\delta}J_{\delta}^{-1})(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}})\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})
0tΩf,δ(s)[((𝒖ψδ))𝒖^][(Iγδ1Jδ)(𝒖^𝒖δ)]\displaystyle-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[(I-\gamma_{\delta}^{-1}J_{\delta})(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})]
=0tΩf,δ(s)[(𝒖^)𝒖^](𝒖^𝒖δ)+R2,1,\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\widehat{\boldsymbol{u}}\cdot\nabla)\widehat{\boldsymbol{u}}]\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})+{{R_{2,1}}}, (161)

where

R2,1=0tΩf,δ(s)(((IγδJδ1)(𝒖ψδ))𝒖^)(𝒖^𝒖δ)0tΩf,δ(s)[((𝒖ψδ))𝒖^][(Iγδ1Jδ)(𝒖^𝒖δ)].\displaystyle{{R_{2,1}}}=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla((I-\gamma_{\delta}J_{\delta}^{-1})(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}})\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[(I-\gamma_{\delta}^{-1}J_{\delta})(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})].

In the following estimates, we will repeatedly use the following inequalities, which hold for a constant CC that is independent of δ\delta:

|γδ1JδI|C(|γδ11|+|γδ|)C||ωωδ||H2(Γ),|\gamma_{\delta}^{-1}J_{\delta}-I|\leq C(|\gamma_{\delta}^{-1}-1|+|\nabla\gamma_{\delta}|)\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)},
|γδJδ1I|C(|γδ1|+|γδ|)C||ωωδ||H2(Γ),|\gamma_{\delta}J_{\delta}^{-1}-I|\leq C(|\gamma_{\delta}-1|+|\nabla\gamma_{\delta}|)\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)},
|(γδJδ1)|C(|xγδ|+|xxγδ|)C(||ωωδ||H2(Γ)+|xx(ωωδ)|),|\nabla(\gamma_{\delta}J_{\delta}^{-1})|\leq C(|\partial_{x}\gamma_{\delta}|+|\partial_{xx}\gamma_{\delta}|)\leq C(||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}+|\partial_{xx}(\omega-\omega_{\delta})|), (162)

so that

||(γδJδ1)||L2(Ωf,δ(t))C||ωωδ||H2(Γ).||\nabla(\gamma_{\delta}J_{\delta}^{-1})||_{L^{2}(\Omega_{f,\delta}(t))}\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}. (163)

To obtain (162), we estimate |xxγδ||\partial_{xx}\gamma_{\delta}| by using the fact that ω\omega is smooth so that |xxω|C|\partial_{xx}\omega|\leq C and a direct computation of xxγδ\partial_{xx}\gamma_{\delta}. Using these estimates, the Leibniz rule, and the smoothness of 𝒖\boldsymbol{u}, we get

|0tΩf,δ(s)(((IγδJδ1)(𝒖ψδ))𝒖^)(𝒖^𝒖δ)|\displaystyle\left|\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla((I-\gamma_{\delta}J_{\delta}^{-1})(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}})\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\right|
C0t||ωωδ||H2(Γ)||𝒖^𝒖δ||L2(Ωf,δ(s))C(0t||ωωδ||2H2(Γ)+0t||𝒖^𝒖δ||2L2(Ωf,δ(s))).\displaystyle\leq C\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}\leq C\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{L^{2}(\Omega_{f,\delta}(s))}\right).

By using (107), and the fact that |Jδ|C|J_{\delta}|\leq C is uniformly bounded, due to the fact that |Jδ|C(1+||ωωδ||H2(Γ))C|J_{\delta}|\leq C(1+||\omega-\omega_{\delta}||_{H^{2}(\Gamma)})\leq C is uniformly bounded, we obtain a similar estimate:

0tΩf,δ(s)[((𝒖ψδ))𝒖^][(Iγδ1Jδ)(𝒖^𝒖δ)]C(0t||ωωδ||2H2(Γ)+0t||𝒖^𝒖δ||2L2(Ωf,δ(s))).\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla(\boldsymbol{u}\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[(I-\gamma_{\delta}^{-1}J_{\delta})(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})]\leq C\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{L^{2}(\Omega_{f,\delta}(s))}\right).

Thus, we obtain

|R1|C(0t||ωωδ||2H2(Γ)+0t||𝒖^𝒖δ||2L2(Ωf,δ(s))).{{|R_{1}|}}\leq C\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{L^{2}(\Omega_{f,\delta}(s))}\right). (164)

We now focus on the second integral in (160). By using (107) we obtain

0tΩf(s)((𝒖)(𝒖𝒖ˇδ))𝒖=0tΩf,δ(s)γδ[(((𝒖𝒖ˇδ)ψδ))Jδ1(𝒖ψδ)](𝒖ψδ)\displaystyle\int_{0}^{t}\int_{\Omega_{f}(s)}((\boldsymbol{u}\cdot\nabla)(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta}))\cdot\boldsymbol{u}=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\gamma_{\delta}[(\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}))J_{\delta}^{-1}(\boldsymbol{u}\circ\psi_{\delta})]\cdot(\boldsymbol{u}\circ\psi_{\delta})
=0tΩf,δ(s)[(((𝒖𝒖ˇδ)ψδ))𝒖^](γδ1Jδ𝒖^)\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot(\gamma_{\delta}^{-1}J_{\delta}\widehat{\boldsymbol{u}})
=0tΩf,δ(s)[(((𝒖𝒖ˇδ)ψδ))𝒖^]𝒖^0tΩf,δ(s)[(((𝒖𝒖ˇδ)ψδ))𝒖^][(Iγδ1Jδ)𝒖^]\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot\widehat{\boldsymbol{u}}-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[(I-\gamma_{\delta}^{-1}J_{\delta})\widehat{\boldsymbol{u}}]
=0tΩf,δ(s)((𝒖^𝒖δ)𝒖^)𝒖^+0tΩf,δ(s)([(IγδJδ1)((𝒖𝒖ˇδ)ψδ)]𝒖^)𝒖^\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\widehat{\boldsymbol{u}})\cdot\widehat{\boldsymbol{u}}+\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla[(I-\gamma_{\delta}J_{\delta}^{-1})((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta})]\widehat{\boldsymbol{u}})\cdot\widehat{\boldsymbol{u}}
0tΩf,δ(s)[(((𝒖𝒖ˇδ)ψδ))𝒖^][(Iγδ1Jδ)𝒖^]\displaystyle-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[(I-\gamma_{\delta}^{-1}J_{\delta})\widehat{\boldsymbol{u}}]
=0tΩf,δ(s)((𝒖^)(𝒖^𝒖δ))𝒖^+R2,2,\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}((\widehat{\boldsymbol{u}}\cdot\nabla)(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}))\cdot\widehat{\boldsymbol{u}}+{{R_{2,2}}}, (165)

where

R2,2:=0tΩf,δ(s)([(IγδJδ1)((𝒖𝒖ˇδ)ψδ)]𝒖^)𝒖^0tΩf,δ(s)[(((𝒖𝒖ˇδ)ψδ))𝒖^][(Iγδ1Jδ)𝒖^].R_{2,2}:=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla[(I-\gamma_{\delta}J_{\delta}^{-1})((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta})]\widehat{\boldsymbol{u}})\cdot\widehat{\boldsymbol{u}}-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}))\widehat{\boldsymbol{u}}]\cdot[(I-\gamma_{\delta}^{-1}J_{\delta})\widehat{\boldsymbol{u}}].

To estimate R2,2R_{2,2}, we will use the following inequalities:

|(𝒖𝒖ˇδ)ψ|\displaystyle|(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi| =|γδ1Jδ(𝒖^𝒖δ)|C|𝒖^𝒖δ|,\displaystyle=|\gamma_{\delta}^{-1}J_{\delta}\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})|\leq C|\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}|,
|((𝒖𝒖ˇδ)ψδ)|\displaystyle|\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta})| =|(γδ1Jδ(𝒖^𝒖δ))||(γδ1Jδ)||𝒖^𝒖δ|+|γδ1Jδ||(𝒖^𝒖δ)|\displaystyle=|\nabla(\gamma_{\delta}^{-1}J_{\delta}\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}))|\leq|\nabla(\gamma_{\delta}^{-1}J_{\delta})|\cdot|\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}|+|\gamma_{\delta}^{-1}J_{\delta}|\cdot|\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})|
C(|(γδ1Jδ)||𝒖^𝒖δ|+|(𝒖^𝒖δ)|).\displaystyle\leq C(|\nabla(\gamma_{\delta}^{-1}J_{\delta})|\cdot|\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}|+|\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})|).

From the fact that max(|Iγδ1Jδ|,|IγδJδ1|)Cmin(1,||ωωδ||H2(Γ))\max\left(|I-\gamma_{\delta}^{-1}J_{\delta}|,|I-\gamma_{\delta}J_{\delta}^{-1}|\right)\leq C\min\left(1,||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}\right), we obtain:

|R2,2|\displaystyle{{|R_{2,2}|}} C(0tΩf,δ(s)|(γδJδ1)||(𝒖𝒖ˇδ)ψδ|+0tΩf,δ(s)|IγδJδ1||((𝒖𝒖ˇδ)ψδ)|\displaystyle\leq C\left(\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}|\nabla(\gamma_{\delta}J_{\delta}^{-1})|\cdot|(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}|+\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}|I-\gamma_{\delta}J_{\delta}^{-1}|\cdot|\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta})|\right.
+0tΩf,δ(s)|Iγδ1Jδ||((𝒖𝒖ˇδ)ψδ)|)\displaystyle\left.+\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}|I-\gamma_{\delta}^{-1}J_{\delta}|\cdot|\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta})|\right)
C(0tΩf,δ(s)(|(γδJδ1)|+|(γδ1Jδ)|)|𝒖^𝒖δ|+0tΩf,δ(s)||ωωδ||H2(Γ)|(𝒖^𝒖δ)|)\displaystyle\leq C\left(\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\left(|\nabla(\gamma_{\delta}J_{\delta}^{-1})|+|\nabla(\gamma_{\delta}^{-1}J_{\delta})|\right)\cdot|\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}|+\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}\cdot|\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})|\right)
ϵ0t||(𝒖^𝒖δ)||L2(Ωf,δ(s))2+C(ϵ)(0t||ωωδ||2H2(Γ)+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).\displaystyle\leq\epsilon\int_{0}^{t}||\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right). (166)

In the last line, we use the following estimates, derived similarly as for (163),

|(γδ1Jδ)|C(|x(γδ1)|+|xγδ|+|xxγδ|)\displaystyle|\nabla(\gamma_{\delta}^{-1}J_{\delta})|\leq C(|\partial_{x}(\gamma_{\delta}^{-1})|+|\partial_{x}\gamma_{\delta}|+|\partial_{xx}\gamma_{\delta}|) C(||ωωδ||H2(Γ)+|xx(ωωδ)|),\displaystyle\leq C(||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}+|\partial_{xx}(\omega-\omega_{\delta})|),
||(γδ1Jδ)||L2(Ωf,δ(t))\displaystyle||\nabla(\gamma_{\delta}^{-1}J_{\delta})||_{L^{2}(\Omega_{f,\delta}(t))} C||ωωδ||H2(Γ).\displaystyle\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}.

Therefore, for the expression in (127), after transferring the integrals (A.2) and (A.2) and estimating R2,1R_{2,1} (164) and R2,2R_{2,2} (A.2), the remaining terms are:

120tΩf,δ(s)[(𝒖^)𝒖^](𝒖^𝒖δ)[(𝒖^)(𝒖^𝒖δ)]𝒖^\displaystyle\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\widehat{\boldsymbol{u}}\cdot\nabla)\widehat{\boldsymbol{u}}]\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})-[(\widehat{\boldsymbol{u}}\cdot\nabla)(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})]\cdot\widehat{\boldsymbol{u}}
120tΩf,δ(s))[(𝒖δ)𝒖δ](𝒖^𝒖δ)[(𝒖δ)(𝒖^𝒖δ)]𝒖δ\displaystyle-\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s))}[(\boldsymbol{u}_{\delta}\cdot\nabla)\boldsymbol{u}_{\delta}]\cdot(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})-[(\boldsymbol{u}_{\delta}\cdot\nabla)(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})]\cdot\boldsymbol{u}_{\delta}
=120tΩf,δ(s)[((𝒖^𝒖δ))𝒖δ]𝒖^120tΩf,δ(s)[((𝒖^𝒖δ))𝒖^]𝒖δ.\displaystyle=\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[((\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\nabla)\boldsymbol{u}_{\delta}]\cdot\widehat{\boldsymbol{u}}-\frac{1}{2}\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[((\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\nabla)\widehat{\boldsymbol{u}}]\cdot\boldsymbol{u}_{\delta}.

In absolute values, the right hand-side can be bounded as follows:

ϵ0t||(𝒖^𝒖δ)||L2(Ωf,δ(s))2+C(ϵ)0t||𝒖^𝒖δ||L2(Ωf,δ(s))2.\leq\epsilon\int_{0}^{t}||\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}.

Combining this estimate with (164) and (A.2) we obtain

|T2|ϵ0t||(𝒖^𝒖δ)||L2(Ωf,δ(s))2+C(ϵ)(0t||ωωδ||2H2(Γ)+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).|T_{2}|\leq\epsilon\int_{0}^{t}||\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Term T3. To estimate term T3T_{3} defined in (128), we start by noting that because 𝒖\boldsymbol{u} and 𝝃\boldsymbol{\xi} are smooth, we can pass to the limit as ν0\nu\to 0 using Proposition 10.1 and the fact that (𝝃δ)ν𝝃δ(\boldsymbol{\xi}_{\delta})_{\nu}\to\boldsymbol{\xi}_{\delta} strongly in L2(0,T;H1(Ωb))L^{2}(0,T;H^{1}(\Omega_{b})), so that we can ultimately just test with 𝒗=𝒖𝒖ˇδ\boldsymbol{v}=\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta} and 𝝍=𝝃𝝃δ\boldsymbol{\psi}=\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}. In the regularized weak formulation for 𝒖δ\boldsymbol{u}_{\delta}, we test with 𝒖\boldsymbol{u} and 𝝃\boldsymbol{\xi}. Note that both test functions 𝒖𝒖ˇδ\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta} and 𝒖^𝒖δ\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta} have the same trace along Γ(t)\Gamma(t) and Γδ(t)\Gamma_{\delta}(t) respectively, which we will formally denote by 𝒖𝒖δ\boldsymbol{u}-\boldsymbol{u}_{\delta} along the reference configuration of the interface Γ\Gamma. Combining the resulting expressions, we have the following contribution of T3T_{3} in the limit as ν0\nu\to 0:

T3\displaystyle T_{3} =120tΓ(s)(𝒖𝒏𝝃𝒏)𝒖(𝒖𝒖ˇδ)120tΓδ(s)(𝒖δ𝒏δ𝝃δ𝒏δ)𝒖δ𝒖^\displaystyle=\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{u}\cdot\boldsymbol{n}-\boldsymbol{\xi}\cdot\boldsymbol{n})\boldsymbol{u}\cdot(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{u}_{\delta}\cdot\boldsymbol{n}_{\delta}-\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta})\boldsymbol{u}_{\delta}\cdot\widehat{\boldsymbol{u}}
+120tΓ(s)|𝒖|2(𝝃𝒏𝒖𝒏)120tΓ(s)|𝒖|2(𝝃δ𝒏𝒖ˇδ𝒏)\displaystyle+\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}|\boldsymbol{u}|^{2}(\boldsymbol{\xi}\cdot\boldsymbol{n}-\boldsymbol{u}\cdot\boldsymbol{n})-\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}|\boldsymbol{u}|^{2}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}-\widecheck{\boldsymbol{u}}_{\delta}\cdot\boldsymbol{n})
120tΓδ(s)|𝒖δ|2(𝝃𝒏δ𝒖^𝒏δ)=120tΓ(s)(𝝃𝒏𝒖𝒏)𝒖𝒖ˇδ\displaystyle-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}|\boldsymbol{u}_{\delta}|^{2}(\boldsymbol{\xi}\cdot\boldsymbol{n}_{\delta}-\widehat{\boldsymbol{u}}\cdot\boldsymbol{n}_{\delta})=\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n}-\boldsymbol{u}\cdot\boldsymbol{n})\boldsymbol{u}\cdot\widecheck{\boldsymbol{u}}_{\delta}
120tΓ(s)(𝝃δ𝒏𝒖ˇδ𝒏)|𝒖|2120tΓδ(s)(𝝃𝒏δ𝒖^𝒏δ)|𝒖δ|2\displaystyle-\frac{1}{2}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}-\widecheck{\boldsymbol{u}}_{\delta}\cdot\boldsymbol{n})|\boldsymbol{u}|^{2}-\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n}_{\delta}-\widehat{\boldsymbol{u}}\cdot\boldsymbol{n}_{\delta})|\boldsymbol{u}_{\delta}|^{2}
+120tΓδ(s)(𝝃δ𝒏δ𝒖δ𝒏δ)𝒖δ𝒖^=R3,1+R3,2,\displaystyle+\frac{1}{2}\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot\boldsymbol{n}_{\delta}-\boldsymbol{u}_{\delta}\cdot\boldsymbol{n}_{\delta})\boldsymbol{u}_{\delta}\cdot\widehat{\boldsymbol{u}}={{R_{3,1}+R_{3,2}}},

where

R3,1=120tΓ[(𝝃𝒖)𝒆y]𝒖δ(𝒖𝒖δ)120tΓ[(𝝃δ𝒖δ)𝒆y]𝒖(𝒖𝒖δ),{{R_{3,1}}}=\frac{1}{2}\int_{0}^{t}\int_{\Gamma}[(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{e}_{y}]\boldsymbol{u}_{\delta}\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta})-\frac{1}{2}\int_{0}^{t}\int_{\Gamma}[(\boldsymbol{\xi}_{\delta}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{e}_{y}]\boldsymbol{u}\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta}),
R3,2=120tΓxω(𝒖𝒆x)𝒖𝒖δ120tΓxω(𝒖δ𝒆x)|𝒖|2120tΓxωδ(𝒖𝒆x)|𝒖δ|2+120tΓxωδ(𝒖δ𝒆x)𝒖𝒖δ.{{R_{3,2}}}=\frac{1}{2}\int_{0}^{t}\int_{\Gamma}\partial_{x}\omega(\boldsymbol{u}\cdot\boldsymbol{e}_{x})\boldsymbol{u}\cdot\boldsymbol{u}_{\delta}-\frac{1}{2}\int_{0}^{t}\int_{\Gamma}\partial_{x}\omega(\boldsymbol{u}_{\delta}\cdot\boldsymbol{e}_{x})|\boldsymbol{u}|^{2}-\frac{1}{2}\int_{0}^{t}\int_{\Gamma}\partial_{x}\omega_{\delta}(\boldsymbol{u}\cdot\boldsymbol{e}_{x})|\boldsymbol{u}_{\delta}|^{2}\\ +\frac{1}{2}\int_{0}^{t}\int_{\Gamma}\partial_{x}\omega_{\delta}(\boldsymbol{u}_{\delta}\cdot\boldsymbol{e}_{x})\boldsymbol{u}\cdot\boldsymbol{u}_{\delta}.

We estimate R3,1R_{3,1} as follows: decompose R3,1R_{3,1} as R3,1=R3,1,1+R3,1,2R_{3,1}=R_{3,1,1}+R_{3,1,2}, where

R3,1,1=120tΓ(𝝃𝒆y)(𝒖𝒖δ)(𝒖𝒖δ)+120tΓ[(𝝃𝝃δ)𝒆y]𝒖(𝒖𝒖δ),{{R_{3,1,1}}}=-\frac{1}{2}\int_{0}^{t}\int_{\Gamma}(\boldsymbol{\xi}\cdot\boldsymbol{e}_{y})(\boldsymbol{u}-\boldsymbol{u}_{\delta})\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta})+\frac{1}{2}\int_{0}^{t}\int_{\Gamma}[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{e}_{y}]\boldsymbol{u}\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta}),
R3,1,2=120tΓ(𝒖𝒆y)(𝒖𝒖δ)(𝒖𝒖δ)120tΓ[(𝒖𝒖δ)𝒆y]𝒖(𝒖𝒖δ).{{R_{3,1,2}}}=\frac{1}{2}\int_{0}^{t}\int_{\Gamma}(\boldsymbol{u}\cdot\boldsymbol{e}_{y})(\boldsymbol{u}-\boldsymbol{u}_{\delta})\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta})-\frac{1}{2}\int_{0}^{t}\int_{\Gamma}[(\boldsymbol{u}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{e}_{y}]\boldsymbol{u}\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta}).

By interpolation,

|R3,1,1|C(0t||𝒖^𝒖δ||1/2L2(Ωf,δ(s))||𝒖^𝒖δ||3/2H1(Ωf,δ(s))+0t||𝝃𝝃δ||L2(Γ)||𝒖^𝒖δ||H1(Ωf,δ(s)))ϵ0t||𝒖^𝒖δ||2H1(Ωf,δ(s))+C(ϵ)(0t||𝒖^𝒖δ||L2(Ωf,δ(s))2+0t||𝝃𝝃δ||L2(Γ)2).{{|R_{3,1,1}|}}\leq C\left(\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{1/2}_{L^{2}(\Omega_{f,\delta}(s))}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{3/2}_{H^{1}(\Omega_{f,\delta}(s))}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||_{L^{2}(\Gamma)}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{H^{1}(\Omega_{f,\delta}(s))}\right)\\ \leq\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{H^{1}(\Omega_{f,\delta}(s))}+C(\epsilon)\left(\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||_{L^{2}(\Gamma)}^{2}\right).

By using the same interpolation inequality, we obtain the following estimate for R3,1,2R_{3,1,2}.

|R3,1,2|ϵ0t||𝒖^𝒖δ||2H1(Ωf,δ(s))+C(ϵ)0t||𝒖^𝒖δ||L2(Ωf,δ(s))2.{{|R_{3,1,2}|}}\leq\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||^{2}_{H^{1}(\Omega_{f,\delta}(s))}+C(\epsilon)\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}.

We estimate R3,2R_{3,2} by first rewriting R3,2R_{3,2} as follows:

R3,2\displaystyle{{R_{3,2}}} =120tΓ(xωxωδ)(𝒖)x𝒖(𝒖𝒖δ)120tΓxωδ(𝒖)x(𝒖𝒖δ)(𝒖𝒖δ)\displaystyle=-\frac{1}{2}\int_{0}^{t}\int_{\Gamma}(\partial_{x}\omega-\partial_{x}\omega_{\delta})(\boldsymbol{u})_{x}\boldsymbol{u}\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta})-\frac{1}{2}\int_{0}^{t}\int_{\Gamma}\partial_{x}\omega_{\delta}(\boldsymbol{u})_{x}(\boldsymbol{u}-\boldsymbol{u}_{\delta})\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta})
+120tΓ(xωxωδ)(𝒖𝒖δ)x|𝒖|2+120tΓxωδ(𝒖𝒖δ)x𝒖(𝒖𝒖δ).\displaystyle+\frac{1}{2}\int_{0}^{t}\int_{\Gamma}(\partial_{x}\omega-\partial_{x}\omega_{\delta})(\boldsymbol{u}-\boldsymbol{u}_{\delta})_{x}|\boldsymbol{u}|^{2}+\frac{1}{2}\int_{0}^{t}\int_{\Gamma}\partial_{x}\omega_{\delta}(\boldsymbol{u}-\boldsymbol{u}_{\delta})_{x}\boldsymbol{u}\cdot(\boldsymbol{u}-\boldsymbol{u}_{\delta}).

By interpolation, by the boundedness of |xω||\partial_{x}\omega| and |xωδ||\partial_{x}\omega_{\delta}|, and by the smoothness of 𝒖\boldsymbol{u}, we get:

|R3,2|ϵ0t||𝒖^𝒖δ||H1(Ωf,δ(s))2+C(ϵ)(0t||ωωδ||2H2(Γ)+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).{{|R_{3,2}|}}\leq\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{H^{1}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Hence, by combining the two estimates we get the final estimate for T3T_{3}:

|T3|ϵ0t||𝒖^𝒖δ||H1(Ωf,δ(s))2+C(ϵ)(0t||ωωδ||2H2(Γ)+0t||𝝃𝝃δ||2L2(Γ)+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).\displaystyle|T_{3}|\leq\epsilon\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{H^{1}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Term T4. To estimate term T4T_{4}, defined in (129), we again use Proposition 10.1 to pass to the limit as ν0\nu\to 0 so that the contribution from T4T_{4} is

T4:=2ν0tΩf(s)𝑫(𝒖):𝑫(𝒖𝒖ˇδ)2ν0tΩf,δ(s)𝑫(𝒖δ):𝑫(𝒖^𝒖δ).T_{4}:=2\nu\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})-2\nu\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\boldsymbol{D}(\boldsymbol{u}_{\delta}):\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}). (167)

We want to transfer the integral on Ω1(t)\Omega_{1}(t) to Ωf,δ(t)\Omega_{f,\delta}(t). Recalling (107), we have that

0tΩf(s)𝑫(𝒖):𝑫(𝒖𝒖ˇδ)=0tΩf,δ(s)γδ[(𝒖ψδ)Jδ1]sym:[((𝒖𝒖ˇδ)ψδ)Jδ1]sym,\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\gamma_{\delta}[\nabla(\boldsymbol{u}\circ\psi_{\delta})J_{\delta}^{-1}]^{sym}:[\nabla((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta})J_{\delta}^{-1}]^{sym},

where the superscript ‘sym’ notation denotes a symmetrization. Following the procedure in [63], we break up the integral as

0tΩf(s)𝑫(𝒖):𝑫(𝒖𝒖ˇδ)=0tΩf,δ(s)𝑫(𝒖^):𝑫(𝒖^𝒖δ)+R4,1+R4,2+R4,3+R4,4,\int_{0}^{t}\int_{\Omega_{f}(s)}\boldsymbol{D}(\boldsymbol{u}):\boldsymbol{D}(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}\boldsymbol{D}(\widehat{\boldsymbol{u}}):\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})+{{R_{4,1}+R_{4,2}+R_{4,3}+R_{4,4}}}, (168)

where

R4,1\displaystyle{{R_{4,1}}} =0tΩf,δ(s)((𝒖ψδ)Jδ1)sym:[(𝒖^𝒖δ)(Jδ1I)+(JδI)(𝒖^𝒖δ)Jδ1]sym,\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla(\boldsymbol{u}\circ\psi_{\delta})J_{\delta}^{-1})^{sym}:[\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})(J_{\delta}^{-1}-I)+(J_{\delta}-I)\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})J_{\delta}^{-1}]^{sym},
R4,2\displaystyle{{R_{4,2}}} =0tΩf,δ(s)[(IγδJδ1)(𝒖ψδ)+(𝒖ψδ)(Jδ1I)]sym:𝑫(𝒖^𝒖δ),\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(I-\gamma_{\delta}J_{\delta}^{-1})\nabla(\boldsymbol{u}\circ\psi_{\delta})+\nabla(\boldsymbol{u}\circ\psi_{\delta})(J_{\delta}^{-1}-I)]^{sym}:\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}),
R4,3\displaystyle{{R_{4,3}}} =0tΩf,δ(s)((𝒖ψδ)Jδ1)sym:(γδ(γδ1Jδ)(𝒖^𝒖δ)Jδ1)sym,\displaystyle=\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}(\nabla(\boldsymbol{u}\circ\psi_{\delta})J_{\delta}^{-1})^{sym}:(\gamma_{\delta}\nabla(\gamma_{\delta}^{-1}J_{\delta})(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})J_{\delta}^{-1})^{sym},
R4,4\displaystyle{{R_{4,4}}} =0tΩf,δ(s)[((γδJδ1))𝒖ψδ]sym:𝑫(𝒖^𝒖δ).\displaystyle=-\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}[(\nabla(\gamma_{\delta}J_{\delta}^{-1}))\boldsymbol{u}\circ\psi_{\delta}]^{sym}:\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}).

To verify this equality, one can use the Leibniz rule, the definition 𝒖^=γδJδ1(𝒖ψδ)\widehat{\boldsymbol{u}}=\gamma_{\delta}J_{\delta}^{-1}\cdot(\boldsymbol{u}\circ\psi_{\delta}), and the identity 𝒖^𝒖δ=γδJδ1((𝒖𝒖ˇδ)ψδ)\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}=\gamma_{\delta}J_{\delta}^{-1}\cdot((\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\circ\psi_{\delta}).

We now estimate the terms R4,1R_{4,1}-R4,4R_{4,4}. For this purpose we will use the following inequalities:

|Jδ1|\displaystyle|J_{\delta}^{-1}| C(1+|xγδ|),|Jδ1I|C(|γδ11|+|xγδ|),\displaystyle\leq C(1+|\partial_{x}\gamma_{\delta}|),\quad|J_{\delta}^{-1}-I|\leq C(|\gamma_{\delta}^{-1}-1|+|\partial_{x}\gamma_{\delta}|),
|JδI|\displaystyle|J_{\delta}-I| C(|γδ1|+|xγδ|),|γδJδ1I|C(|γδ1|+|xγδ|).\displaystyle\leq C(|\gamma_{\delta}-1|+|\partial_{x}\gamma_{\delta}|),\quad|\gamma_{\delta}J_{\delta}^{-1}-I|\leq C(|\gamma_{\delta}-1|+|\partial_{x}\gamma_{\delta}|).

and, recalling the definition of γδ\gamma_{\delta} in (106), we have the following inequalities:

|γδ1|C||ωωδ||H2(Γ),|γδ11|C||ωωδ||H2(Γ),|\gamma_{\delta}-1|\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)},\qquad|\gamma_{\delta}^{-1}-1|\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)},
|xγδ|C||ωωδ||H2(Γ),|x(γδ1)|C||ωωδ||H2(Γ).|\partial_{x}\gamma_{\delta}|\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)},\qquad|\partial_{x}(\gamma_{\delta}^{-1})|\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}.

Because |Jδ1|C(1+|xγδ|)C|J_{\delta}^{-1}|\leq C(1+|\partial_{x}\gamma_{\delta}|)\leq C since |xγδ||\partial_{x}\gamma_{\delta}| is bounded, and because 𝒖\boldsymbol{u} is smooth,

|R4,1|\displaystyle{{|R_{4,1}|}} C0t||(𝒖^𝒖δ)||L2(Ωf,δ(s))(||γδ11||L2(Ωf,δ(s))+||γδ1||L2(Ωf,δ(s))+||xγδ||L2(Ωf,δ(s)))\displaystyle\leq C\int_{0}^{t}||\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}(||\gamma_{\delta}^{-1}-1||_{L^{2}(\Omega_{f,\delta}(s))}+||\gamma_{\delta}-1||_{L^{2}(\Omega_{f,\delta}(s))}+||\partial_{x}\gamma_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))})
ϵ0t||(𝒖^𝒖δ)||L2(Ωf,δ(s))2+C(ϵ)0t||ωωδ||2H2(Γ).\displaystyle\leq\epsilon\int_{0}^{t}||\nabla(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}.

We also have that

|R4,2|ϵ0t||𝑫(𝒖^𝒖δ)||L2(Ωf,δ(s))2+C(ϵ)0t||ωωδ||2H2(Γ).{{|R_{4,2}|}}\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}.

For R4,3R_{4,3} and R4,4R_{4,4}, we compute that

(γδ1Jδ)=(γδ10(R+y)γδ1xγδ1),(γδJδ1)=(γδ0(R+y)xγδ1).\nabla(\gamma_{\delta}^{-1}J_{\delta})=\nabla\begin{pmatrix}\gamma_{\delta}^{-1}&0\\ (R+y)\gamma_{\delta}^{-1}\partial_{x}\gamma_{\delta}&1\\ \end{pmatrix},\qquad\nabla(\gamma_{\delta}J_{\delta}^{-1})=\nabla\begin{pmatrix}\gamma_{\delta}&0\\ -(R+y)\partial_{x}\gamma_{\delta}&1\\ \end{pmatrix}.

Therefore,

|(γδ1Jδ)|C(|x(γδ1)|+|xγδ|+|xxγδ|),|(γδJδ1)|C(|xγδ|+|xxγδ|),|\nabla(\gamma_{\delta}^{-1}J_{\delta})|\leq C(|\partial_{x}(\gamma_{\delta}^{-1})|+|\partial_{x}\gamma_{\delta}|+|\partial_{xx}\gamma_{\delta}|),\qquad|\nabla(\gamma_{\delta}J_{\delta}^{-1})|\leq C(|\partial_{x}\gamma_{\delta}|+|\partial_{xx}\gamma_{\delta}|),

where we can estimate

|xxγδ|C(||ωωδ||H2(Γ)|xxω|+|xx(ωωδ)|+||ωωδ||H2(Γ)).|\partial_{xx}\gamma_{\delta}|\leq C(||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}|\partial_{xx}\omega|+|\partial_{xx}(\omega-\omega_{\delta})|+||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}).

So since ||xxω||L2(Ωf,δ(t))C||\partial_{xx}\omega||_{L^{2}(\Omega_{f,\delta}(t))}\leq C since ω\omega is uniformly bounded in H2(Γ)H^{2}(\Gamma), we have that

|R4,3|\displaystyle{{|R_{4,3}|}} C0t||(γδ1Jδ)||L2(Ωf,δ(s))||𝒖^𝒖δ||L2(Ωf,δ(s))C0t||ωωδ||H2(Γ)||𝒖^𝒖δ||L2(Ωf,δ(s))\displaystyle\leq C\int_{0}^{t}||\nabla(\gamma_{\delta}^{-1}J_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}\leq C\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}
C(0t||ωωδ||H2(Γ)2+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).\displaystyle\leq C\left(\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Similarly, using ||(γδJδ1)||L2(Ωf,δ(t))C||ωωδ||H2(Γ)||\nabla(\gamma_{\delta}J_{\delta}^{-1})||_{L^{2}(\Omega_{f,\delta}(t))}\leq C||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}, we have the following estimate for R4R_{4}:

|R4,4|\displaystyle{{|R_{4,4}|}} C0t||(γδJδ1)||L2(Ωf,δ(s))||𝑫(𝒖^𝒖δ)||L2(Ωf,δ(s))\displaystyle\leq C\int_{0}^{t}||\nabla(\gamma_{\delta}J_{\delta}^{-1})||_{L^{2}(\Omega_{f,\delta}(s))}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}
ϵ0t||𝑫(𝒖^𝒖δ)||L2(Ωf,δ(s))2+C(ϵ)0t||ωωδ||2H2(Γ).\displaystyle\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}.

We now have the final estimate of T4T_{4}, obtained after using (167) and (168) as follows:

T4=2ν0tΩf,δ(s)|𝑫(𝒖^𝒖δ)|2+R4,T_{4}=2\nu\int_{0}^{t}\int_{\Omega_{f,\delta}(s)}|\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})|^{2}+{{R_{4}}},

where

|R4|ϵ0t||𝑫(𝒖^𝒖δ)||2L2(Ωf,δ(s))+C(ϵ)(0t||ωωδ||2H2(Γ)+0t||𝒖^𝒖δ||L2(Ωf,δ(s))2).{{|R_{4}|}}\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||^{2}_{L^{2}(\Omega_{f,\delta}(s))}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta}||_{L^{2}(\Omega_{f,\delta}(s))}^{2}\right).

Term T5. Similarly as before, after passing to the limit as ν0\nu\to 0 in term T5T_{5}, defined by (169), the contribution of this term is

T5=β0tΓ(s)(𝝃𝒖)𝝉(s)[(𝝃𝝃δ)𝝉(s)(𝒖𝒖ˇδ)𝝉(s)]β0tΓδ(s)(𝝃δ𝒖δ)𝝉δ(s)[(𝝃𝝃δ)𝝉δ(s)(𝒖^𝒖δ)𝝉δ(s)].T_{5}=\beta\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}(s)[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}(s)-(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\cdot\boldsymbol{\tau}(s)]\\ -\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}_{\delta}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)-(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)]. (169)

We note that when we test the weak formulation for 𝒖\boldsymbol{u} with 𝒗=𝒖(𝒖ˇδ)ν\boldsymbol{v}=\boldsymbol{u}-(\widecheck{\boldsymbol{u}}_{\delta})_{\nu} and 𝝍=𝝃(𝝃δ)ν\boldsymbol{\psi}=\boldsymbol{\xi}-(\boldsymbol{\xi}_{\delta})_{\nu}, we can pass to the limit as ν0\nu\to 0 to obtain the first term in T5T_{5} above, by using similar arguments involving Proposition 10.1, as for the previously considered terms. This term can now be rewritten as follows:

T5=β0tΓδ(s)|(𝝃𝝃δ)𝝉δ(s)(𝒖^𝒖δ)𝝉δ(s)|2+R5,T_{5}=\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}|(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)-(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)|^{2}+R_{5},

where

R5=β0tΓ(s)(𝝃𝒖)𝝉(s)[(𝝃𝝃δ)𝝉(s)(𝒖𝒖ˇδ)𝝉(s)]β0tΓδ(s)(𝝃𝒖^)𝝉δ(s)[(𝝃𝝃δ)𝝉δ(s)(𝒖^𝒖δ)𝝉δ(s)].R_{5}=\beta\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{\tau}(s)[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}(s)-(\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta})\cdot\boldsymbol{\tau}(s)]\\ -\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}(\boldsymbol{\xi}-\widehat{\boldsymbol{u}})\cdot\boldsymbol{\tau}_{\delta}(s)[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)-(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)].

Denote the arc length elements of Γ(t)\Gamma(t) and Γδ(t)\Gamma_{\delta}(t) respectively by 𝒥ωΓ=1+|xω|2\mathcal{J}^{\omega}_{\Gamma}=\sqrt{1+|\partial_{x}\omega|^{2}} and 𝒥ωδΓ=1+|xωδ|2\mathcal{J}^{\omega_{\delta}}_{\Gamma}=\sqrt{1+|\partial_{x}\omega_{\delta}|^{2}}, and we recall that we denote the tangent vectors to Γ(t)\Gamma(t) and Γδ(t)\Gamma_{\delta}(t) respectively by 𝝉(t)=1𝒥ωΓ(1,xω)\boldsymbol{\tau}(t)=\frac{1}{\mathcal{J}^{\omega}_{\Gamma}}(1,\partial_{x}\omega) and 𝝉δ(t)=1𝒥ωδΓ(1,xωδ)\boldsymbol{\tau}_{\delta}(t)=\frac{1}{\mathcal{J}^{\omega_{\delta}}_{\Gamma}}(1,\partial_{x}\omega_{\delta}). We can now rewrite R5R_{5} by writing everything in terms of the xx and yy components. For this purpose, recall that 𝝃\boldsymbol{\xi} and 𝝃δ\boldsymbol{\xi}_{\delta} along the interface displace in only the yy direction. We formally express the common trace of 𝒖𝒖ˇδ\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta} and 𝒖^𝒖δ\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta} along the reference configuration of the interface Γ\Gamma by 𝒖𝒖δ\boldsymbol{u}-\boldsymbol{u}_{\delta}. Thus,

R5\displaystyle R_{5} =β0tΓ(𝝃𝒖)(1,xω)[(𝝃𝝃δ)(𝒖𝒖δ)]𝝉(s)\displaystyle=\beta\int_{0}^{t}\int_{\Gamma}(\boldsymbol{\xi}-\boldsymbol{u})\cdot(1,\partial_{x}\omega)[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})-(\boldsymbol{u}-\boldsymbol{u}_{\delta})]\cdot\boldsymbol{\tau}(s)
β0tΓ(𝝃𝒖)(1,xωδ)[(𝝃𝝃δ)(𝒖𝒖δ)]𝝉δ(s).\displaystyle-\beta\int_{0}^{t}\int_{\Gamma}(\boldsymbol{\xi}-\boldsymbol{u})\cdot(1,\partial_{x}\omega_{\delta})[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})-(\boldsymbol{u}-\boldsymbol{u}_{\delta})]\cdot\boldsymbol{\tau}_{\delta}(s).

In the previous step, we used the fact that when transferred back to the reference configuration Ωf\Omega_{f}, 𝒖^𝒖δ\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta} and 𝒖𝒖ˇδ\boldsymbol{u}-\widecheck{\boldsymbol{u}}_{\delta} have the same trace along Γ\Gamma. Thus, R5=R5,1+R5,2R_{5}=R_{5,1}+R_{5,2}, where

R5,1\displaystyle R_{5,1} =β0tΓ(𝝃𝒖)𝒆y(xωxωδ)[(𝝃𝝃δ)(𝒖𝒖δ)]𝝉(s),\displaystyle=\beta\int_{0}^{t}\int_{\Gamma}(\boldsymbol{\xi}-\boldsymbol{u})\cdot\boldsymbol{e}_{y}(\partial_{x}\omega-\partial_{x}\omega_{\delta})[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})-(\boldsymbol{u}-\boldsymbol{u}_{\delta})]\cdot\boldsymbol{\tau}(s),
R5,2\displaystyle R_{5,2} =β0tΓ(𝝃𝒖)(1,xωδ)[(𝝃𝝃δ)(𝒖𝒖δ)](𝝉(s)𝝉δ(s)).\displaystyle=\beta\int_{0}^{t}\int_{\Gamma}(\boldsymbol{\xi}-\boldsymbol{u})\cdot(1,\partial_{x}\omega_{\delta})[(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})-(\boldsymbol{u}-\boldsymbol{u}_{\delta})]\cdot(\boldsymbol{\tau}(s)-\boldsymbol{\tau}_{\delta}(s)).

We can use the fact that |xω||\partial_{x}\omega| and |xωδ||\partial_{x}\omega_{\delta}| are uniformly bounded to obtain the following estimates:

|R5,1|ϵ0t||𝑫(𝒖^𝒖δ)||L2(Ωf,δ(s))+C(ϵ)(0t||ωωδ||H2(Γ)2+0t||𝝃𝝃δ||L2(Γ)2),|R_{5,1}|\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||_{L^{2}(\Gamma)}^{2}\right),

where we used the trace inequality, Poincare’s inequality, and Korn’s inequality for the fluid. For the second term R5,2R_{5,2}, we use the estimate |𝝉(s)𝝉δ(s)|C|xωxωδ||\boldsymbol{\tau}(s)-\boldsymbol{\tau}_{\delta}(s)|\leq C|\partial_{x}\omega-\partial_{x}\omega_{\delta}| to obtain

|R5,2|ϵ0t||𝑫(𝒖^𝒖δ)||L2(Ωf,δ(s))+C(ϵ)(0t||ωωδ||H2(Γ)2+0t||𝝃𝝃δ||L2(Γ)2).|R_{5,2}|\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||_{L^{2}(\Gamma)}^{2}\right).

Hence,

T5=β0tΓδ(s)|(𝝃𝝃δ)𝝉δ(s)(𝒖^𝒖δ)𝝉δ(s)|2+R5,T_{5}=\beta\int_{0}^{t}\int_{\Gamma_{\delta}(s)}|(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)-(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})\cdot\boldsymbol{\tau}_{\delta}(s)|^{2}+R_{5},

where

|R5|ϵ0t||𝑫(𝒖^𝒖δ)||L2(Ωf,δ(s))+C(ϵ)(0t||ωωδ||H2(Γ)2+0t||𝝃𝝃δ||L2(Γ)2).|R_{5}|\leq\epsilon\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||_{L^{2}(\Omega_{f,\delta}(s))}+C(\epsilon)\left(\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||_{L^{2}(\Gamma)}^{2}\right).

Terms T6-T8. We will present estimates only for term T6T_{6}, defined in (130), as the same procedure will hold for T7T_{7} and T8T_{8}. Since ζ\zeta and ζδ\zeta_{\delta} are weakly continuous in L2(Γ)L^{2}(\Gamma), by the weak formulation, we get:

0tΓζt[(ζδ)ν]+0tΓζδtζ\displaystyle\int_{0}^{t}\int_{\Gamma}\zeta\cdot\partial_{t}\left[(\zeta_{\delta})_{\nu}\right]+\int_{0}^{t}\int_{\Gamma}\zeta_{\delta}\cdot\partial_{t}\zeta =0tΓζt[(ζδ)ν]+0tΓζδt[(ζ)ν]0tΓζδt[(ζ)νζ]\displaystyle=\int_{0}^{t}\int_{\Gamma}\zeta\cdot\partial_{t}\left[(\zeta_{\delta})_{\nu}\right]+\int_{0}^{t}\int_{\Gamma}\zeta_{\delta}\cdot\partial_{t}\left[(\zeta)_{\nu}\right]-\int_{0}^{t}\int_{\Gamma}\zeta_{\delta}\cdot\partial_{t}\left[(\zeta)_{\nu}-\zeta\right]
Γζ(t)ζδ(t)Γ|ζ0|2.\displaystyle\to\int_{\Gamma}\zeta(t)\cdot\zeta_{\delta}(t)-\int_{\Gamma}|\zeta_{0}|^{2}.

This follows from Lemma 2.5 in [63], which implies:

0tΓζt[(ζδ)ν]+0tΓζδt[(ζ)ν]Γζ(t)ζδ(t)Γ|ζ0|2, as ν0,\int_{0}^{t}\int_{\Gamma}\zeta\cdot\partial_{t}\left[(\zeta_{\delta})_{\nu}\right]+\int_{0}^{t}\int_{\Gamma}\zeta_{\delta}\cdot\partial_{t}\left[(\zeta)_{\nu}\right]\to\int_{\Gamma}\zeta(t)\cdot\zeta_{\delta}(t)-\int_{\Gamma}|\zeta_{0}|^{2},\qquad\text{ as }\nu\to 0,

and from the fact that ζ\zeta is smooth in space and time, which implies

0tΓζδt[(ζ)νζ]0, as ν0.\int_{0}^{t}\int_{\Gamma}\zeta_{\delta}\cdot\partial_{t}\left[(\zeta)_{\nu}-\zeta\right]\to 0,\qquad\text{ as }\nu\to 0.

Furthermore, because ζ(0)=ζδ(0)=ζ0\zeta(0)=\zeta_{\delta}(0)=\zeta_{0} weak continuity of ζδ\zeta_{\delta} at t=0t=0 implies that Γζ(0)[ζ(0)(ζδ)ν(0)]0\displaystyle\int_{\Gamma}\zeta(0)\cdot[\zeta(0)-(\zeta_{\delta})_{\nu}(0)]\to 0 as ν0\nu\to 0. Similarly, Γζ(s)[ζ(s)(ζδ)ν(s)]0\displaystyle\int_{\Gamma}\zeta(s)\cdot[\zeta(s)-(\zeta_{\delta})_{\nu}(s)]\to 0 as ν0\nu\to 0 for almost every s[0,T]s\in[0,T]. Hence, as ν0\nu\to 0, the contribution from T6T_{6} is

T6=12ρpΓ|(ζζδ)(t)|2.T_{6}=\frac{1}{2}\rho_{p}\int_{\Gamma}|(\zeta-\zeta_{\delta})(t)|^{2}.

Similarly, the contributions from T7T_{7} and T8T_{8} as ν0\nu\to 0 are

T7=12Γ|Δ(ωωδ)(t)|2,T8=12ρbΩb|(𝝃𝝃δ)(t)|2.T_{7}=\frac{1}{2}\int_{\Gamma}|\Delta(\omega-\omega_{\delta})(t)|^{2},\qquad T_{8}=\frac{1}{2}\rho_{b}\int_{\Omega_{b}}|(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})(t)|^{2}.

Terms T9-T12. Since these calculations are straight forward, a discussion about the limiting expressions as ν0\nu\to 0 for terms T9T_{9}-T12T_{12} was presented earlier, just under (133).

Term T13. Similarly as before, by taking the limit as ν0\nu\to 0, we have that

T13=α0tΩb(s)p(𝝃𝝃δ)+α0tΩδb,δ(s)pδ(𝝃𝝃δ).T_{13}=-{\alpha}\int_{0}^{t}\int_{\Omega_{b}(s)}p\nabla\cdot(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})+{\alpha}\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}p_{\delta}\nabla\cdot(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}).

To estimate this term we use (21) and the matrix identity 𝑩1=1det(𝑩)𝑩C\boldsymbol{B}^{-1}=\frac{1}{\text{det}(\boldsymbol{B})}\boldsymbol{B}^{C} to obtain

|T13|=α|0tΩb𝒥ηbpηb(𝝃𝝃δ)0tΩb𝒥ηδδbpδηδδb(𝝃𝝃δ)|=α|0tΩbptr((𝝃𝝃δ)(𝑰+𝜼)C)0tΩbpδtr((𝝃𝝃δ)(𝑰+𝜼δδ)C)|R13,1+R13,2,|T_{13}|={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{\eta}_{b}p\nabla^{\eta}_{b}\cdot(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})-\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}p_{\delta}\nabla^{{\eta}^{\delta}_{\delta}}_{b}\cdot(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\right|\\ ={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}p\cdot\text{tr}\left(\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{C}\right)-\int_{0}^{t}\int_{\Omega_{b}}p_{\delta}\cdot\text{tr}\left(\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right)\right|\leq{{R_{13,1}+R_{13,2}}},

where the superscript ``C"``C" denotes the cofactor matrix. The integrals R13,1R_{13,1} and R13,2R_{13,2} are defined as follows:

R13,1\displaystyle{{R_{13,1}}} =α|0tΩbptr((𝝃𝝃δ)((𝜼𝜼δδ))C)|,\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}p\cdot\text{tr}\left(\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot(\nabla(\boldsymbol{\eta}-{\boldsymbol{\eta}}^{\delta}_{\delta}))^{C}\right)\right|,
R13,2\displaystyle{{R_{13,2}}} =α|0tΩb(ppδ)tr((𝝃𝝃δ)(𝑰+𝜼δδ)C)|.\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}(p-p_{\delta})\cdot\text{tr}\left(\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right)\right|.

In the previous calculations, we observe that the cofactor matrix operation is linear when the matrices are two by two. Using the fact that pp is smooth, the assumption (117), and the fact that

||𝜼δ𝜼δδ||L2(Ωb)C||𝜼𝜼δ||L2(Ω~b)C(||𝜼𝜼δ||L2(Ωb)+||ωωδ||H2(Γ))||\nabla{\boldsymbol{\eta}}^{\delta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}||_{L^{2}(\Omega_{b})}\leq{{C||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\tilde{\Omega}_{b})}}}\leq C\left(||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}+||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}\right) (170)

for a constant CC independent of δ\delta, by Young’s convolution inequality and the definition of odd extension to the larger domain Ω~b\tilde{\Omega}_{b} in Definition 5.1, we obtain the estimates on R13,1R_{13,1} and R13,2R_{13,2}:

R13,1\displaystyle{{R_{13,1}}} ϵ0t||(𝝃𝝃δ)||L2(Ωb)2+C(ϵ)(0t||𝜼𝜼δδ||L2(Ωb)2)\displaystyle\leq\epsilon\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||_{L^{2}(\Omega_{b})}^{2}+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}\right)
C(ϵ)0t||𝜼𝜼δ||L2(Ωb)2+ϵ0t||(𝝃𝝃δ)||L2(Ωb)2\displaystyle\leq C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}+\epsilon\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||_{L^{2}(\Omega_{b})}^{2}
+C(ϵ)(0t||𝜼𝜼δ||2L2(Ωb)+0t||ωωδ||2H2(Γ)),\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}\right),
R13,2\displaystyle{{R_{13,2}}} ϵ0t||(𝝃𝝃δ)||L2(Ωb)2+C(ϵ)(0t||ppδ||2L2(Ωb)).\displaystyle\leq\epsilon\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||_{L^{2}(\Omega_{b})}^{2}+C(\epsilon)\left(\int_{0}^{t}||p-p_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right).

Therefore, the final estimate for T13T_{13} is as follows:

|T13|\displaystyle|T_{13}| C(ϵ)0t||𝜼𝜼δ||L2(Ωb)2+ϵ0t||(𝝃𝝃δ)||L2(Ωb)2\displaystyle\leq C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}+\epsilon\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||_{L^{2}(\Omega_{b})}^{2}
+C(ϵ)(0t||𝜼𝜼δ||2L2(Ωb)+0t||ωωδ||2H2(Γ)+0t||ppδ||2L2(Ωb)).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||p-p_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right).

Term T14. This term can be handled in the same way as terms T6T_{6}-T8T_{8}.

Term T15. We pass to the limit as ν0\nu\to 0 in (135) to obtain:

T15=α0tΩb(s)DDt𝜼(ppδ)+α0tΩδb,δ(s)DδDt𝜼δ(ppδ).T_{15}=-{\alpha}\int_{0}^{t}\int_{\Omega_{b}(s)}\frac{D}{Dt}\boldsymbol{\eta}\cdot\nabla(p-p_{\delta})+\alpha\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}\frac{{D}^{\delta}}{Dt}\boldsymbol{\eta}_{\delta}\cdot\nabla(p-p_{\delta}).

To estimate this term we pull back to the reference domain and use (21) and the cofactor formula for the matrix inverse to obtain:

|T15|\displaystyle|T_{15}| =α|0tΩb𝒥ηbt𝜼ηb(ppδ)0tΩb𝒥ηδδbt𝜼δηδδb(ppδ)|\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{\eta}_{b}\partial_{t}\boldsymbol{\eta}\cdot\nabla^{\eta}_{b}(p-p_{\delta})-\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}\partial_{t}\boldsymbol{\eta}_{\delta}\cdot\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta})\right|
=α|0tΩbt𝜼[(ppδ)(𝑰+𝜼)C]0tΩbt𝜼δ[(ppδ)(𝑰+𝜼δδ)C]|R15,1+R15,2,\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\partial_{t}\boldsymbol{\eta}\cdot\left[\nabla(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{C}\right]-\int_{0}^{t}\int_{\Omega_{b}}\partial_{t}\boldsymbol{\eta}_{\delta}\cdot\left[\nabla(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]\right|\leq{{R_{15,1}+R_{15,2}}},

where

R15,1\displaystyle{{R_{15,1}}} =α|0tΩbt𝜼[(ppδ)(𝜼𝜼δδ)C]|,\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\partial_{t}\boldsymbol{\eta}\cdot\left[\nabla(p-p_{\delta})\cdot(\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]\right|,
R15,2\displaystyle{{R_{15,2}}} =α|0tΩb(t𝜼t𝜼δ)[(ppδ)(𝑰+𝜼δδ)C]|.\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}(\partial_{t}\boldsymbol{\eta}-\partial_{t}\boldsymbol{\eta}_{\delta})\cdot\left[\nabla(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]\right|.

To estimate R15,1R_{15,1}, we use (115), (116), and the convolution inequality (170) to obtain:

R15,1\displaystyle{{R_{15,1}}} ϵ0t||(ppδ)||L2(Ωδb,δ(s))2+C(ϵ)0t||𝜼𝜼δ||L2(Ωb)2\displaystyle\leq\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}
+C(ϵ)(0t||𝜼𝜼δ||L2(Ωb)2+0t||ωωδ||2H2(Γ)).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}\right).

Here, we also used the following estimate on the norm of the gradient of the pressure on the reference domain and on the moving domain, which is obtained by using (115), (116), and (21):

||(ppδ)(t)||2L2(Ωb)\displaystyle||\nabla(p-p_{\delta})(t)||^{2}_{L^{2}(\Omega_{b})} =Ωb|(ppδ)|2=Ωb𝒥ηδδb|ηδδb(ppδ)(𝑰+𝜼δδ)|2(𝒥ηδδb)1\displaystyle=\int_{\Omega_{b}}|\nabla(p-p_{\delta})|^{2}=\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}|\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})|^{2}\cdot(\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b})^{-1}
CΩb𝒥ηδδb|ηδδb(ppδ)|2=C||(ppδ)(t)||L2(Ωδb,δ(t))2,\displaystyle\leq C\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}|\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta})|^{2}=C||\nabla(p-p_{\delta})(t)||_{L^{2}({\Omega}^{\delta}_{b,\delta}(t))}^{2}, (171)

where constant CC is independent of δ\delta and t[0,Tδ]t\in[0,T_{\delta}].

The estimate of R15,2R_{15,2} is straight forward:

R15,2ϵ0t||(ppδ)||L2(Ωδb,δ(s))2+C(ϵ)0t||t𝜼t𝜼δ||2L2(Ωb).{{R_{15,2}}}\leq\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\partial_{t}\boldsymbol{\eta}-\partial_{t}\boldsymbol{\eta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}.

From here we get the final estimate of T15T_{15}:

|T15|\displaystyle|T_{15}| ϵ0t||(ppδ)||L2(Ωδb,δ(s))2+C(ϵ)0t||𝜼𝜼δ||L2(Ωb)2\displaystyle\leq\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}
+C(ϵ)(0t||𝜼𝜼δ||L2(Ωb)2+0t||ωωδ||2H2(Γ)+0t||t𝜼t𝜼δ||2L2(Ωb)).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}+\int_{0}^{t}||\partial_{t}\boldsymbol{\eta}-\partial_{t}\boldsymbol{\eta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right).

Term T16. To estimate T16T_{16} defined in (136) we start by passing to the limit as ν0\nu\to 0 to obtain

T16=α0tΓ(s)(𝝃𝒏)(ppδ)+α0tΓδδ(s)(𝝃δ𝒏δδ)(ppδ),T_{16}=-{\alpha}\int_{0}^{t}\int_{\Gamma(s)}(\boldsymbol{\xi}\cdot\boldsymbol{n})(p-p_{\delta})+{\alpha}\int_{0}^{t}\int_{{\Gamma}^{\delta}_{\delta}(s)}(\boldsymbol{\xi}_{\delta}\cdot{\boldsymbol{n}}^{\delta}_{\delta})(p-p_{\delta}),

where 𝒏δδ{\boldsymbol{n}}^{\delta}_{\delta} is the upward pointing normal vector to Γδδ(t){\Gamma}^{\delta}_{\delta}(t). We integrate by parts to obtain that |T16|R16,1+R16,2|T_{16}|\leq R_{16,1}+R_{16,2}, where

R16,1:=α|0tΩb(s)(𝝃)(ppδ)0tΩδb,δ(s)(𝝃δ)(ppδ)|,{{R_{16,1}}}:={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}(s)}(\nabla\cdot\boldsymbol{\xi})(p-p_{\delta})-\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}(\nabla\cdot\boldsymbol{\xi}_{\delta})(p-p_{\delta})\right|,
R16,2:=α|0tΩb(s)𝝃(ppδ)0tΩδb,δ(s)𝝃δ(ppδ)|.{{R_{16,2}}}:={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}(s)}\boldsymbol{\xi}\cdot\nabla(p-p_{\delta})-\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}\boldsymbol{\xi}_{\delta}\cdot\nabla(p-p_{\delta})\right|.

By using (21) and the bootstrap assumption (117), we have that

R16,1\displaystyle{{R_{16,1}}} =α|0tΩb𝒥ηb(tr(ηb𝝃))(ppδ)0tΩb𝒥ηδδb(tr(ηδδb𝝃δ))(ppδ)|\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{\eta}_{b}(\text{tr}(\nabla^{\eta}_{b}\boldsymbol{\xi}))(p-p_{\delta})-\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}(\text{tr}(\nabla^{{\eta}^{\delta}_{\delta}}_{b}\boldsymbol{\xi}_{\delta}))(p-p_{\delta})\right|
=α|0tΩbtr(𝝃(𝑰+𝜼)C)(ppδ)0tΩbtr(𝝃δ(𝑰+𝜼δδ)C)(ppδ)|\displaystyle={\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\text{tr}(\nabla\boldsymbol{\xi}\cdot(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{C})(p-p_{\delta})-\int_{0}^{t}\int_{\Omega_{b}}\text{tr}(\nabla\boldsymbol{\xi}_{\delta}\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C})(p-p_{\delta})\right|
α|0tΩbtr(𝝃(𝜼𝜼δδ)C)(ppδ)|+α|0tΩbtr((𝝃𝝃δ)(𝑰+𝜼δδ)C)(ppδ)|\displaystyle\leq{\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\text{tr}(\nabla\boldsymbol{\xi}\cdot(\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C})(p-p_{\delta})\right|+{\alpha}\left|\int_{0}^{t}\int_{\Omega_{b}}\text{tr}(\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C})(p-p_{\delta})\right|
C0t||𝜼𝜼δδ||L2(Ωb)||ppδ||L2(Ωb)+C0t||𝝃𝝃δ||L2(Ωb)||ppδ||L2(Ωb).\displaystyle\leq C\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}||_{L^{2}(\Omega_{b})}\cdot||p-p_{\delta}||_{L^{2}(\Omega_{b})}+C\int_{0}^{t}||\nabla\boldsymbol{\xi}-\nabla\boldsymbol{\xi}_{\delta}||_{L^{2}(\Omega_{b})}\cdot||p-p_{\delta}||_{L^{2}(\Omega_{b})}.

For R16,2R_{16,2}, we compute

R16,2\displaystyle{{R_{16,2}}} =α|0tΩb𝝃[(ppδ)(𝑰+𝜼)C]0tΩb𝝃δ[(ppδ)(𝑰+𝜼δδ)C]|\displaystyle=\alpha\left|\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{\xi}\cdot\left[\nabla(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{C}\right]-\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{\xi}_{\delta}\cdot\left[\nabla(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]\right|
α|0tΩb𝝃[(ppδ)(𝜼𝜼δδ)C]|+α|0tΩb(𝝃𝝃δ)[(ppδ)(𝑰+𝜼δδ)C]|\displaystyle\leq\alpha\left|\int_{0}^{t}\int_{\Omega_{b}}\boldsymbol{\xi}\cdot\left[\nabla(p-p_{\delta})\cdot(\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]\right|+\alpha\left|\int_{0}^{t}\int_{\Omega_{b}}(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})\cdot\left[\nabla(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]\right|
C0t||ppδ||L2(Ωb)||𝜼𝜼δδ||L2(Ωb)+C0t||𝝃𝝃δ||L2(Ωb)||ppδ||L2(Ωb).\displaystyle\leq C\int_{0}^{t}||\nabla p-\nabla p_{\delta}||_{L^{2}(\Omega_{b})}\cdot||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}||_{L^{2}(\Omega_{b})}+C\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||_{L^{2}(\Omega_{b})}\cdot||\nabla p-\nabla p_{\delta}||_{L^{2}(\Omega_{b})}.

By the convolution inequality (170) and the previous estimate on the gradient of the pressure (A.2), we conclude that

|T16|ϵ(0t||𝝃𝝃δ||2L2(Ωb)+0t||ppδ||L2(Ωδb,δ(s))2)+C(ϵ)(0t||ppδ||2L2(Ωb)+0t||𝜼𝜼δ||L2(Ωb)2+0t||𝜼𝜼δ||L2(Ωb)2+0t||ωωδ||H2(Γ)2+0t||𝝃𝝃δ||2L2(Ωb)).|T_{16}|\leq\epsilon\left(\int_{0}^{t}||\nabla\boldsymbol{\xi}-\nabla\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\nabla p-\nabla p_{\delta}||_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}^{2}\right)+C(\epsilon)\left(\int_{0}^{t}||p-p_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right.\\ \left.+\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}+\int_{0}^{t}||\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta}||^{2}_{L^{2}(\Omega_{b})}\right).

Term T17. To estimate term T17T_{17} defined in (137) we use (21) to compute

T17\displaystyle T_{17} =κ0tΩb𝒥ηbηbpηb(ppδ)κ0tΩb𝒥ηδδbηδδbpδηδδb(ppδ)\displaystyle=\kappa\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{\eta}_{b}\nabla^{\eta}_{b}p\cdot\nabla^{\eta}_{b}(p-p_{\delta})-\kappa\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}\nabla^{{\eta}^{\delta}_{\delta}}_{b}p_{\delta}\cdot\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta})
=κ0tΩb𝒥ηδδbηδδb(ppδ)ηδδb(ppδ)+I1+I2=κ0tΩδb,δ(t)|(ppδ)|2+R17,1+R17,2,\displaystyle=\kappa\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta})\cdot\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta})+I_{1}+I_{2}=\kappa\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(t)}|\nabla(p-p_{\delta})|^{2}+{{R_{17,1}+R_{17,2}}},

where

R17,1\displaystyle{{R_{17,1}}} =κ0tΩb𝒥ηbηbpηb(ppδ)κ0tΩb𝒥ηδδbηbpηδδb(ppδ),\displaystyle=\kappa\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{\eta}_{b}\nabla^{\eta}_{b}p\cdot\nabla^{\eta}_{b}(p-p_{\delta})-\kappa\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}\nabla^{\eta}_{b}p\cdot\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta}),
R17,2\displaystyle{{R_{17,2}}} =κ0tΩb𝒥ηδδb(ηbpηδδbp)ηδδb(ppδ).\displaystyle=\kappa\int_{0}^{t}\int_{\Omega_{b}}\mathcal{J}^{{\eta}^{\delta}_{\delta}}_{b}(\nabla^{\eta}_{b}p-\nabla^{{\eta}^{\delta}_{\delta}}_{b}p)\cdot\nabla^{{\eta}^{\delta}_{\delta}}_{b}(p-p_{\delta}).

To estimate R17,1R_{17,1}, we use (21) to obtain

R17,1=κ0tΩbηbp((ppδ)[(𝑰+𝜼)C(𝑰+𝜼δδ)C]).{{R_{17,1}}}=\kappa\int_{0}^{t}\int_{\Omega_{b}}\nabla^{\eta}_{b}p\cdot\Big{(}\nabla(p-p_{\delta})\cdot\left[(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{C}-(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]\Big{)}.

Because 𝜼\boldsymbol{\eta} is smooth, |ηbp|C|\nabla^{\eta}_{b}p|\leq C uniformly in space and time. Therefore,

|R17,1|C0t||(ppδ)||L2(Ωb)||(𝜼𝜼δδ)C||L2(Ωb).{{|R_{17,1}|}}\leq C\int_{0}^{t}||\nabla(p-p_{\delta})||_{L^{2}(\Omega_{b})}\cdot||(\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}||_{L^{2}(\Omega_{b})}.

Using the estimate in (A.2), we obtain the desired estimate that

|R17,1|ϵ0t||(ppδ)||L2(Ωδb,δ(s))2+C(ϵ)0t||𝜼𝜼δδ||2L2(Ωb).{{|R_{17,1}|}}\leq\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}^{2}+C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}.

To estimate R17,2R_{17,2}, we use the bootstrap assumption (117) that there exists a constant CC (independent of δ\delta) such that |𝜼δδ|C|\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}|\leq C pointwise for t[0,Tδ]t\in[0,T_{\delta}]. Therefore, |(𝑰+𝜼δ)C||(\boldsymbol{I}+\nabla\boldsymbol{\eta}_{\delta})^{C}| is pointwise uniformly bounded in space and time on the time interval [0,Tδ][0,T_{\delta}]. Thus, by (21),

R17,2=κ0tΩb(ηbpηδδbp)[(ppδ)(𝑰+𝜼δδ)C]{{R_{17,2}}}=\kappa\int_{0}^{t}\int_{\Omega_{b}}(\nabla^{\eta}_{b}p-\nabla^{{\eta}^{\delta}_{\delta}}_{b}p)\cdot\left[\nabla(p-p_{\delta})\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{C}\right]

and hence

|R17,2|C0t||ηbpηδδbp||L2(Ωb)||(ppδ)||L2(Ωb).{{|R_{17,2}|}}\leq C\int_{0}^{t}||\nabla^{\eta}_{b}p-\nabla^{{\eta}^{\delta}_{\delta}}_{b}p||_{L^{2}(\Omega_{b})}\cdot||\nabla(p-p_{\delta})||_{L^{2}(\Omega_{b})}.

We estimate the first pressure term by using (21) to obtain

||ηbpηδδbp||L2(Ωb)2\displaystyle||\nabla^{\eta}_{b}p-\nabla^{{\eta}^{\delta}_{\delta}}_{b}p||_{L^{2}(\Omega_{b})}^{2} =Ωb|p[(𝑰+𝜼)1(𝑰+𝜼δδ)1]|2\displaystyle=\int_{\Omega_{b}}\left|\nabla p\cdot\left[(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{-1}-(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{-1}\right]\right|^{2}
=Ωb|p(𝑰+𝜼δδ)1[(𝑰+𝜼δδ)(𝑰+𝜼)1𝑰]|2\displaystyle=\int_{\Omega_{b}}\left|\nabla p\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{-1}[(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{-1}-\boldsymbol{I}]\right|^{2}
=Ωb|p(𝑰+𝜼δδ)1[(𝑰+𝜼δδ)(𝑰+𝜼)](𝑰+𝜼)1|2\displaystyle=\int_{\Omega_{b}}\left|\nabla p\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{-1}[(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})-(\boldsymbol{I}+\nabla\boldsymbol{\eta})](\boldsymbol{I}+\nabla\boldsymbol{\eta})^{-1}\right|^{2}
=Ωb|p(𝑰+𝜼δδ)1(𝜼δδ𝜼)(𝑰+𝜼)1|2.\displaystyle=\int_{\Omega_{b}}\left|\nabla p\cdot(\boldsymbol{I}+\nabla{\boldsymbol{\eta}}^{\delta}_{\delta})^{-1}(\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}-\nabla\boldsymbol{\eta})(\boldsymbol{I}+\nabla\boldsymbol{\eta})^{-1}\right|^{2}.

Using the fact that pp is smooth and the bootstrap assumption (116), we have that

||ηbpηδδbp||L2(Ωb)2C||𝜼δδ𝜼||L2(Ωb)2.||\nabla^{\eta}_{b}p-\nabla^{{\eta}^{\delta}_{\delta}}_{b}p||_{L^{2}(\Omega_{b})}^{2}\leq C||\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}-\nabla\boldsymbol{\eta}||_{L^{2}(\Omega_{b})}^{2}.

Therefore, combining this with (A.2) we obtain

R17,2ϵ0t||(ppδ)||2L2(Ωδb,δ(s))+C(ϵ)0t||𝜼𝜼δδ||2L2(Ωb).{{R_{17,2}}}\leq\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||^{2}_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}+C(\epsilon)\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}_{\delta}||^{2}_{L^{2}(\Omega_{b})}.

The final estimate of T17T_{17} now follows after the application of the convolution inequality (170):

T17κ0tΩδb,δ(s)|(ppδ)|2+R17,T_{17}\leq\kappa\int_{0}^{t}\int_{{\Omega}^{\delta}_{b,\delta}(s)}|\nabla(p-p_{\delta})|^{2}+{{R_{17}}},

where the remainder is bounded by

|R17|\displaystyle{{|R_{17}|}} ϵ0t||(ppδ)||2L2(Ωδb,δ(s))\displaystyle\leq\epsilon\int_{0}^{t}||\nabla(p-p_{\delta})||^{2}_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}
+C(ϵ)(0t||𝜼𝜼δ||L2(Ωb)2+0t||𝜼𝜼δ||L2(Ωb)2+0t||ωωδ||H2(Γ)2).\displaystyle+C(\epsilon)\left(\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla{\boldsymbol{\eta}}^{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\nabla\boldsymbol{\eta}-\nabla\boldsymbol{\eta}_{\delta}||_{L^{2}(\Omega_{b})}^{2}+\int_{0}^{t}||\omega-\omega_{\delta}||_{H^{2}(\Gamma)}^{2}\right).

Term 18. Here want to estimate

T18\displaystyle T_{18} =0tΓ(s)p(𝒖𝝃)𝒏0tΓ(s)p(𝒖δ𝝃δ)𝒏0tΓδ(s)pδ(𝒖𝝃)𝒏δ+0tΓδ(s)pδ(𝒖δ𝝃δ)𝒏δ\displaystyle=\int_{0}^{t}\int_{\Gamma(s)}p(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n}-\int_{0}^{t}\int_{\Gamma(s)}p(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n}-\int_{0}^{t}\int_{\Gamma_{\delta}(s)}p_{\delta}(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n}_{\delta}+\int_{0}^{t}\int_{\Gamma_{\delta}(s)}p_{\delta}(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n}_{\delta}
0tΓ(s)((𝒖𝝃)𝒏)(ppδ)+0tΓδ(s)((𝒖δ𝝃δ)𝒏δ)(ppδ)\displaystyle-\int_{0}^{t}\int_{\Gamma(s)}((\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n})(p-p_{\delta})+\int_{0}^{t}\int_{\Gamma_{\delta}(s)}((\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n}_{\delta})(p-p_{\delta})
=0tΓ(s)p(𝒖δ𝝃δ)𝒏0tΓδ(s)pδ(𝒖𝝃)𝒏δ+0tΓ(s)((𝒖𝝃)𝒏)pδ+0tΓδ(s)((𝒖δ𝝃δ)𝒏δ)p.\displaystyle=-\int_{0}^{t}\int_{\Gamma(s)}p(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n}-\int_{0}^{t}\int_{\Gamma_{\delta}(s)}p_{\delta}(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n}_{\delta}+\int_{0}^{t}\int_{\Gamma(s)}((\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{n})p_{\delta}+\int_{0}^{t}\int_{\Gamma_{\delta}(s)}((\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{n}_{\delta})p.

By mapping all of the integrals back to the reference domain Γ\Gamma, we obtain

T18\displaystyle T_{18} =0tΓp(𝒖δ𝝃δ)(xω,1)0tΓpδ(𝒖𝝃)(xωδ,1)\displaystyle=-\int_{0}^{t}\int_{\Gamma}p(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot(-\partial_{x}\omega,1)-\int_{0}^{t}\int_{\Gamma}p_{\delta}(\boldsymbol{u}-\boldsymbol{\xi})\cdot(-\partial_{x}\omega_{\delta},1)
+0tΓpδ(𝒖𝝃)(xω,1)+0tΓp(𝒖δ𝝃δ)(xωδ,1)\displaystyle+\int_{0}^{t}\int_{\Gamma}p_{\delta}(\boldsymbol{u}-\boldsymbol{\xi})\cdot(-\partial_{x}\omega,1)+\int_{0}^{t}\int_{\Gamma}p(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot(-\partial_{x}\omega_{\delta},1)
=0tΓp(𝒖δ𝝃δ)𝒆x(xωxωδ)0tΓpδ(𝒖𝝃)𝒆x(xωxωδ)\displaystyle=\int_{0}^{t}\int_{\Gamma}p(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{e}_{x}(\partial_{x}\omega-\partial_{x}\omega_{\delta})-\int_{0}^{t}\int_{\Gamma}p_{\delta}(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{e}_{x}(\partial_{x}\omega-\partial_{x}\omega_{\delta})
=0tΓp[(𝒖𝝃)𝒆x(𝒖δ𝝃δ)𝒆x](xωxωδ)+0tΓ(ppδ)(𝒖𝝃)𝒆x(xωxωδ).\displaystyle=-\int_{0}^{t}\int_{\Gamma}p[(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{e}_{x}-(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{e}_{x}](\partial_{x}\omega-\partial_{x}\omega_{\delta})+\int_{0}^{t}\int_{\Gamma}(p-p_{\delta})(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{e}_{x}(\partial_{x}\omega-\partial_{x}\omega_{\delta}).

The absolute value is bounded as follows:

|0tΓp[(𝒖𝝃)𝒆x(𝒖δ𝝃δ)𝒆x](xωxωδ)|+|0tΓ(ppδ)(𝒖𝝃)𝒆x(xωxωδ)|C(0t||(𝒖𝝃)𝒆x(𝒖δ𝝃δ)𝒆x||L2(Γ)||xωxωδ||L2(Γ)+0t||ppδ||L2(Γ)||xωxωδ||L2(Γ)).\left|\int_{0}^{t}\int_{\Gamma}p[(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{e}_{x}-(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{e}_{x}](\partial_{x}\omega-\partial_{x}\omega_{\delta})\right|+\left|\int_{0}^{t}\int_{\Gamma}(p-p_{\delta})(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{e}_{x}(\partial_{x}\omega-\partial_{x}\omega_{\delta})\right|\\ \leq C\left(\int_{0}^{t}||(\boldsymbol{u}-\boldsymbol{\xi})\cdot\boldsymbol{e}_{x}-(\boldsymbol{u}_{\delta}-\boldsymbol{\xi}_{\delta})\cdot\boldsymbol{e}_{x}||_{L^{2}(\Gamma)}||\partial_{x}\omega-\partial_{x}\omega_{\delta}||_{L^{2}(\Gamma)}+\int_{0}^{t}||p-p_{\delta}||_{L^{2}(\Gamma)}||\partial_{x}\omega-\partial_{x}\omega_{\delta}||_{L^{2}(\Gamma)}\right).

After the application of the trace theorem, Poincare’s inequality, and Korn’s inequality we obtain the final estimate:

|T18|\displaystyle|T_{18}| ϵ(0t||𝑫(𝒖^𝒖δ)||2L2(Ωf,δ(s))+0t||(𝝃𝝃δ)||2L2(Ωb)+0t||(ppδ)||2L2(Ωδb,δ(s)))\displaystyle\leq\epsilon\left(\int_{0}^{t}||\boldsymbol{D}(\widehat{\boldsymbol{u}}-\boldsymbol{u}_{\delta})||^{2}_{L^{2}(\Omega_{f,\delta}(s))}+\int_{0}^{t}||\nabla(\boldsymbol{\xi}-\boldsymbol{\xi}_{\delta})||^{2}_{L^{2}(\Omega_{b})}+\int_{0}^{t}||\nabla(p-p_{\delta})||^{2}_{L^{2}({\Omega}^{\delta}_{b,\delta}(s))}\right)
+C(ϵ)0t||ωωδ||2H2(Γ).\displaystyle+C(\epsilon)\int_{0}^{t}||\omega-\omega_{\delta}||^{2}_{H^{2}(\Gamma)}.

A.3 Generalized Aubin-Lions Compactness Theorem [57]

To help the reader follow the results from Section 8.5.2 we state here the Generalized Aubin-Lions Compactness Theorem, i.e., Theorem 3.1 of [57].

Theorem A.1.

(The Generalized Aubin-Lions Compactness Theorem) Let VV and HH be Hilbert spaces such that VHV\subset\subset H. Suppose that {𝐮Δt}L2(0,T;H)\{{\bf u}_{\Delta t}\}\subset L^{2}(0,T;H) is a sequence such that 𝐮Δt(t,)=𝐮Δtn()on((n1)Δt,nΔt],n=1,,N,{\bf u}_{\Delta t}(t,\cdot)={\bf u}_{\Delta t}^{n}(\cdot)\ {\rm on}\;((n-1){\Delta t},n{\Delta t}],\;n=1,\dots,N, with NΔt=TN\Delta t=T. Let VΔtnV_{\Delta t}^{n} and QΔtnQ_{\Delta t}^{n} be Hilbert spaces such that (VΔtn,QΔtn)V×V(V_{\Delta t}^{n},Q_{\Delta t}^{n})\hookrightarrow V\times V, where the embeddings are uniformly continuous w.r.t. Δt{\Delta t} and nn, and VΔtnQΔtn¯H(QΔtn)V_{\Delta t}^{n}\subset\subset\overline{Q_{\Delta t}^{n}}^{H}\hookrightarrow(Q_{\Delta t}^{n})^{\prime}. Let 𝐮ΔtnVΔtn{\bf u}_{\Delta t}^{n}\in V_{\Delta t}^{n}, n=1,,Nn=1,\dots,N. If the following is true:

  • (A)

    There exists a universal constant C>0C>0 such that for every Δt{\Delta t}

    1. A1.

      n=1N𝐮nΔtVnΔt2ΔtC,\sum_{n=1}^{N}\|{\bf u}^{n}_{\Delta t}\|_{V^{n}_{\Delta t}}^{2}{\Delta t}\leq C,

    2. A2.

      𝐮ΔtL(0,T;H)C,\|{\bf u}_{\Delta t}\|_{L^{\infty}(0,T;H)}\leq C,

    3. A3.

      τΔt𝐮Δt𝐮Δt2L2(Δt,T;H)CΔt.\|\tau_{\Delta t}{\bf u}_{\Delta t}-{\bf u}_{\Delta t}\|^{2}_{L^{2}({\Delta t},T;H)}\leq C{\Delta t}.

  • (B)

    There exists a universal constant C>0C>0 such that

    PnΔt𝐮n+1Δt𝐮nΔtΔt(QnΔt)C(𝐮n+1ΔtVn+1Δt+1),n=0,,N1,\|P^{n}_{\Delta t}\frac{{\bf u}^{n+1}_{\Delta t}-{\bf u}^{n}_{\Delta t}}{{\Delta t}}\|_{(Q^{n}_{\Delta t})^{\prime}}\leq C(\|{\bf u}^{n+1}_{\Delta t}\|_{V^{n+1}_{\Delta t}}+1),\;n=0,\dots,N-1,

    where PnΔtP^{n}_{\Delta t} is the orthogonal projector onto QnΔt¯H\overline{Q^{n}_{\Delta t}}^{H}.

  • (C)

    The function spaces QΔtnQ_{\Delta t}^{n} and VΔtnV_{\Delta t}^{n} depend smoothly on time in the following sense:

    1. C1.

      For every Δt>0{\Delta t}>0, and for every l{1,,N}l\in\{1,\dots,N\} and n{1,,Nl}n\in\{1,\dots,N-l\}, there exists a space Qn,lΔtVQ^{n,l}_{\Delta t}\subset V and the operators JiΔt,l,n:Qn,lΔtQn+iΔt,i=0,1,,l,{{J}}^{i}_{\Delta t,l,n}:Q^{n,l}_{\Delta t}\to Q^{n+i}_{\Delta t},i=0,1,\dots,l, such that JiΔt,l,n𝐪Qn+iΔtC𝐪Qn,lΔt,𝐪Qn,lΔt\|J^{i}_{\Delta t,l,n}{\bf q}\|_{Q^{n+i}_{\Delta t}}\leq C\|{\bf q}\|_{Q^{n,l}_{\Delta t}},\ \forall{\bf q}\in Q^{n,l}_{\Delta t}, and

      ((Jj+1Δt,l,n𝐪JjΔt,l,n𝐪),𝐮n+j+1Δt)HCΔt𝐪Qn,lΔt𝐮n+j+1ΔtVn+j+1Δt,j{0,,l1},\Big{(}(J^{j+1}_{\Delta t,l,n}{\bf q}-J^{j}_{\Delta t,l,n}{\bf q}),{\bf u}^{n+j+1}_{\Delta t}\Big{)}_{H}\leq C{\Delta t}\|{\bf q}\|_{Q^{n,l}_{\Delta t}}\|{\bf u}^{n+j+1}_{\Delta t}\|_{V^{n+j+1}_{\Delta t}},\quad j\in\{0,\dots,l-1\}, (172)
      JiΔt,l,n𝐪𝐪HClΔt𝐪Qn,lΔt,i{0,,l},\|J^{i}_{\Delta t,l,n}{\bf q}-{\bf q}\|_{H}\leq C\sqrt{l{\Delta t}}\|{\bf q}\|_{Q^{n,l}_{\Delta t}},\quad i\in\{0,\dots,l\}, (173)

      where C>0C>0 is independent of Δt,n{\Delta t},n and ll.

    2. C2.

      Let Vn,lΔt=Qn,lΔt¯VV^{n,l}_{\Delta t}=\overline{Q^{n,l}_{\Delta t}}^{V}. There exist the functions IiΔt,l,n:Vn+iΔtVn,lΔt,i=0,1,,l,I^{i}_{{\Delta t},l,n}:V^{n+i}_{\Delta t}\to V^{n,l}_{\Delta t},\ i=0,1,\dots,l, and a universal constant C>0C>0, such that for every 𝐯Vn+iΔt{\bf v}\in V^{n+i}_{\Delta t}

      IiΔt,l,n𝐯Vn,lΔtC𝐯Vn+iΔt,i{0,,l},\|I^{i}_{{\Delta t},l,n}{\bf v}\|_{V^{n,l}_{\Delta t}}\leq C\|{\bf v}\|_{V^{n+i}_{\Delta t}},\quad i\in\{0,\dots,l\}, (174)
      IiΔt,l,n𝐯𝐯Hg(lΔt)𝐯Vn+iΔt,i{0,,l},\|I^{i}_{{\Delta t},l,n}{\bf v}-{\bf v}\|_{H}\leq g(l{\Delta t})\|{\bf v}\|_{V^{n+i}_{\Delta t}},\quad i\in\{0,\dots,l\}, (175)

      where g:++g:{\mathbb{R}}_{+}\to{\mathbb{R}}_{+} is a universal, monotonically increasing function such that g(h)0g(h)\to 0 as h0h\to 0.

    3. C3.

      Uniform Ehrling property: For every δ>0\delta>0 there exists a constant C(δ)C(\delta) independent of n,ln,l and Δt{\Delta t}, such that

      𝐯Hδ𝐯Vn,lΔt+C(δ)𝐯(Qn,lΔt),𝐯Vn,lΔt.\|{\bf v}\|_{H}\leq\delta\|{\bf v}\|_{V^{n,l}_{\Delta t}}+C(\delta)\|{\bf v}\|_{(Q^{n,l}_{\Delta t})^{\prime}},\quad{\bf v}\in V^{n,l}_{\Delta t}. (176)

then {𝐮Δt}\{{\bf u}_{\Delta t}\} is relatively compact in L2(0,T;H)L^{2}(0,T;H).

References

  • [1] R. A. Adams. Sobolev spaces, volume 65 of Pure and Applied Mathematics. Academic Press, New York-London, 1975.
  • [2] I. Ambartsumyan, V. J. Ervin, T. Nguyen, and I. Yotov. A nonlinear Stokes-Biot model for the interaction of a non-Newtonian fluid with poroelastic media. ESAIM Math. Model. Numer. Anal., 53(6):1915–1955, 2019.
  • [3] S. Badia, A. Quaini, and A. Quarteroni. Coupling Biot and Navier-Stokes equations for modelling fluid-poroelastic media interaction. Journal of Computational Physics, 228(21):7986–8014, 2009.
  • [4] V. Barbu, Z. Grujić, I. Lasiecka, and A. Tuffaha. Existence of the energy-level weak solutions for a nonlinear fluid-structure interaction model. In Fluids and waves, volume 440 of Contemp. Math., pages 55–82. Amer. Math. Soc., Providence, RI, 2007.
  • [5] V. Barbu, Z. Grujić, I. Lasiecka, and A. Tuffaha. Smoothness of weak solutions to a nonlinear fluid-structure interaction model. Indiana Univ. Math. J., 57(3):1173–1207, 2008.
  • [6] H. Barucq, M. Madaune-Tort, and P. Saint-Macary. Theoretical aspects of wave propagation for Biot’s consolidation problem. Monografías del Seminario Matemático García de Galdeano, 31:449–458, 2004.
  • [7] H. Barucq, M. Madaune-Tort, and P. Saint-Macary. On nonlinear Biot’s consolidation models. Nonlinear Anal., 63:e985–e995, 2005.
  • [8] H. Beirão da Veiga. On the existence of strong solutions to a coupled fluid-structure evolution problem. J. Math. Fluid Mech., 6(1):21–52, 2004.
  • [9] M. A. Biot. General theory of three-dimensional consolidation. J. Appl. Phys., 12(2):155–164, 1941.
  • [10] M. A. Biot. Theory of elasticity and consolidation for a porous anisotropic solid. J. Appl. Phys., 26(2):182–185, 1955.
  • [11] L. Bociu, G. Guidoboni, R. Sacco, and J. T. Webster. Analysis of nonlinear poro-elastic and poro-visco-elastic models. Arch. Ration. Mech. Anal., 222(3):1445–1519, 2016.
  • [12] L. Bociu, B. Muha, and J. T. Webster. Weak solutions in nonlinear poroelasticity with incompressible constituents. Nonlinear Anal. Real World Appl., 67:Paper No. 103563, 22, 2022.
  • [13] L. Bociu, B. Muha, and J. T. Webster. Mathematical effects of linear visco-elasticity in quasi-static Biot models. J. Math. Anal. Appl., 527(2):Paper No. 127462, 2023.
  • [14] L. Bociu, S. Čanić, B. Muha, and J. T. Webster. Multilayered poroelasticity interacting with Stokes flow. SIAM J. Math. Anal., 53(6):6243–6279, 2021.
  • [15] L. Bociu and J. T. Webster. Nonlinear quasi-static poroelasticity. J. Differential Equations, 296:242–278, 2021.
  • [16] S. C. Brenner and L. R. Scott. The Mathematical Theory of Finite Element Methods, volume 15 of Texts in Applied Mathematics. Springer Science+Business Media, LLC, New York, third edition, 2008.
  • [17] M. Bukac, P. Zunino, and I. Yotov. Explicit partitioning strategies for the interaction between a fluid and a multilayered poroelastic structure: an operator-splitting approach. Journal of Computational Physics, 228(21):7986–8014, 2013.
  • [18] S. Canic, Y. Wang, and M. Bukač. A next-generation mathematical model for drug eluting stents. SIAM J. Appl. Math., 81(4):1503–1529, 2021.
  • [19] P. Causin, G. Guidoboni, A. Harris, D. Prada, R. Sacco, and S. Terragni. A poroelastic model for the perfusion of the lamina cribrosa in the optic nerve head. Math. Biosci., 257:33–41, 2014.
  • [20] A. Cesmelioglu. Analysis of the coupled Navier-Stokes/Biot problem. J. Math. Anal. Appl., 456(2):970–991, 2017.
  • [21] 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.
  • [22] N. V. Chemetov, Š. Nečasová, and B. Muha. Weak-strong uniqueness for fluid-rigid body interaction problem with slip boundary condition. J. Math. Phys., 60(1):011505, 13, 2019.
  • [23] C. H. A. Cheng, D. Coutand, and S. Shkoller. Navier-Stokes equations interacting with a nonlinear elastic biofluid shell. SIAM J. Math. Anal., 39(3):742–800, 2007.
  • [24] C. H. A. Cheng and S. Shkoller. The interaction of the 3D Navier-Stokes equations with a moving nonlinear Koiter elastic shell. SIAM J. Math. Anal., 42(3):1094–1155, 2010.
  • [25] P. G. Ciarlet. Mathematical Elasticity Volume I: Three-Dimensional Elasticity, volume 20 of Studies in Mathematics and Its Applications. Elsevier Science Publishers B.V., Amsterdam, 1988.
  • [26] D. Cioranescu and J. Saint Jean Paulin. Homogenization of reticulated structures, volume 136 of Applied Mathematical Sciences. Springer-Verlag, New York, 1999.
  • [27] C. Conca, F. Murat, and O. Pironneau. The Stokes and Navier-Stokes equations with boundary conditions involving the pressure. Japan. J. Math., 20(2):279–318, 1994.
  • [28] D. Coutand and S. Shkoller. Motion of an elastic solid inside an incompressible viscous fluid. Arch. Ration. Mech. Anal., 176(1):25–102, 2005.
  • [29] D. Coutand and S. Shkoller. The interaction between quasilinear elastodynamics and the Navier-Stokes equations. Arch. Ration. Mech. Anal., 179(3):303–352, 2006.
  • [30] M. Discacciati and A. Quarteroni. Navier-Stokes/Darcy coupling: modeling, analysis, and numerical approximation. Rev. Mat. Complut., 22(2):315–426, 2009.
  • [31] M. Dreher and A. Jüngel. Compact families of piecewise constant functions in Lp(0,T;B){L}^{p}(0,{T};{B}). Nonlinear Anal., 75(6):3072–3077, 2012.
  • [32] L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, second edition, 2010.
  • [33] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. Studies in Advanced Mathematics. CRC Press, Boca Raton, FL, 1992.
  • [34] V. Girault, M. F. Wheeler, B. Ganis, and M. E. Mear. A lubrication fracture model in a poro-elastic medium. Math. Models Methods Appl. Sci., 25(4):587–645, 2015.
  • [35] R. Glowinski. Finite element methods for incompressible viscous flow, in: P.G.Ciarlet, J.-L.Lions (Eds), Handbook of numerical analysis, volume 9. North-Holland, Amsterdam, 2003.
  • [36] 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.
  • [37] 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.
  • [38] C. Grandmont, M. Lukáčová-Medvid’ová, and Š. Nečasová. Mathematical and numerical analysis of some FSI problems. In T. Bodnár, G. P. Galdi, and Š. Nečasová, editors, Fluid-structure interaction and biomedical applications, Advances in Mathematical Fluid Mechanics, pages 1–77. Birkhäuser, 2014.
  • [39] M. Ignatova, I. Kukavica, I. Lasiecka, and A. Tuffaha. On well-posedness for a free boundary fluid-structure model. J. Math. Phys., 53(11):115624, 13, 2012.
  • [40] M. Ignatova, I. Kukavica, I. Lasiecka, and A. Tuffaha. On well-posedness and small data global existence for an interface damped free boundary fluid-structure model. Nonlinearity, 27(3):467–499, 2014.
  • [41] A. Inoue and M. Wakimoto. On existence of solutions of the Navier-Stokes equation in a time dependent domain. J. Fac. Sci. Univ. Tokyo Sect. IA Math., 24(2):303–319, 1977.
  • [42] W. Jäger and A. Mikelić. On the boundary conditions at the contact interface between a porous medium and a free fluid. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 23(3):403–465, 1996.
  • [43] W. Jäger and A. Mikelić. On the interface boundary condition of Beavers, Joseph, and Saffman. SIAM J. Appl. Math., 60(4):1111–1127, 2000.
  • [44] J. Kuan, S. Čanić, and B. Muha. Existence of a weak solution to a regularized moving boundary fluid-structure interaction problem with poroelastic media. Comptes Rendus Mécanique, 351(S1):1–30, 2023.
  • [45] I. Kukavica and A. Tuffaha. Solutions to a fluid-structure interaction free boundary problem. DCDS-A, 32(4):1355–1389, 2012.
  • [46] I. Kukavica, A. Tuffaha, and M. Ziane. Strong solutions for a fluid structure interaction system. Adv. Differential Equations, 15(3-4):231–254, 2010.
  • [47] D. Lengeler and M. Růžička. Weak solutions for an incompressible Newtonian fluid interacting with a Koiter type shell. Arch. Ration. Mech. Anal., 211(1):205–255, 2014.
  • [48] J. Lequeurre. Existence of strong solutions to a fluid-structure system. SIAM J. Math. Anal., 43(1):389–410, 2011.
  • [49] M. Lesinigo, C. D’Angelo, and A. Quarteroni. A multiscale Darcy-Brinkman model for fluid flow in fractured porous media. Numer. Math., 117(4):717–752, 2011.
  • [50] S. E. Mikhailov. Traces, extensions and co-normal derivatives for elliptic systems on Lipschitz domains. J. Math. Anal. Appl., 378:324–342, 2011.
  • [51] B. Muha. A note on the Trace Theorem for domains which are locally subgraph of a hölder continuous function. Netw. Heterog. Media, 9(1):191–196, 2014.
  • [52] 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.
  • [53] B. Muha and S. Čanić. A nonlinear, 3D fluid-structure interaction problem driven by the time-dependent dynamic pressure data: a constructive existence proof. Commun. Inf. Syst., 13(3):357–397, 2013.
  • [54] 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.
  • [55] B. Muha and S. Čanić. Fluid-structure interaction between an incompressible, viscous 3D fluid and an elastic shell with nonlinear Koiter membrane energy. Interfaces Free Bound., 17(4):465–495, 2015.
  • [56] 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.
  • [57] 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.
  • [58] 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.
  • [59] B. Muha, Š. Nečasová, and A. Radošević. A uniqueness result for 3D incompressible fluid-rigid body interaction problem. J. Math. Fluid Mech., 23(1):Paper No. 1, 39, 2021.
  • [60] S. Owczarek. A Galerkin method for Biot consolidation model. Math. Mech. Solids, 15(1):42–56, 2010.
  • [61] J.-P. Raymond and M. Vanninathan. A fluid-structure model coupling the Navier-Stokes equations and the Lamé system. J. Math. Pures Appl. (9), 102(3):546–596, 2014.
  • [62] A. Scharf, S. Čanić, and Y. Wang. A partitioned scheme for fluid-structure interaction with multilayered poroelastic media. In draft form., 2024.
  • [63] 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.
  • [64] A. Seboldt, O. Oyekole, J. Tambača, and M. Bukač. Numerical modeling of the fluid-porohyperelastic structure interaction. SIAM J. Sci. Comput., 43(4):A2923–A2948, 2021.
  • [65] R. E. Showalter. Diffusion in poro-elastic media. J. Math. Anal. Appl., 251(1):310–340, 2000.
  • [66] R. E. Showalter. Poroelastic filtration coupled to Stokes flow. In Control theory of partial differential equations, volume 242 of Lect. Notes Pure Appl. Math., pages 229–241. Chapman & Hall/CRC, Boca Raton, FL, 2005.
  • [67] R. E. Showalter and N. Su. Partially saturated flow in a poroelastic medium. Discrete Contin. Dyn. Syst. Ser. B, 1(4):403–420, 2001.
  • [68] S. Čanić. Fluid-structure interaction with incompressible fluids. In L. C. Berselli and M. Ružička, editors, Progress in Mathematical Fluid Dynamics, volume 2272 of Lecture Notes in Mathematics, pages 15–87. Springer, 2020.
  • [69] A. Ženíšek. The existence and uniqueness theorem in Biot’s consolidation theory. Aplikace Matematiky, 29(3):194–211, 1984.
  • [70] Y. Wang, S. Čanić, M. Bukač, C. Blaha, and S. Roy. Mathematical and computational modeling of a poroelastic cell scaffold in a bioartificial pancreas. Fluids, 7(7):222, 2022.
  • [71] J. Young, B. Rivière, Jr. C. S. Cox, and K. Uray. A mathematical model of intestinal oedema formation. Math. Med. Biol., 31(1):1–15, 2014.
  • [72] R. Zakerzadeh and P. Zunino. A computational framework for fluid-porous structure interaction with large structural deformation. Meccanica, 54:101–121, 2019.