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

Homogenization of a two-phase problem with nonlinear dynamic Wentzell-interface condition for connected-disconnected porous media

M. Gahn
Abstract

We investigate a reaction-diffusion problem in a two-component porous medium with a nonlinear interface condition between the different components. One component is connected and the other one is disconnected. The ratio between the microscopic pore scale and the size of the whole domain is described by the small parameter ϵ\epsilon. On the interface between the components we consider a dynamic Wentzell-boundary condition, where the normal fluxes from the bulk domains are given by a reaction-diffusion equation for the traces of the bulk-solutions, including nonlinear reaction-kinetics depending on the solutions on both sides of the interface. Using two-scale techniques, we pass to the limit ϵ0\epsilon\to 0 and derive macroscopic models, where we need homogenization results for surface diffusion. To cope with the nonlinear terms we derive strong two-scale convergence results.

Keywords: Homogenization; Two-scale convergence; Reaction-diffusion equations; Nonlinear interface conditions; Surface-diffusion.

MSC: 35K57; 35B27

footnotetext: Interdisciplinary Center for Scientific Computing, University of Heidelberg, Im Neuenheimer Feld 205, 69120 Heidelberg, Germany, [email protected].

1 Introduction

In this paper we derive homogenized models for nonlinear reaction-diffusion problems with dynamic Wentzell-boundary conditions in multi-component porous media. The domain consists of two components Ωϵ1\Omega_{\epsilon}^{1} and Ωϵ2\Omega_{\epsilon}^{2}, where Ωϵ1\Omega_{\epsilon}^{1} is connected, and Ωϵ2\Omega_{\epsilon}^{2} is disconnected and consists of periodically distributed inclusions. The small scaling parameter ϵ\epsilon represents the ratio between the length of an inclusion an the size of the whole domain. At the interface Γϵ\Gamma_{\epsilon} between the two components we assume a dynamic Wentzell-boundary condition, i. e., the normal flux at the surface is given by a reaction-diffusion equation on Γϵ\Gamma_{\epsilon}. More precisely, this boundary/interface condition describes processes like reactions, adsorption, desorption, and diffusion at the interface Γϵ\Gamma_{\epsilon}. Further it takes into account exchange of species between the different compartments, what can be modeled by nonlinear reaction-kinetics depending on the solutions on both sides of Γϵ\Gamma_{\epsilon}. The aim is the derivation of macroscopic models with homogenized diffusion coefficients for ϵ0\epsilon\to 0, the solution of which is an approximation of the microscopic solution. An additional focus of the paper is to provide general strong two-scale compactness results, which are based on a priori estimates for the microscopic solution.

Reaction-diffusion processes play an important role in many applications, and our model is motivated by metabolic and regulatory processes in living cells. Here, an important example is the carbohydrate metabolism in plant cells, where biochemical species are diffusing and reacting within the (connected) cytoplasm and the (disconnected) organelles (like chloroplasts and mitochondria), and are exchanged between different cellular compartments. At the outer mitochondrial membrane takes place the process of metabolic channeling, where intermediates in metabolic pathways are passed directly from enzyme to enzyme without equilibrating in the bulk-solution phase of the cell [32]. This effect can be modeled by the dynamic Wentzell-boundary condition, see [15, Chapter 4] for more details about the modeling and the derivation of these boundary conditions, which can be derived by asymptotic analysis.

To pass to the limit ϵ0\epsilon\to 0 in the variational equation for the microscopic problem we have to cope with several difficulties. The main challenges are the coupled bulk-surface diffusion in the perforated domains, as well as the treatment of the nonlinear terms, especially on the oscillating surface Γϵ\Gamma_{\epsilon}. To overcome these problems we make use of the two-scale method in perforated domains and on oscillating surfaces, where we need two-scale compactness results for diffusion processes on surfaces. To pass to the limit in the nonlinear terms we need strong convergence results. Such results are quite standard for the connected domain, but the usual methods fail for the disconnected domain. Here we make use of the unfolding method, which gives us a characterization for the two-scale convergence via functions defined on fixed domains, and a Kolmogorov-Simon-type compactness result for Banach-valued function spaces. Additionally, due to the nonlinear structure of the problem and the weak assumptions on the data, we have to deal with low regularity for the time-derivative.

There exists a large amount of papers dealing with homogenization problems for parabolic equations in multi-component porous media. However, results for the connected-disconnected case for nonlinear problems, especially for nonlinear interface conditions, seem to be rare. In [18] and [19] systems of reaction-diffusion problems are considered with nonlinear interface conditions. In [18] surface concentration is included and an additional focus lies on the modeling part of the carbohydrate metabolism and the specific structure of the nonlinear reaction kinetics. In the present paper we extend those models to problems including an additional surface diffusion for the traces of the bulk-solutions in Ωϵ1\Omega_{\epsilon}^{1} and Ωϵ2\Omega_{\epsilon}^{2}. The stationary case for different scalings with a continuous normal flux condition at the interface, given by a nonlinear monotone function depending on the jump of the solutions on both sides, is treated in [13]. There, the nonlinear terms in the disconnected domain only occurs for particular scalings and it is not straightforward to generalize those results to systems.

A double porosity model, where the diffusion inside the disconnected domain is of order ϵ2\epsilon^{2}, is considered in [8, 29] for continuous transmission conditions at the interface for the solutions and the normal fluxes. In [8] a nonlinear diffusion coefficient is considered, and the convergence of the nonlinear term is obtained by using the Kirchhoff-transformation and comparing the microscopic and the macroscopic equation, where the last one was obtained by a formal asymptotic expansion. Nonlinear reaction-kinetics in the bulk domains and an additional ordinary differential equation on the interface is considered in [29], where the strong convergence is proved by showing that the unfolded sequence of the microscopic solution is a Cauchy-sequence. A similar model with different kind of interface conditions is considered in [24], where the method of two-scale convergence is used and a variational principle to identify the limits of the nonlinear terms.

To pass to the limit in the diffusive terms on the interface Γϵ\Gamma_{\epsilon} arising from the Wentzell-boundary condition, compactness results for the surface gradient on an oscillating manifolds are needed. For such kind of problems in [4, 21] two-scale compactness results are derived for connected surfaces, where in [21] the method of unfolding is used. Compactness results for a coupled bulk-surface problem when the evolution of the trace of the bulk-solution on the surface Γϵ\Gamma_{\epsilon} is described by a diffusion equation, are treated in [5, 16]. In [5] continuity of the traces across the interface is assumed, where in [16] also jumps across the interface are allowed and also compactness results for the disconnected domain Ωϵ2\Omega_{\epsilon}^{2} are derived. In [5], the convergence results are applied to a linear problem with a dynamic Wentzell-interface condition. A reaction-diffusion problem including dynamic Wentzell-boundary conditions and nonlinear reaction-rates in the bulk domain and on the surface is considered in [6] for a connected perforated domain.

In this paper we start with the microscopic model and establish existence and uniqueness of a weak solution. The appropriate function space for a weak solution is the space of Sobolev functions of first order with H1H^{1}-traces on the interface Γϵ\Gamma_{\epsilon}, which we denote by j,ϵ\mathbb{H}_{j,\epsilon} for j=1,2j=1,2. To pass to the limit ϵ0\epsilon\to 0 we make use of the method of two-scale convergence for domains and surfaces, see [2, 3, 26, 28]. For the treatment of the diffusive terms on the oscillating surface we use the methods developed in [16] for the spaces j,ϵ\mathbb{H}_{j,\epsilon}. Those two-scale compactness results are based on a priori estimates for the microscopic solution depending explicitly on ϵ\epsilon. However, to pass to the limit in the nonlinear terms, the usual (weak) two-scale convergence is not enough and we need strong two-scale convergence, what leads to difficulties especially in the disconnected domain Ωϵ2\Omega_{\epsilon}^{2}. The strong convergence is obtained be applying the unfolding operator, see [11] for an overview of the unfolding method, to the microscopic solution and use a Kolmogorov-Simon-type compactness result for the unfolded sequence. We derive a general strong two-scale compactness result that is based only on a priori estimates and estimates for the difference between the solution and discrete shifts (with respect to the microscopic cells) of the solution. Since we only take into account linear shifts, which are not well defined for general surfaces, we use a Banach-valued Kolmogorov-Simon-compactness result, see [17]. Further, for our microscopic model we only obtain low regularity results for the time-derivative (which is a functional on j,ϵ\mathbb{H}_{j,\epsilon}), what leads to additional difficulties in the control of the time variable in the proof of the strong convergence.

The paper is organized as follows: In Section 2 we introduce the geometrical setting and the microscopic model. In Section 3 we show existence and uniqueness of a microscopic solution, and derive a priori estimates depending explicitly on ϵ\epsilon. In Section 4 we prove general strong two-scale compactness results for the connected and disconnected domain. In Section 5 we state the convergence results for the microscopic solution, formulate the macroscopic model, and show that the limit of the micro-solutions solves the macro-model. In the Appendix A we repeat the definition of the two-scale convergence and the unfolding operator and summarize some well known results from the literature.

2 The microscopic model

In this section we introduce the microscopic problem. We start with the definition of the microscopic domains Ωϵ1\Omega_{\epsilon}^{1} and Ωϵ2\Omega_{\epsilon}^{2}, as well as the interface Γϵ\Gamma_{\epsilon}, and explain some geometrical properties. Then we state the microscopic equation for given ϵ\epsilon and give the assumptions on the data.

2.1 The microscopic geometry

Let Ωn\Omega\subset\mathbb{R}^{n} with Lipschitz-boundary and ϵ>0\epsilon>0 a sequence with ϵ1\epsilon^{-1}\in\mathbb{N}. We define the unit cube Y:=(0,1)nY:=(0,1)^{n} and Y2YY_{2}\subset Y such that Y2¯Y\overline{Y_{2}}\subset Y, so Y2Y_{2} strictly included in YY. Further, we define Y1:=YY2¯Y_{1}:=Y\setminus\overline{Y_{2}} and Γ:=Y2\Gamma:=\partial Y_{2}, and we suppose that ΓC1,1\Gamma\in C^{1,1}. We assume that Y1Y_{1} is connected and for the sake of simplicity we also assume that Y2Y_{2} is connected. The general case of disconnected Y2Y_{2} is easily obtained by considering the connected components of Y2Y_{2}, see also Remark 3. Now, the microscopic domains Ωϵ1\Omega_{\epsilon}^{1} and Ωϵ2\Omega_{\epsilon}^{2} are defined by scaled and shifted reference elements YjY_{j} for j=1,2j=1,2. Let Kϵ:={kn:ϵ(k+Y)Ω}K_{\epsilon}:=\{k\in\mathbb{Z}^{n}\,:\,\epsilon(k+Y)\subset\Omega\} and define

Ωϵ2:=kKϵϵ(Y2+k),Ωϵ1:=ΩΩϵ2¯,Γϵ:=Ωϵ2.\displaystyle\Omega_{\epsilon}^{2}:=\bigcup_{k\in K_{\epsilon}}\epsilon(Y_{2}+k),\quad\quad\Omega_{\epsilon}^{1}:=\Omega\setminus\overline{\Omega_{\epsilon}^{2}},\quad\quad\Gamma_{\epsilon}:=\partial\Omega_{\epsilon}^{2}.

Hence, Γϵ\Gamma_{\epsilon} denotes the oscillating interface between Ωϵ1\Omega_{\epsilon}^{1} and Ωϵ2\Omega_{\epsilon}^{2}. Due to the assumptions on Y1Y_{1} and Y2Y_{2} it holds that Ωϵ1\Omega_{\epsilon}^{1} is connected and Ωϵ2\Omega_{\epsilon}^{2} is disconnected, and ΓϵC1,1\Gamma_{\epsilon}\in C^{1,1} is not touching the outer boundary Ω\partial\Omega.

2.2 The microscopic model

We are looking for a solution (uϵ1,uϵ2)(u_{\epsilon}^{1},u_{\epsilon}^{2}) with uϵj:(0,T)×Ωϵju_{\epsilon}^{j}:(0,T)\times\Omega_{\epsilon}^{j}\rightarrow\mathbb{R} for j=1,2j=1,2, such that it holds that

tuϵj(Dϵjuϵj)=fϵj(uϵj) in (0,T)×Ωϵj,ϵ(tuϵjΓϵ(DΓϵjΓϵuϵj)hϵj(uϵ1,uϵ2))=Dϵjuϵjν on (0,T)×Γϵ,Dϵjuϵjν=0 on (0,T)×Ω,uϵj(0)=uϵ,ij in Ωϵj,uϵj|Γϵ(0)=uϵ,i,Γϵj on Γϵ,\displaystyle\begin{aligned} \partial_{t}u_{\epsilon}^{j}-\nabla\cdot\big{(}D^{j}_{\epsilon}\nabla u_{\epsilon}^{j}\big{)}&=f_{\epsilon}^{j}(u_{\epsilon}^{j})&\mbox{ in }&(0,T)\times\Omega_{\epsilon}^{j},\\ \epsilon\big{(}\partial_{t}u_{\epsilon}^{j}-\nabla_{\Gamma_{\epsilon}}\cdot\big{(}D_{\Gamma_{\epsilon}}^{j}\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{j}\big{)}-h_{\epsilon}^{j}(u_{\epsilon}^{1},u_{\epsilon}^{2})\big{)}&=-D^{j}_{\epsilon}\nabla u_{\epsilon}^{j}\cdot\nu&\mbox{ on }&(0,T)\times\Gamma_{\epsilon},\\ -D^{j}_{\epsilon}\nabla u_{\epsilon}^{j}\cdot\nu&=0&\mbox{ on }&(0,T)\times\partial\Omega,\\ u_{\epsilon}^{j}(0)&=u_{\epsilon,i}^{j}&\mbox{ in }&\Omega_{\epsilon}^{j},\\ u_{\epsilon}^{j}|_{\Gamma_{\epsilon}}(0)&=u_{\epsilon,i,\Gamma_{\epsilon}}^{j}&\mbox{ on }&\Gamma_{\epsilon},\end{aligned} (1)

where ν\nu denotes the outer unit normal (we neglect a subscript for the underlying domain, since this should be clear from the context), and uϵj|Γϵu_{\epsilon}^{j}|_{\Gamma_{\epsilon}} denotes the trace of uϵju_{\epsilon}^{j} on Γϵ\Gamma_{\epsilon}. If it is clear from the context, we use the same notation for a function and its trace, for example we just write uϵju_{\epsilon}^{j} for uϵj|Γϵu_{\epsilon}^{j}|_{\Gamma_{\epsilon}}. The precise weak formulation of the micro-model above is stated in Section 3, see (3)\eqref{VariationalEquationMicroscopicProblem}, after introducing the necessary function spaces.

In the following, with TyΓT_{y}\Gamma and TxΓϵT_{x}\Gamma_{\epsilon} for yΓy\in\Gamma and xΓϵx\in\Gamma_{\epsilon} we denote the tangent spaces of Γ\Gamma at yy and Γϵ\Gamma_{\epsilon} at xx, respectively. The orthogonal projection PΓ(y):nTyΓP_{\Gamma}(y):\mathbb{R}^{n}\rightarrow T_{y}\Gamma for yΓy\in\Gamma is given by

PΓ(y)ξ=ξ(ξν(y))ν(y)for ξn,\displaystyle P_{\Gamma}(y)\xi=\xi-(\xi\cdot\nu(y))\nu(y)\quad\quad\mbox{for }\xi\in\mathbb{R}^{n},

where ν(y)\nu(y) denotes the outer unit normal at yΓy\in\Gamma. Let us extend the unit normal YY-periodically. Then, the orthogonal projection PΓϵ(x):nTxΓϵP_{\Gamma_{\epsilon}}(x):\mathbb{R}^{n}\rightarrow T_{x}\Gamma_{\epsilon} for xΓϵx\in\Gamma_{\epsilon} is given by

PΓϵ(x)ξ=ξ(ξν(xϵ))ν(xϵ)for ξn.\displaystyle P_{\Gamma_{\epsilon}}(x)\xi=\xi-\left(\xi\cdot\nu\left(\frac{x}{\epsilon}\right)\right)\nu\left(\frac{x}{\epsilon}\right)\quad\quad\mbox{for }\xi\in\mathbb{R}^{n}.

Assumptions on the data:
In the following let j{1,2}j\in\{1,2\}.

  1. (A1)

    For the bulk-diffusion we have Dϵj(x):=Dj(xϵ)D^{j}_{\epsilon}(x):=D^{j}\left(\frac{x}{\epsilon}\right) with DjLper(Yj)n×nD^{j}\in L^{\infty}_{\mathrm{per}}(Y_{j})^{n\times n} symmetric and coercive, i. e., there exits c0>0c_{0}>0 such that for almost every yYjy\in Y_{j}

    Dj(y)ξξc0|ξ|2 for all ξn.\displaystyle D^{j}(y)\xi\cdot\xi\geq c_{0}|\xi|^{2}\quad\mbox{ for all }\xi\in\mathbb{R}^{n}.
  2. (A2)

    For the surface-diffusion we suppose DΓϵj(x):=DΓj(xϵ)D_{\Gamma_{\epsilon}}^{j}(x):=D_{\Gamma}^{j}\left(\frac{x}{\epsilon}\right) with DΓjLper(Γ)n×nD_{\Gamma}^{j}\in L^{\infty}_{\mathrm{per}}(\Gamma)^{n\times n} symmetric and DΓj(y)|TyΓ:TyΓTyΓD_{\Gamma}^{j}(y)|_{T_{y}\Gamma}:T_{y}\Gamma\rightarrow T_{y}\Gamma for almost every yΓy\in\Gamma. Further, we assume that DΓjD_{\Gamma}^{j} is coercive, i. e., there exists c0>0c_{0}>0 such that for almost every yΓy\in\Gamma

    DΓj(y)ξξc0|ξ|2 for all ξTyΓ.\displaystyle D_{\Gamma}^{j}(y)\xi\cdot\xi\geq c_{0}|\xi|^{2}\quad\mbox{ for all }\xi\in T_{y}\Gamma.
  3. (A3)

    For the reaction-rates in the bulk domains we suppose that fϵj(t,x,z):=fj(t,xϵ,z)f_{\epsilon}^{j}(t,x,z):=f^{j}\left(t,\frac{x}{\epsilon},z\right) with fjL((0,T)×Yj×)f^{j}\in L^{\infty}((0,T)\times Y_{j}\times\mathbb{R}) is YY-periodic with respect to the second variable and uniformly Lipschitz continuous with respect to the last variable, i. e., there exists C>0C>0 such that for all z,wz,w\in\mathbb{R} and almost every (t,y)(0,T)×Yj(t,y)\in(0,T)\times Y_{j} it holds that

    |fj(t,y,z)fj(t,y,w)|C|zw|.\displaystyle|f^{j}(t,y,z)-f^{j}(t,y,w)|\leq C|z-w|.
  4. (A4)

    For the reaction-rates on the surface we suppose that hϵj(t,x,z1,z2):=hj(t,xϵ,z1,z2)h_{\epsilon}^{j}(t,x,z_{1},z_{2}):=h^{j}\left(t,\frac{x}{\epsilon},z_{1},z_{2}\right) with hjL((0,T)×Γ×2)h^{j}\in L^{\infty}((0,T)\times\Gamma\times\mathbb{R}^{2}) is YY-periodic with respect to the second variable and uniformly Lipschitz continuous with respect to the last variable, i. e., there exists C>0C>0 such that for all z1,z2,w1,w2z_{1},z_{2},w_{1},w_{2}\in\mathbb{R} and almost every (t,y)(0,T)×Γ(t,y)\in(0,T)\times\Gamma it holds that

    |hj(t,y,z1,z2)hj(t,y,w1,w2)|C(|z1w1|+|z2w2|).\displaystyle|h^{j}(t,y,z_{1},z_{2})-h^{j}(t,y,w_{1},w_{2})|\leq C\big{(}|z_{1}-w_{1}|+|z_{2}-w_{2}|\big{)}.
  5. (A5)

    For the initial conditions we assume uϵ,ijL2(Ωϵj)u_{\epsilon,i}^{j}\in L^{2}(\Omega_{\epsilon}^{j}) and uϵ,i,ΓϵjL2(Γϵ)u_{\epsilon,i,\Gamma_{\epsilon}}^{j}\in L^{2}(\Gamma_{\epsilon}) and there exists u0,ijL2(Ω)u_{0,i}^{j}\in L^{2}(\Omega) and u0,i,ΓjL2(Ω)u_{0,i,\Gamma}^{j}\in L^{2}(\Omega) such that

    uϵ,ij\displaystyle u_{\epsilon,i}^{j} u0,ij\displaystyle\rightarrow u_{0,i}^{j} in the two-scale sense ,
    uϵ,i,Γϵj\displaystyle u_{\epsilon,i,\Gamma_{\epsilon}}^{j} u0,i,Γj\displaystyle\rightarrow u_{0,i,\Gamma}^{j} in the two-scale sense on Γϵ\displaystyle\mbox{ in the two-scale sense on }\Gamma_{\epsilon} ,

    and it holds that

    ϵuϵ,i,ΓϵjL2(Γϵ)C.\displaystyle\sqrt{\epsilon}\|u_{\epsilon,i,\Gamma_{\epsilon}}^{j}\|_{L^{2}(\Gamma_{\epsilon})}\leq C.

    Additionally we assume that the sequences uϵ,i2u_{\epsilon,i}^{2} and uϵ,i,Γϵ2u_{\epsilon,i,\Gamma_{\epsilon}}^{2} converge strongly in the two-scale sense.

For the definition of the two-scale convergence see Section 4. We emphasize that due to the convergences in (A5) it holds that

uϵ,ijL2(Ωϵj)+ϵuϵ,i,ΓϵjL2(Γϵ)C.\displaystyle\|u_{\epsilon,i}^{j}\|_{L^{2}(\Omega_{\epsilon}^{j})}+\sqrt{\epsilon}\|u_{\epsilon,i,\Gamma_{\epsilon}}^{j}\|_{L^{2}(\Gamma_{\epsilon})}\leq C.

3 Existence of a weak solution and a priori estimates

The aim of this section is the investigation of the microscopic problem (1)\eqref{MicroscopicModel}. We introduce appropriate function spaces and show existence and uniqueness of a microscopic solution. Further, we derive a priori estimates for the solution depending explicitly on ϵ\epsilon. These estimates form the basis for the derivation of the macroscopic problem (12)\eqref{MacroscopicModel} by using the compactness results from Section 4.

3.1 Function spaces

Due to the Laplace-Beltrami operator in the boundary condition in (1)\eqref{MicroscopicModel}, it is not enough to consider the usual Sobolev space H1(Ωϵj)H^{1}(\Omega_{\epsilon}^{j}) as a solution space, because we need more regular traces. This gives rise to deal with the following function spaces for j=1,2j=1,2:

j,ϵ:={ϕϵjH1(Ωϵj):ϕϵj|ΓϵH1(Γϵ)},j:={ϕH1(Yj):ϕ|ΓH1(Γ)},\displaystyle\begin{aligned} \mathbb{H}_{j,\epsilon}&:=\left\{\phi_{\epsilon}^{j}\in H^{1}(\Omega_{\epsilon}^{j})\,:\,\phi_{\epsilon}^{j}|_{\Gamma_{\epsilon}}\in H^{1}(\Gamma_{\epsilon})\right\},\\ \mathbb{H}_{j}&:=\left\{\phi\in H^{1}(Y_{j})\,:\,\phi|_{\Gamma}\in H^{1}(\Gamma)\right\},\end{aligned} (2)

with the inner products

(ϕϵj,ψϵj)j,ϵ\displaystyle(\phi_{\epsilon}^{j},\psi_{\epsilon}^{j})_{\mathbb{H}_{j,\epsilon}} :=(ϕϵj,ψϵj)H1(Ωϵj)+ϵ(ϕϵj,ψϵj)H1(Γϵ),\displaystyle:=(\phi_{\epsilon}^{j},\psi_{\epsilon}^{j})_{H^{1}(\Omega_{\epsilon}^{j})}+\epsilon(\phi_{\epsilon}^{j},\psi_{\epsilon}^{j})_{H^{1}(\Gamma_{\epsilon})},
(ϕ,ψ)j\displaystyle(\phi,\psi)_{\mathbb{H}_{j}} :=(ϕ,ψ)H1(Yj)+(ϕ,ψ)H1(Γ).\displaystyle:=(\phi,\psi)_{H^{1}(Y_{j})}+(\phi,\psi)_{H^{1}(\Gamma)}.

The associated norms are denoted by j,ϵ\|\cdot\|_{\mathbb{H}_{j,\epsilon}} and j\|\cdot\|_{\mathbb{H}_{j}}. Obviously, the spaces j,ϵ\mathbb{H}_{j,\epsilon} and j\mathbb{H}_{j} are separable Hilbert spaces and we have the dense embeddings C(Ωϵj¯)j,ϵC^{\infty}(\overline{\Omega_{\epsilon}^{j}})\subset\mathbb{H}_{j,\epsilon} and C(Yj¯)jC^{\infty}(\overline{Y_{j}})\subset\mathbb{H}_{j}, see [16, Proposition 5]. We also define the space

𝕃j,ϵ:=L2(Ωϵj)×L2(Γϵ),𝕃j:=L2(Yj)×L2(Γ)\displaystyle\mathbb{L}_{j,\epsilon}:=L^{2}(\Omega_{\epsilon}^{j})\times L^{2}(\Gamma_{\epsilon}),\quad\quad\mathbb{L}_{j}:=L^{2}(Y_{j})\times L^{2}(\Gamma)

with inner products

(ϕϵj,ψϵj)𝕃j,ϵ\displaystyle(\phi_{\epsilon}^{j},\psi_{\epsilon}^{j})_{\mathbb{L}_{j,\epsilon}} :=(ϕϵj,ψϵj)L2(Ωϵj)+ϵ(ϕϵj,ψϵj)L2(Γϵ),\displaystyle:=(\phi_{\epsilon}^{j},\psi_{\epsilon}^{j})_{L^{2}(\Omega_{\epsilon}^{j})}+\epsilon(\phi_{\epsilon}^{j},\psi_{\epsilon}^{j})_{L^{2}(\Gamma_{\epsilon})},
(ϕ,ψ)𝕃j\displaystyle(\phi,\psi)_{\mathbb{L}_{j}} :=(ϕ,ψ)L2(Yj)+(ϕ,ψ)L2(Γ),\displaystyle:=(\phi,\psi)_{L^{2}(Y_{j})}+(\phi,\psi)_{L^{2}(\Gamma)},

and again denote the associated norms by 𝕃j,ϵ\|\cdot\|_{\mathbb{L}_{j,\epsilon}} and 𝕃j\|\cdot\|_{\mathbb{L}_{j}}. Obviously, we have

j,ϵ={(uϵ,vϵ)H1(Ωϵj)×H1(Γϵ):uϵ|Γϵ=vϵ},\displaystyle\mathbb{H}_{j,\epsilon}\overset{\sim}{=}\left\{(u_{\epsilon},v_{\epsilon})\in H^{1}(\Omega_{\epsilon}^{j})\times H^{1}(\Gamma_{\epsilon})\,:\,u_{\epsilon}|_{\Gamma_{\epsilon}}=v_{\epsilon}\right\},

and a similar result for j\mathbb{H}_{j}. Therefore, we have the following Gelfand-triples:

j,ϵ𝕃j,ϵj,ϵ,j𝕃jj.\displaystyle\mathbb{H}_{j,\epsilon}\hookrightarrow\mathbb{L}_{j,\epsilon}\hookrightarrow\mathbb{H}_{j,\epsilon}^{\prime},\quad\quad\mathbb{H}_{j}\hookrightarrow\mathbb{L}_{j}\hookrightarrow\mathbb{H}_{j}.

We will also make use for α(12,1]\alpha\in\left(\frac{1}{2},1\right] of the function space

jα:={ϕHα(Yj):ϕ|ΓHα(Γ)}\displaystyle\mathbb{H}_{j}^{\alpha}:=\left\{\phi\in H^{\alpha}(Y_{j})\,:\,\phi|_{\Gamma}\in H^{\alpha}(\Gamma)\right\}

with inner product

(ϕ,ψ)jα:=(ϕ,ψ)Hα(Yj)+(ϕ,ψ)Hα(Γ).\displaystyle(\phi,\psi)_{\mathbb{H}_{j}^{\alpha}}:=(\phi,\psi)_{H^{\alpha}(Y_{j})}+(\phi,\psi)_{H^{\alpha}(\Gamma)}.

By definition we have j=j1\mathbb{H}_{j}=\mathbb{H}_{j}^{1}. Obviously, we have the compact embedding jjα\mathbb{H}_{j}\hookrightarrow\mathbb{H}_{j}^{\alpha} for α(12,1)\alpha\in\left(\frac{1}{2},1\right).

3.2 Existence and uniqueness of a weak solution

A weak solution of Problem (1)\eqref{MicroscopicModel} is defined in the following way: The tuple (uϵ1,uϵ2)(u_{\epsilon}^{1},u_{\epsilon}^{2}) is a weak solution of (1)\eqref{MicroscopicModel} if for j=1,2j=1,2

uϵjL2((0,T),j,ϵ)H1((0,T),j,ϵ),\displaystyle u_{\epsilon}^{j}\in L^{2}((0,T),\mathbb{H}_{j,\epsilon})\cap H^{1}((0,T),\mathbb{H}_{j,\epsilon}^{\prime}),

uϵju_{\epsilon}^{j} and uϵj|Γϵu_{\epsilon}^{j}|_{\Gamma_{\epsilon}} fulfill the initial condition uϵj(0)=uϵ,iju_{\epsilon}^{j}(0)=u_{\epsilon,i}^{j} and uϵj|Γϵ(0)=uϵ,i,Γϵju_{\epsilon}^{j}|_{\Gamma_{\epsilon}}(0)=u_{\epsilon,i,\Gamma_{\epsilon}}^{j}, and for all ϕϵjj,ϵ\phi_{\epsilon}^{j}\in\mathbb{H}_{j,\epsilon} it holds almost everywhere in (0,T)(0,T)

tuϵj,ϕϵjj,ϵ,j,ϵ+(Dϵjuϵj,ϕϵj)Ωϵj+ϵ(DΓϵjΓϵuϵj,Γϵϕϵj)Γϵ=(fϵj(uϵj),ϕϵj)Ωϵj+ϵ(hϵj(uϵ1,uϵ2),ϕϵj)Γϵ.\displaystyle\begin{aligned} \langle\partial_{t}u_{\epsilon}^{j},\phi_{\epsilon}^{j}&\rangle_{\mathbb{H}_{j,\epsilon}^{\prime},\mathbb{H}_{j,\epsilon}}+(D^{j}_{\epsilon}\nabla u_{\epsilon}^{j},\nabla\phi_{\epsilon}^{j})_{\Omega_{\epsilon}^{j}}+\epsilon(D_{\Gamma_{\epsilon}}^{j}\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{j},\nabla_{\Gamma_{\epsilon}}\phi_{\epsilon}^{j})_{\Gamma_{\epsilon}}\\ &=(f_{\epsilon}^{j}(u_{\epsilon}^{j}),\phi_{\epsilon}^{j})_{\Omega_{\epsilon}^{j}}+\epsilon(h_{\epsilon}^{j}(u_{\epsilon}^{1},u_{\epsilon}^{2}),\phi_{\epsilon}^{j})_{\Gamma_{\epsilon}}.\end{aligned} (3)

Here (,)U(\cdot,\cdot)_{U} stands for the inner product on L2(U)L^{2}(U), for a suitable set UnU\subset\mathbb{R}^{n}, and for a Banach space XX and its dual XX^{\prime} we write ,X,X\langle\cdot,\cdot\rangle_{X^{\prime},X} for the duality pairing between XX^{\prime} and XX. The scaling factor ϵ\epsilon for the time-derivative in (1)\eqref{MicroscopicModel} is included in the duality pairing ,j,ϵ,j,ϵ\langle\cdot,\cdot\rangle_{\mathbb{H}_{j,\epsilon}^{\prime},\mathbb{H}_{j,\epsilon}}. In fact, if additionally it holds that tuϵjL2((0,T),H1(Ωϵj))\partial_{t}u_{\epsilon}^{j}\in L^{2}((0,T),H^{1}(\Omega_{\epsilon}^{j})^{\prime}) and tuϵj|ΓϵL2((0,T),H1(Γϵ))\partial_{t}u_{\epsilon}^{j}|_{\Gamma_{\epsilon}}\in L^{2}((0,T),H^{1}(\Gamma_{\epsilon})^{\prime}) with respect to the Gelfand-triples H1(Ωϵj)L2(Ωϵj)H1(Ωϵj)H^{1}(\Omega_{\epsilon}^{j})\hookrightarrow L^{2}(\Omega_{\epsilon}^{j})\hookrightarrow H^{1}(\Omega_{\epsilon}^{j})^{\prime} and H1(Γϵ)L2(Γϵ)H1(Γϵ)H^{1}(\Gamma_{\epsilon})\hookrightarrow L^{2}(\Gamma_{\epsilon})\hookrightarrow H^{1}(\Gamma_{\epsilon})^{\prime}, we get for all ϕϵjj,ϵ\phi_{\epsilon}^{j}\in\mathbb{H}_{j,\epsilon}

tuϵj,ϕϵjj,ϵ,j,ϵ=tuϵj,ϕϵjH1(Ωϵj),H1(Ωϵj)+ϵtuϵj|Γϵ,ϕϵj|ΓϵH1(Γϵ),H1(Γϵ).\displaystyle\langle\partial_{t}u_{\epsilon}^{j},\phi_{\epsilon}^{j}\rangle_{\mathbb{H}_{j,\epsilon}^{\prime},\mathbb{H}_{j,\epsilon}}=\langle\partial_{t}u_{\epsilon}^{j},\phi_{\epsilon}^{j}\rangle_{H^{1}(\Omega_{\epsilon}^{j})^{\prime},H^{1}(\Omega_{\epsilon}^{j})}+\epsilon\langle\partial_{t}u_{\epsilon}^{j}|_{\Gamma_{\epsilon}},\phi_{\epsilon}^{j}|_{\Gamma_{\epsilon}}\rangle_{H^{1}(\Gamma_{\epsilon})^{\prime},H^{1}(\Gamma_{\epsilon})}.
Proposition 1.

There exists a unique weak solution uϵ=(uϵ1,uϵ2)u_{\epsilon}=(u_{\epsilon}^{1},u_{\epsilon}^{2}) of the microscopic problem (1)\eqref{MicroscopicModel}.

Proof.

This is an easy consequence of the Galerkin-method and the Leray-Schauder principle, where we have to use similar estimates as in Proposition 2 below. The uniqueness follows from standard energy estimates. For more details see [15]. ∎

3.3 A priori estimates

We derive a priori estimates for the microscopic solution depending explicitly on ϵ\epsilon. These estimates are necessary for the application of the two-scale compactness results from Section 4 to derive the macroscopic model. In a first step, we give estimates in the spaces L2((0,T),j,ϵ)L^{2}((0,T),\mathbb{H}_{j,\epsilon}) and H1((0,T),j,ϵ)H^{1}((0,T),\mathbb{H}_{j,\epsilon}^{\prime}). Such kind of estimates are also needed to establish the existence of a weak solution via the Galerkin-method. In a second step, we derive estimates for the difference of shifted microscopic solution with respect to the macroscopic variable. These estimates are necessary for strong two-scale compactness results in the disconnected domain.

The following trace inequality for perforated domains will be used frequently throughout the paper and follows easily by a standard decomposition argument and the trace inequality on the reference element YjY_{j}, see also [20, Theorem II.4.1 and Exercise II.4.1]: For every θ>0\theta>0 there exists a C(θ)>0C(\theta)>0 such that for every vϵH1(Ωϵj)v_{\epsilon}\in H^{1}(\Omega_{\epsilon}^{j}) it holds that

vϵL2(Γϵ)C(θ)ϵvϵL2(Ωϵj)+θϵvϵL2(Ωϵj).\displaystyle\|v_{\epsilon}\|_{L^{2}(\Gamma_{\epsilon})}\leq\frac{C(\theta)}{\sqrt{\epsilon}}\|v_{\epsilon}\|_{L^{2}(\Omega_{\epsilon}^{j})}+\theta\sqrt{\epsilon}\|\nabla v_{\epsilon}\|_{L^{2}(\Omega_{\epsilon}^{j})}. (4)
Proposition 2.

The weak solution uϵ=(uϵ1,uϵ2)u_{\epsilon}=(u_{\epsilon}^{1},u_{\epsilon}^{2}) of the microscopic problem (1)\eqref{MicroscopicModel} fulfills the following a priori estimate

tuϵjL2((0,T),j,ϵ)+uϵjL2((0,T),j,ϵ)C,\displaystyle\|\partial_{t}u_{\epsilon}^{j}\|_{L^{2}((0,T),\mathbb{H}_{j,\epsilon}^{\prime})}+\|u_{\epsilon}^{j}\|_{L^{2}((0,T),\mathbb{H}_{j,\epsilon})}\leq C,

for a constant C>0C>0 independent of ϵ\epsilon.

Proof.

We choose uϵju_{\epsilon}^{j} as a test-function in (3)\eqref{SectionExistence} (for j=1,2j=1,2) to obtain with the Assumptions (A3) and (A4) on fjf^{j} and hjh^{j}

12ddtuϵj𝕃j,ϵ2+(Dϵj\displaystyle\frac{1}{2}\frac{d}{dt}\|u_{\epsilon}^{j}\|_{\mathbb{L}_{j,\epsilon}}^{2}+(D^{j}_{\epsilon} uϵj,uϵj)Ωϵj+ϵ(DΓϵjΓϵuϵj,Γϵuϵj)Γϵ\displaystyle\nabla u_{\epsilon}^{j},\nabla u_{\epsilon}^{j})_{\Omega_{\epsilon}^{j}}+\epsilon(D_{\Gamma_{\epsilon}}^{j}\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{j},\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{j})_{\Gamma_{\epsilon}}
=(fϵj(uϵj),uϵj)Ωϵj+ϵ(hϵj(uϵ1,uϵ2),uϵj)Γϵ\displaystyle=(f_{\epsilon}^{j}(u_{\epsilon}^{j}),u_{\epsilon}^{j})_{\Omega_{\epsilon}^{j}}+\epsilon(h_{\epsilon}^{j}(u_{\epsilon}^{1},u_{\epsilon}^{2}),u_{\epsilon}^{j})_{\Gamma_{\epsilon}}
C(1+uϵjL2(Ωϵj)2+ϵuϵ1L2(Γϵ)2+ϵuϵ2L2(Γϵ)2)\displaystyle\leq C\left(1+\|u_{\epsilon}^{j}\|_{L^{2}(\Omega_{\epsilon}^{j})}^{2}+\epsilon\|u_{\epsilon}^{1}\|^{2}_{L^{2}(\Gamma_{\epsilon})}+\epsilon\|u_{\epsilon}^{2}\|^{2}_{L^{2}(\Gamma_{\epsilon})}\right)
C(1+uϵ1𝕃1,ϵ2+uϵ2𝕃2,ϵ2).\displaystyle\leq C\left(1+\|u_{\epsilon}^{1}\|^{2}_{\mathbb{L}_{1,\epsilon}}+\|u_{\epsilon}^{2}\|^{2}_{\mathbb{L}_{2,\epsilon}}\right).

Using the coercivity of DϵjD^{j}_{\epsilon} and DΓϵjD_{\Gamma_{\epsilon}}^{j} from the Assumptions (A1) and (A2), we obtain for j=1,2j=1,2

ddtuϵj𝕃j,ϵ2+uϵjL2(Ωϵj)2+ϵΓϵuϵjL2(Γϵ)2C(1+uϵ1𝕃1,ϵ2+uϵ2𝕃2,ϵ2).\displaystyle\frac{d}{dt}\|u_{\epsilon}^{j}\|_{\mathbb{L}_{j,\epsilon}}^{2}+\|\nabla u_{\epsilon}^{j}\|_{L^{2}(\Omega_{\epsilon}^{j})}^{2}+\epsilon\|\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{j}\|_{L^{2}(\Gamma_{\epsilon})}^{2}\leq C\left(1+\|u_{\epsilon}^{1}\|^{2}_{\mathbb{L}_{1,\epsilon}}+\|u_{\epsilon}^{2}\|^{2}_{\mathbb{L}_{2,\epsilon}}\right).

Summing over j=1,2j=1,2, integrating with respect to time, the Assumption (A5), and the Gronwall-inequality implies the boundedness of uϵjL2((0,T),j,ϵ)\|u_{\epsilon}^{j}\|_{L^{2}((0,T),\mathbb{H}_{j,\epsilon})} uniformly with respect to ϵ\epsilon.

It remains to check the bound for the time-derivative tuϵj\partial_{t}u_{\epsilon}^{j}. As a test-function in (3)\eqref{VariationalEquationMicroscopicProblem} we choose ϕϵjj,ϵ\phi_{\epsilon}^{j}\in\mathbb{H}_{j,\epsilon} with ϕϵjj,ϵ1\|\phi_{\epsilon}^{j}\|_{\mathbb{H}_{j,\epsilon}}\leq 1 to obtain (using the boundedness of the diffusion tensors and again the growth condition for hjh^{j} and fjf^{j}):

tuϵj,ϕϵjj,ϵ,j,ϵC(uϵ11,ϵ+uϵ22,ϵ)ϕϵjj,ϵC(uϵ11,ϵ+uϵ22,ϵ).\displaystyle\langle\partial_{t}u_{\epsilon}^{j},\phi_{\epsilon}^{j}\rangle_{\mathbb{H}_{j,\epsilon}^{\prime},\mathbb{H}_{j,\epsilon}}\leq C\left(\|u_{\epsilon}^{1}\|_{\mathbb{H}_{1,\epsilon}}+\|u_{\epsilon}^{2}\|_{\mathbb{H}_{2,\epsilon}}\right)\|\phi_{\epsilon}^{j}\|_{\mathbb{H}_{j,\epsilon}}\leq C\left(\|u_{\epsilon}^{1}\|_{\mathbb{H}_{1,\epsilon}}+\|u_{\epsilon}^{2}\|_{\mathbb{H}_{2,\epsilon}}\right).

Squaring, integrating with respect to time, and the boundedness of uϵju_{\epsilon}^{j} for j=1,2j=1,2 already obtained above implies the desired result. ∎

Next, we derive estimates for the difference of the shifted functions. First of all, we introduce some additional notations. For h>0h>0 let us define

Ωh\displaystyle\Omega_{h} :={xΩ:dist(x,Ω)>h},\displaystyle:=\{x\in\Omega\,:\,\mathrm{dist}(x,\partial\Omega)>h\},
Kϵ,h\displaystyle K_{\epsilon,h} :={kn:ϵ(Y+k)Ωh},\displaystyle:=\{k\in\mathbb{Z}^{n}\,:\,\epsilon(Y+k)\subset\Omega_{h}\},
Ωϵ,h\displaystyle\Omega_{\epsilon,h} :=intkKϵ,hϵ(Y¯+k),\displaystyle:=\mathrm{int}\bigcup_{k\in K_{\epsilon,h}}\epsilon\big{(}\overline{Y}+k\big{)},

and the related perforated domains and the related surface

Ωϵ,h2\displaystyle\Omega_{\epsilon,h}^{2} :=kKϵ,hϵ(Y2+k),Ωϵ,h1:=Ωϵ,hΩϵ,h2¯,Γϵ,h:=Ωϵ,h2.\displaystyle:=\bigcup_{k\in K_{\epsilon,h}}\epsilon\big{(}Y_{2}+k\big{)},\quad\Omega_{\epsilon,h}^{1}:=\Omega_{\epsilon,h}\setminus\overline{\Omega_{\epsilon,h}^{2}},\quad\Gamma_{\epsilon,h}:=\partial\Omega_{\epsilon,h}^{2}.

For lnl\in\mathbb{Z}^{n} with |lϵ|<h|l\epsilon|<h and Gϵ,h{Ωϵ,h,Ωϵ,h1,Ωϵ,h2}G_{\epsilon,h}\in\{\Omega_{\epsilon,h},\Omega_{\epsilon,h}^{1},\Omega_{\epsilon,h}^{2}\}, we define for an arbitrary function vϵ:Gϵ,hv_{\epsilon}:G_{\epsilon,h}\rightarrow\mathbb{R} the shifted function

vϵl(x):=vϵ(x+lϵ),\displaystyle v_{\epsilon}^{l}(x):=v_{\epsilon}(x+l\epsilon),

and the difference between the shifted function and the function itself

δlvϵ(x):=δvϵ(x):=vϵl(x)vϵ(x)=vϵ(x+lϵ)vϵ(x).\displaystyle\delta_{l}v_{\epsilon}(x):=\delta v_{\epsilon}(x):=v_{\epsilon}^{l}(x)-v_{\epsilon}(x)=v_{\epsilon}(x+l\epsilon)-v_{\epsilon}(x). (5)

Here, in the writing δvϵ\delta v_{\epsilon} we neglect the dependence on ll if it is clear from the context. Further, we define j,ϵ,h\mathbb{H}_{j,\epsilon,h} in the same way as j,ϵ\mathbb{H}_{j,\epsilon} in (2)\eqref{DefinitionHilbertraum} by replacing Ωϵj\Omega_{\epsilon}^{j} and Γϵ\Gamma_{\epsilon} with Ωϵ,hj\Omega_{\epsilon,h}^{j} and Γϵ,h\Gamma_{\epsilon,h}. In the same way we define 𝕃j,ϵ,h\mathbb{L}_{j,\epsilon,h}. Further, for any function ϕϵ,h2,ϵ,h\phi_{\epsilon,h}\in\mathbb{H}_{2,\epsilon,h} we write ϕ¯ϵ,h\overline{\phi}_{\epsilon,h} for the zero extension to Ωϵ2\Omega_{\epsilon}^{2}. Especially it holds that ϕ¯ϵ,h2,ϵ\overline{\phi}_{\epsilon,h}\in\mathbb{H}_{2,\epsilon}, since Ωϵ2\Omega_{\epsilon}^{2} is disconnected.

Proposition 3.

Let 0<h10<h\ll 1, then for all lnl\in\mathbb{Z}^{n} with |ϵl|<h|\epsilon l|<h, it holds that

δuϵ2L((0,T),𝕃2,ϵ,h)+\displaystyle\|\delta u_{\epsilon}^{2}\|_{L^{\infty}((0,T),\mathbb{L}_{2,\epsilon,h})}+\| δuϵ2L2((0,T),𝕃2,ϵ,h)\displaystyle\nabla\delta u_{\epsilon}^{2}\|_{L^{2}((0,T),\mathbb{L}_{2,\epsilon,h})}
C(δuϵ1L2((0,T),L2(Ωϵ,h1))+δ(uϵ,i2,uϵ,i,Γϵ2)𝕃2,ϵ,h+ϵ),\displaystyle\leq C\left(\|\delta u_{\epsilon}^{1}\|_{L^{2}((0,T),L^{2}(\Omega_{\epsilon,h}^{1}))}+\big{\|}\delta\big{(}u_{\epsilon,i}^{2},u_{\epsilon,i,\Gamma_{\epsilon}}^{2}\big{)}\big{\|}_{\mathbb{L}_{2,\epsilon,h}}+\epsilon\right),

for a constant C>0C>0 independent of h,ϵ,h,\,\epsilon, and ll.

Proof.

Let 0<h10<h\ll 1 and lnl\in\mathbb{Z}^{n} with |ϵl|<h|\epsilon l|<h, and we shortly write uϵ2,l:=(uϵ2|Ωϵ,h2)lu_{\epsilon}^{2,l}:=\left(u_{\epsilon}^{2}|_{\Omega_{\epsilon,h}^{2}}\right)^{l}, i. e., the shifts with respect to lϵl\epsilon of the restriction uϵ2|Ωϵ,h2u_{\epsilon}^{2}|_{\Omega_{\epsilon,h}^{2}} (we neglect the index hh). In the same way we define uϵ1,lu_{\epsilon}^{1,l}. Let ϕϵ,h2,ϵ,h\phi_{\epsilon,h}\in\mathbb{H}_{2,\epsilon,h}. Then, for xΩϵ2(Ωϵ,h2+ϵl)x\in\Omega_{\epsilon}^{2}\setminus(\Omega_{\epsilon,h}^{2}+\epsilon l) it holds that xlϵΩϵ,h2x-l\epsilon\notin\Omega_{\epsilon,h}^{2} and therefore ϕ¯ϵ,hl(x)=0\overline{\phi}_{\epsilon,h}^{-l}(x)=0 and similar from xΓϵ(Γϵ,h+ϵl)x\in\Gamma_{\epsilon}\setminus(\Gamma_{\epsilon,h}+\epsilon l) it follows ϕ¯ϵ,hl(x)=0\overline{\phi}_{\epsilon,h}^{-l}(x)=0 . This implies for all ψC0(0,T)\psi\in C_{0}^{\infty}(0,T)

0T\displaystyle\int_{0}^{T} (uϵ2,l,ϕϵ,h)𝕃2,ϵ,hψ(t)dt\displaystyle\left(u_{\epsilon}^{2,l},\phi_{\epsilon,h}\right)_{\mathbb{L}_{2,\epsilon,h}}\psi^{\prime}(t)dt
=0T[Ωϵ,h2uϵ2(t,x+lϵ)ϕϵ,h(x)𝑑x+ϵΓϵ,huϵ2(t,x+lϵ)ϕϵ,h(x)𝑑σ]ψ(t)𝑑t\displaystyle=\int_{0}^{T}\left[\int_{\Omega_{\epsilon,h}^{2}}u_{\epsilon}^{2}(t,x+l\epsilon)\phi_{\epsilon,h}(x)dx+\epsilon\int_{\Gamma_{\epsilon,h}}u_{\epsilon}^{2}(t,x+l\epsilon)\phi_{\epsilon,h}(x)d\sigma\right]\psi^{\prime}(t)dt
=0T[Ωϵ,h2+lϵuϵ2(t,x)ϕϵ,hl(x)𝑑x+ϵΓϵ,h+lϵuϵ2(t,x)ϕϵ,hl(x)𝑑σ]ψ(t)𝑑t\displaystyle=\int_{0}^{T}\left[\int_{\Omega_{\epsilon,h}^{2}+l\epsilon}u_{\epsilon}^{2}(t,x)\phi_{\epsilon,h}^{-l}(x)dx+\epsilon\int_{\Gamma_{\epsilon,h}+l\epsilon}u_{\epsilon}^{2}(t,x)\phi_{\epsilon,h}^{-l}(x)d\sigma\right]\psi^{\prime}(t)dt
=0T[Ωϵ2uϵ2(t,x)ϕϵ,hl(x)𝑑x+ϵΓϵuϵ2(t,x)ϕϵ,hl(x)𝑑σ]ψ(t)𝑑t\displaystyle=\int_{0}^{T}\left[\int_{\Omega_{\epsilon}^{2}}u_{\epsilon}^{2}(t,x)\phi_{\epsilon,h}^{-l}(x)dx+\epsilon\int_{\Gamma_{\epsilon}}u_{\epsilon}^{2}(t,x)\phi_{\epsilon,h}^{-l}(x)d\sigma\right]\psi^{\prime}(t)dt
=0T(uϵ2,ϕ¯ϵ,hl)𝕃2,ϵψ(t)𝑑t=0Ttuϵ2,ϕ¯ϵ,hl2,ϵ,2,ϵψ(t)𝑑t.\displaystyle=\int_{0}^{T}\left(u_{\epsilon}^{2},\overline{\phi}_{\epsilon,h}^{-l}\right)_{\mathbb{L}_{2,\epsilon}}\psi^{\prime}(t)dt=-\int_{0}^{T}\left\langle\partial_{t}u_{\epsilon}^{2},\overline{\phi}_{\epsilon,h}^{-l}\right\rangle_{\mathbb{H}_{2,\epsilon}^{\prime},\mathbb{H}_{2,\epsilon}}\psi(t)dt.

Hence, we have tuϵ2,lL2((0,T),2,ϵ,h)\partial_{t}u_{\epsilon}^{2,l}\in L^{2}((0,T),\mathbb{H}_{2,\epsilon,h}^{\prime}) with

tuϵ2,l,ϕϵ,h2,ϵ,h,2,ϵ,h=tuϵ2,ϕ¯ϵ,hl2,ϵ,2,ϵ\displaystyle\big{\langle}\partial_{t}u_{\epsilon}^{2,l},\phi_{\epsilon,h}\big{\rangle}_{\mathbb{H}_{2,\epsilon,h}^{\prime},\mathbb{H}_{2,\epsilon,h}}=\big{\langle}\partial_{t}u_{\epsilon}^{2},\overline{\phi}_{\epsilon,h}^{-l}\big{\rangle}_{\mathbb{H}_{2,\epsilon}^{\prime},\mathbb{H}_{2,\epsilon}}

almost everywhere in (0,T)(0,T). Using ϕ¯ϵ,hl2,ϵ\overline{\phi}_{\epsilon,h}^{-l}\in\mathbb{H}_{2,\epsilon} as a test-function in (3)\eqref{VariationalEquationMicroscopicProblem}, we obtain using the periodicity of D2D^{2}, DΓ2D^{2}_{\Gamma}, f2f^{2}, and h2h^{2}, by an elemental calculation

tuϵ2,ϕ¯ϵ,hl2,ϵ,2,ϵ=\displaystyle\big{\langle}\partial_{t}u_{\epsilon}^{2},\overline{\phi}_{\epsilon,h}^{-l}\big{\rangle}_{\mathbb{H}_{2,\epsilon}^{\prime},\mathbb{H}_{2,\epsilon}}= (Dϵ2uϵ2,l,ϕϵ,h)Ωϵ,h2ϵ(DΓϵ2Γϵuϵ2,l,Γϵϕϵ,h)Γϵ,h\displaystyle-\big{(}D_{\epsilon}^{2}\nabla u_{\epsilon}^{2,l},\nabla\phi_{\epsilon,h}\big{)}_{\Omega_{\epsilon,h}^{2}}-\epsilon\big{(}D_{\Gamma_{\epsilon}}^{2}\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{2,l},\nabla_{\Gamma_{\epsilon}}\phi_{\epsilon,h}\big{)}_{\Gamma_{\epsilon,h}}
+(fϵ2(uϵ2,l),ϕϵ,h)Ωϵ,h2+ϵ(hϵ2(uϵ1,l,uϵ2,l),ϕϵ,h)Γϵ,h.\displaystyle+\big{(}f_{\epsilon}^{2}(u_{\epsilon}^{2,l}),\phi_{\epsilon,h}\big{)}_{\Omega_{\epsilon,h}^{2}}+\epsilon\big{(}h_{\epsilon}^{2}(u_{\epsilon}^{1,l},u_{\epsilon}^{2,l}),\phi_{\epsilon,h}\big{)}_{\Gamma_{\epsilon,h}}.

Subtracting the above equation for l=0l=0 and arbitrary lnl\in\mathbb{Z}^{n} with |ϵl|<h|\epsilon l|<h we obtain

t\displaystyle\langle\partial_{t} δuϵ2,ϕϵ,h2,ϵ,h,2,ϵ,h+(Dϵ2δuϵ2,ϕϵ,h)Ωϵ,h2+ϵ(DΓϵ2Γϵδuϵ2,Γϵϕϵ,h)Γϵ,h\displaystyle\delta u_{\epsilon}^{2},\phi_{\epsilon,h}\rangle_{\mathbb{H}_{2,\epsilon,h}^{\prime},\mathbb{H}_{2,\epsilon,h}}+\big{(}D_{\epsilon}^{2}\nabla\delta u_{\epsilon}^{2},\nabla\phi_{\epsilon,h}\big{)}_{\Omega_{\epsilon,h}^{2}}+\epsilon\big{(}D_{\Gamma_{\epsilon}}^{2}\nabla_{\Gamma_{\epsilon}}\delta u_{\epsilon}^{2},\nabla_{\Gamma_{\epsilon}}\phi_{\epsilon,h}\big{)}_{\Gamma_{\epsilon,h}}
=(fϵ2((uϵ2)l)fϵ2(uϵ2),ϕϵ,h)Ωϵ,h2+ϵ(hϵ2((uϵ1)l,(uϵ2)l)hϵ2(uϵ1,uϵ2),ϕϵ,h)Γϵ,h.\displaystyle=\big{(}f_{\epsilon}^{2}\big{(}(u_{\epsilon}^{2})^{l}\big{)}-f_{\epsilon}^{2}(u_{\epsilon}^{2}),\phi_{\epsilon,h}\big{)}_{\Omega_{\epsilon,h}^{2}}+\epsilon\big{(}h_{\epsilon}^{2}\big{(}(u_{\epsilon}^{1})^{l},(u_{\epsilon}^{2})^{l}\big{)}-h_{\epsilon}^{2}(u_{\epsilon}^{1},u_{\epsilon}^{2}),\phi_{\epsilon,h}\big{)}_{\Gamma_{\epsilon,h}}.

Choosing ϕϵ,h:=δuϵ2\phi_{\epsilon,h}:=\delta u_{\epsilon}^{2} (more precisely we take the restriction of uϵ2u_{\epsilon}^{2} to Ωϵ,h2\Omega_{\epsilon,h}^{2}) we obtain with the coercivity of D2D^{2} and DΓ2D^{2}_{\Gamma}, as well as the Lipschitz continuity of f2f^{2} and h2h^{2}

12ddtδuϵ2𝕃2,ϵ,h2+c0δuϵ2𝕃2,ϵ,h2\displaystyle\frac{1}{2}\frac{d}{dt}\|\delta u_{\epsilon}^{2}\|^{2}_{\mathbb{L}_{2,\epsilon,h}}+c_{0}\|\nabla\delta u_{\epsilon}^{2}\|_{\mathbb{L}_{2,\epsilon,h}}^{2} C(δuϵ2L2(Ωϵ,h2)2+ϵj=12δuϵjL2(Γϵ,h)2)\displaystyle\leq C\left(\|\delta u_{\epsilon}^{2}\|^{2}_{L^{2}(\Omega_{\epsilon,h}^{2})}+\epsilon\sum_{j=1}^{2}\|\delta u_{\epsilon}^{j}\|^{2}_{L^{2}(\Gamma_{\epsilon,h})}\right)
j=12(C(θ)δuϵjL2(Ωϵ,hj)2+ϵθuϵjL2(Ωϵ,hj)2),\displaystyle\leq\sum_{j=1}^{2}\left(C(\theta)\|\delta u_{\epsilon}^{j}\|_{L^{2}(\Omega_{\epsilon,h}^{j})}^{2}+\epsilon\theta\|\nabla u_{\epsilon}^{j}\|^{2}_{L^{2}(\Omega_{\epsilon,h}^{j})}\right),

for arbitrary θ>0\theta>0, where we used the trace-inequality (4)\eqref{TraceInequality}. Choosing θ\theta small enough the gradient term for j=2j=2 can be absorbed from the left-hand side. Integrating with respect to time, using the a priori estimates from Proposition 2 for the gradients of uϵ1u_{\epsilon}^{1}, as well as the Gronwall-inequality, we obtain the desired result. ∎

4 Two-scale compactness results

In this section we prove general strong two-scale compactness results for functions in the space L2((0,T),j,ϵ)H1((0,T),j,ϵ)L^{2}((0,T),\mathbb{H}_{j,\epsilon})\cap H^{1}((0,T),\mathbb{H}_{j,\epsilon}^{\prime}) for j{1,2}j\in\{1,2\} based on suitable a priori estimates. These estimates are fulfilled by the microscopic solution uϵ=(uϵ1,uϵ2)u_{\epsilon}=(u_{\epsilon}^{1},u_{\epsilon}^{2}) which fulfills Proposition 2 and Proposition 3, but are not restricted to them. The connected and disconnected case are completely different and are therefore treated separately. These strong compactness results are enough to pass to the limit in the nonlinear terms in the microscopic equation (3)\eqref{VariationalEquationMicroscopicProblem}, in fact we have:

Lemma 1.

Let p(1,)p\in(1,\infty).

  1. (i)

    For j{1,2}j\in\{1,2\} let (uϵj)Lp((0,T)×Ωϵj)(u_{\epsilon}^{j})\subset L^{p}((0,T)\times\Omega_{\epsilon}^{j}) be a sequence converging strongly in the two-scale sense to u0jLp((0,T)×Ω×Yj)u_{0}^{j}\in L^{p}((0,T)\times\Omega\times Y_{j}). Further f:[0,T]×Yj×f:[0,T]\times Y_{j}\times\mathbb{R}\rightarrow\mathbb{R} is continuous, YY-periodic with respect to the second variable, and fulfills the growth condition

    |f(t,y,z)|C(1+|z|) for all (t,y,z)[0,T]×Yj×.\displaystyle|f(t,y,z)|\leq C(1+|z|)\quad\mbox{ for all }(t,y,z)\in[0,T]\times Y_{j}\times\mathbb{R}.

    Then it holds up to a subsequence

    f(t,xϵ,uϵj)\displaystyle f\left(\cdot_{t},\frac{\cdot_{x}}{\epsilon},u_{\epsilon}^{j}\right) f(t,y,u0j)\displaystyle\rightarrow f\left(\cdot_{t},\cdot_{y},u_{0}^{j}\right) in the two-scale sense in Lp.\displaystyle\mbox{ in the two-scale sense in }L^{p}.
  2. (ii)

    Let (uϵ)Lp((0,T)×Γϵ)(u_{\epsilon})\subset L^{p}((0,T)\times\Gamma_{\epsilon}) be a sequence converging strongly in the two-scale sense on Γϵ\Gamma_{\epsilon} to u0Lp((0,T)×Ω×Γ)u_{0}\in L^{p}((0,T)\times\Omega\times\Gamma). Further h:[0,T]×Γ×h:[0,T]\times\Gamma\times\mathbb{R}\rightarrow\mathbb{R} is continuous, YY-periodic with respect to the second variable, and fulfills the growth condition

    |h(t,y,z)|C(1+|z|) for all (t,y,z)[0,T]×Γ×.\displaystyle|h(t,y,z)|\leq C(1+|z|)\quad\mbox{ for all }(t,y,z)\in[0,T]\times\Gamma\times\mathbb{R}.

    Then it holds up to a subsequence

    h(t,xϵ,uϵ)\displaystyle h\left(\cdot_{t},\frac{\cdot_{x}}{\epsilon},u_{\epsilon}\right) h(t,y,u0)\displaystyle\rightarrow h(\cdot_{t},\cdot_{y},u_{0}) in the two-scale sense on Γϵ in Lp.\displaystyle\mbox{ in the two-scale sense on }\Gamma_{\epsilon}\mbox{ in }L^{p}.

We emphasize that for functions ff and hh uniformly Lipschitz-continuous with respect to the last variable, the growth conditions are fulfilled. For such Lipschitz-continuous functions we also easily obtain the strong two-scale convergence of the whole sequence.

Proof.

We only prove (ii). The other statement follows the same way. Due to Lemma 7 the sequence 𝒯ϵuϵ\mathcal{T}_{\epsilon}u_{\epsilon} converges in Lp((0,T)×Ω×Γ)L^{p}((0,T)\times\Omega\times\Gamma) to u0u_{0}. Hence, up to a subsequence, 𝒯ϵuϵu0\mathcal{T}_{\epsilon}u_{\epsilon}\rightarrow u_{0} almost everywhere in (0,T)×Ω×Γ(0,T)\times\Omega\times\Gamma. Further we have

𝒯ϵ(h(t,xϵ,uϵ))=h(t,y,𝒯ϵuϵ)h(t,y,u0) a.e. in (0,T)×Ω×Γ.\displaystyle\mathcal{T}_{\epsilon}\left(h\left(\cdot_{t},\frac{\cdot_{x}}{\epsilon},u_{\epsilon}\right)\right)=h\left(\cdot_{t},y,\mathcal{T}_{\epsilon}u_{\epsilon}\right)\rightarrow h(\cdot_{t},\cdot_{y},u_{0})\quad\mbox{ a.e. in }(0,T)\times\Omega\times\Gamma.

The growth condition on hh implies 𝒯ϵ(h(t,xϵ,uϵ))\mathcal{T}_{\epsilon}\left(h\left(\cdot_{t},\frac{\cdot_{x}}{\epsilon},u_{\epsilon}\right)\right) bounded in Lp((0,T)×Ω×Γ)L^{p}((0,T)\times\Omega\times\Gamma). Egorov’s theorem (see also [23, Theorem 13.44]) implies

𝒯ϵ(h(t,xϵ,uϵ))h(t,y,u0) weakly in Lp((0,T)×Ω×Γ).\displaystyle\mathcal{T}_{\epsilon}\left(h\left(\cdot_{t},\frac{\cdot_{x}}{\epsilon},u_{\epsilon}\right)\right)\rightharpoonup h(\cdot_{t},\cdot_{y},u_{0})\quad\mbox{ weakly in }L^{p}((0,T)\times\Omega\times\Gamma).

As a direct consequence we obtain by density:

Lemma 2.

Let ff be as in Lemma 1. Further, let {vϵ}\{v_{\epsilon}\} be a bounded sequence in L2((0,T)×Ωϵj)L^{2}((0,T)\times\Omega_{\epsilon}^{j}) which converges strongly in the LpL^{p}-two-scale sense to v0L2((0,T)×Ω×Y)v_{0}\in L^{2}((0,T)\times\Omega\times Y) for p[1,2)p\in[1,2). Then it holds that

f(t,xϵ,vϵ(x))\displaystyle f\left(\cdot_{t},\frac{\cdot_{x}}{\epsilon},v_{\epsilon}(\cdot_{x})\right) f(t,y,v0(x))\displaystyle\rightarrow f(\cdot_{t},\cdot_{y},v_{0}(\cdot_{x})) in the two-scale sense in L2.\displaystyle\mbox{ in the two-scale sense in }L^{2}.

A similar result holds on the oscillating surface Γϵ\Gamma_{\epsilon}.

4.1 The connected domain Ωϵ1\Omega_{\epsilon}^{1}

Here we give a strong compactness result for a sequence in the connected domain Ωϵ1\Omega_{\epsilon}^{1} under suitable a priori estimates. The case of a connected perforated domain can be treated more easily than a disconnected domain, because we can extend a bounded sequence in H1(Ωϵ)H^{1}(\Omega_{\epsilon}) to a bounded sequence in H1(Ω)H^{1}(\Omega), due to [1, 12]. Hence, we can work in fixed Bochner spaces (not depending on ϵ\epsilon) and use standard methods from functional analysis. For this we need control for the time-variable, which can be obtained from the uniform bound of the time-derivative tuϵ1\partial_{t}u_{\epsilon}^{1}. However, since tuϵ1\partial_{t}u_{\epsilon}^{1} is pointwise only an element in the space 1,ϵ\mathbb{H}_{1,\epsilon}^{\prime} it is not clear if the time-derivative of the extension of uϵ1u_{\epsilon}^{1} exists and if it is bounded uniformly with respect to ϵ\epsilon (Unfortunately, this circumstance is often overseen in the existing literature). The following Lemma gives us an estimate for the difference of the shifts with respect to time for functions with generalized time-derivative. It is just an easy generalization of [19, Lemma 9].

Lemma 3.

Let VV and HH be Hilbert spaces and we assume that (V,H,V)(V,H,V^{\prime}) is a Gelfand-triple. Let vL2((0,T),V)H1((0,T),V)v\in L^{2}((0,T),V)\cap H^{1}((0,T),V^{\prime}). Then, for every ϕV\phi\in V and almost every t(0,T)t\in(0,T), s(T,T)s\in(-T,T), such that t+s(0,T)t+s\in(0,T), we have

|(v(t+s)v(t),ϕ)H||s|ϕVtvL2((t,t+s),V).\displaystyle\big{|}(v(t+s)-v(t),\phi)_{H}\big{|}\leq\sqrt{|s|}\|\phi\|_{V}\|\partial_{t}v\|_{L^{2}((t,t+s),V^{\prime})}.

Especially, it holds that

v(t+s)v(t)H2|s|v(t+s)v(t)VtvL2((t,t+s),V).\displaystyle\big{\|}v(t+s)-v(t)\big{\|}^{2}_{H}\leq\sqrt{|s|}\big{\|}v(t+s)-v(t)\big{\|}_{V}\|\partial_{t}v\|_{L^{2}((t,t+s),V^{\prime})}.
Proof.

The proof follows the same lines as the proof of [19, Lemma 9], if we replace the Gelfand-triple (H1(Ωjϵ),L2(Ωjϵ),H1(Ωjϵ))(H^{1}(\Omega_{j}^{\epsilon}),L^{2}(\Omega_{j}^{\epsilon}),H^{1}(\Omega_{j}^{\epsilon})^{\prime}) by the Gelfand-triple (V,H,V)(V,H,V^{\prime}). ∎

In the following, for vϵH1(Ωϵ1)v_{\epsilon}\in H^{1}(\Omega_{\epsilon}^{1}) we denote by v~ϵH1(Ω)\tilde{v}_{\epsilon}\in H^{1}(\Omega) the extension from [1, 12] with

v~ϵL2(Ω)CvϵL2(Ωϵ1),v~ϵL2(Ω)CvϵL2(Ωϵ1),\displaystyle\|\tilde{v}_{\epsilon}\|_{L^{2}(\Omega)}\leq C\|v_{\epsilon}\|_{L^{2}(\Omega_{\epsilon}^{1})},\quad\|\nabla\tilde{v}_{\epsilon}\|_{L^{2}(\Omega)}\leq C\|\nabla v_{\epsilon}\|_{L^{2}(\Omega_{\epsilon}^{1})},

with a constant C>0C>0 independent of ϵ\epsilon.

Proposition 4.

Let (vϵ)L2((0,T),1,ϵ)H1((0,T),1,ϵ)(v_{\epsilon})\subset L^{2}((0,T),\mathbb{H}_{1,\epsilon})\cap H^{1}((0,T),\mathbb{H}_{1,\epsilon}^{\prime}) be a sequence with

tvϵL2((0,T),1,ϵ)+vϵL2((0,T),1,ϵ)C.\displaystyle\|\partial_{t}v_{\epsilon}\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon}^{\prime})}+\|v_{\epsilon}\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon})}\leq C. (6)

There exists v0L2((0,T),H1(Ω))v_{0}\in L^{2}((0,T),H^{1}(\Omega)) such that for all β(12,1)\beta\in\left(\frac{1}{2},1\right) up to a subsequence it holds that

v~ϵv0 in L2((0,T),Hβ(Ω)).\displaystyle\tilde{v}_{\epsilon}\rightarrow v_{0}\quad\mbox{ in }L^{2}((0,T),H^{\beta}(\Omega)).

Further, it holds that (up to a subsequence)

𝒯ϵvϵv0 in L2((0,T)×Ω,1).\displaystyle\mathcal{T}_{\epsilon}v_{\epsilon}\rightarrow v_{0}\quad\mbox{ in }L^{2}((0,T)\times\Omega,\mathbb{H}_{1}).
Proof.

Since v~ϵ\tilde{v}_{\epsilon} is bounded in L2((0,T),H1(Ω))L^{2}((0,T),H^{1}(\Omega)) there exists v0L2((0,T),H1(Ω))v_{0}\in L^{2}((0,T),H^{1}(\Omega)), such that up to a subsequence vϵv_{\epsilon} converges weakly to v0v_{0} in L2((0,T),H1(Ω))L^{2}((0,T),H^{1}(\Omega)). Lemma 3 and inequality (6)\eqref{PropStarkeKonvergenzZshgGebiet} imply for 0<h00<h\ll 0

0Thvϵ(t+h)vϵ𝕃1,ϵ2𝑑tChtvϵL2((0,T),1,ϵ)0Thvϵ(t+h)vϵ1,ϵ𝑑tCh.\displaystyle\begin{aligned} \int_{0}^{T-h}\|v_{\epsilon}(t+h)-v_{\epsilon}\|^{2}_{\mathbb{L}_{1,\epsilon}}dt&\leq C\sqrt{h}\|\partial_{t}v_{\epsilon}\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon}^{\prime})}\int_{0}^{T-h}\|v_{\epsilon}(t+h)-v_{\epsilon}\|_{\mathbb{H}_{1,\epsilon}}dt\\ &\leq C\sqrt{h}.\end{aligned} (7)

Now, from the properties of the extension v~ϵ\tilde{v}_{\epsilon} we obtain

0Thv~ϵ(t+h)v~ϵL1(Ω)2𝑑tC0Thvϵ(t+h)vϵL2(Ωϵ1)2𝑑tCh.\displaystyle\int_{0}^{T-h}\|\tilde{v}_{\epsilon}(t+h)-\tilde{v}_{\epsilon}\|^{2}_{L^{1}(\Omega)}dt\leq C\int_{0}^{T-h}\|v_{\epsilon}(t+h)-v_{\epsilon}\|^{2}_{L^{2}(\Omega_{\epsilon}^{1})}dt\leq C\sqrt{h}.

Since H1(Ω)Hβ(Ω)H^{1}(\Omega)\hookrightarrow H^{\beta}(\Omega) is compact for β(12,1)\beta\in\left(\frac{1}{2},1\right) we can apply [30, Theorem 1] to (v~ϵ)(\tilde{v}_{\epsilon}) as a sequence in L2((0,T),Hβ(Ω))L^{2}((0,T),H^{\beta}(\Omega)) and obtain the strong convergence of v~ϵ\tilde{v}_{\epsilon} to v0v_{0} in L2((0,T),Hβ(Ω))L^{2}((0,T),H^{\beta}(\Omega)).

Now we prove the convergence of 𝒯ϵvϵ\mathcal{T}_{\epsilon}v_{\epsilon}. It holds that

𝒯ϵvϵv0L2((0,T)×Ω,1)2=𝒯ϵvϵv0L2((0,T)×Ω,H1(Y1))2+𝒯ϵvϵv0L2((0,T)×Ω,H1(Γ))2.\displaystyle\|\mathcal{T}_{\epsilon}v_{\epsilon}-v_{0}\|_{L^{2}((0,T)\times\Omega,\mathbb{H}_{1})}^{2}=\|\mathcal{T}_{\epsilon}v_{\epsilon}-v_{0}\|_{L^{2}((0,T)\times\Omega,H^{1}(Y_{1}))}^{2}+\|\mathcal{T}_{\epsilon}v_{\epsilon}-v_{0}\|_{L^{2}((0,T)\times\Omega,H^{1}(\Gamma))}^{2}.

We only prove the convergence to zero for the second term, since the first one can be treated in a similar way. We obtain from the properties of the unfolding operator from Lemma 6, the trace inequality, and the inequality (6)\eqref{PropStarkeKonvergenzZshgGebiet}

\displaystyle\| 𝒯ϵvϵv0L2((0,T)×Ω,H1(Γ))C𝒯ϵvϵv0L2((0,T)×Ω×Γ)+CΓ,y𝒯ϵvϵL2((0,T)×Ω×Γ)\displaystyle\mathcal{T}_{\epsilon}v_{\epsilon}-v_{0}\|_{L^{2}((0,T)\times\Omega,H^{1}(\Gamma))}\leq C\|\mathcal{T}_{\epsilon}v_{\epsilon}-v_{0}\|_{L^{2}((0,T)\times\Omega\times\Gamma)}+C\|\nabla_{\Gamma,y}\mathcal{T}_{\epsilon}v_{\epsilon}\|_{L^{2}((0,T)\times\Omega\times\Gamma)}
C(𝒯ϵvϵv0L2((0,T)×Ω×Y1)+ϵvϵL2((0,T)×Ωϵ1)+ϵ32ΓϵvϵL2((0,T)×Γϵ))\displaystyle\leq C\left(\|\mathcal{T}_{\epsilon}v_{\epsilon}-v_{0}\|_{L^{2}((0,T)\times\Omega\times Y_{1})}+\epsilon\|\nabla v_{\epsilon}\|_{L^{2}((0,T)\times\Omega_{\epsilon}^{1})}+\epsilon^{\frac{3}{2}}\|\nabla_{\Gamma_{\epsilon}}v_{\epsilon}\|_{L^{2}((0,T)\times\Gamma_{\epsilon})}\right)
C(vϵv0L2((0,T)×Ωϵ1)+𝒯ϵv0v0L2((0,T)×Ω×Y1)+ϵ).\displaystyle\leq C\left(\|v_{\epsilon}-v_{0}\|_{L^{2}((0,T)\times\Omega_{\epsilon}^{1})}+\|\mathcal{T}_{\epsilon}v_{0}-v_{0}\|_{L^{2}((0,T)\times\Omega\times Y_{1})}+\epsilon\right).

The first term converges to zero for ϵ0\epsilon\to 0, due to the strong convergence of v~ϵ\tilde{v}_{\epsilon} to v0v_{0}, and the second term because of [11, Prop. 4.4]. This gives the desired result. ∎

Remark 1.

  1. (i)

    The extension of a bounded sequence vϵv_{\epsilon} in L2((0,T),H1(Ωϵ1))L^{2}((0,T),H^{1}(\Omega_{\epsilon}^{1})) to v~ϵL2((0,T),H1(Ω))\tilde{v}_{\epsilon}\in L^{2}((0,T),H^{1}(\Omega)) preserving the boundedness is only possible for the connected domain Ωϵ1\Omega_{\epsilon}^{1}. This implies in a simple way the strong convergence of v~ϵ\tilde{v}_{\epsilon}, if there is also a control for the time-derivative. For the disconnected domain this method fails and therefore we use a Kolmogorov-Simon-compactness result for the unfolded sequence, where we need an additional condition to control the shifts with respect to the macroscopic variable, see Theorem 1.

  2. (ii)

    If we replace in the assumptions of Proposition 6 the space 1,ϵ\mathbb{H}_{1,\epsilon} by H1(Ωϵ1)H^{1}(\Omega_{\epsilon}^{1}), we obtain the well known result that a bounded sequence in L2((0,T),H1(Ωϵ1))H1((0,T),H1(Ωϵ1))L^{2}((0,T),H^{1}(\Omega_{\epsilon}^{1}))\cap H^{1}((0,T),H^{1}(\Omega_{\epsilon}^{1})^{\prime}) has an extension converging strongly in L2((0,T)×Ω)L^{2}((0,T)\times\Omega), see [25], and also in L2((0,T),Hβ(Ω))L^{2}((0,T),H^{\beta}(\Omega)) for β(12,1)\beta\in\left(\frac{1}{2},1\right), see [19]. The proof above for the strong convergence of v~ϵ\tilde{v}_{\epsilon} can be easily adapted to this situation. However, we emphasize that in our situation we cannot guarantee for the solution (uϵ1,uϵ2)(u_{\epsilon}^{1},u_{\epsilon}^{2}) that tuϵ1L2((0,T),H1(Ωϵ1))\partial_{t}u_{\epsilon}^{1}\in L^{2}((0,T),H^{1}(\Omega_{\epsilon}^{1})^{\prime}).

4.2 The disconnected domain Ωϵ2\Omega_{\epsilon}^{2}

In this section we give a strong two-scale compactness result for the disconnected domain Ωϵ2\Omega_{\epsilon}^{2} of Kolmogorov-Simon-type, i. e., it is based on a priori estimates for the difference of discrete shifts, see condition (ii) in Theorem 1. As already mentioned above it is in general not possible to find an extension for a function in H1(Ωϵ2)H^{1}(\Omega_{\epsilon}^{2}) to the whole domain Ω\Omega which preserves the a priori estimates. Hence, the method from Section 4.1 for the connected domain fails. Therefore, we consider the unfolded sequence in the Bochner space Lp(Ω,L2((0,T),2β))L^{p}(\Omega,L^{2}((0,T),\mathbb{H}_{2}^{\beta})) with β(12,1)\beta\in\left(\frac{1}{2},1\right) and p(1,2)p\in(1,2), and apply the Kolmogorov-Simon-compactness result from [17], which gives an extension of [30, Theorem 1] to higher-dimensional domains of definition. Here, a crucial point is the estimate for the shifts. An important reason to work here with general Bochner spaces, i. e., Banach-valued functions spaces, is that we are dealing with manifolds and therefore linear shifts with respect to the space variable are not well defined.

In the following Lemma we estimate the shifts of the unfolded sequence with respect to the macroscopic variable by the shifts of the function itself, see again Section 3.3 for the notations.

Lemma 4.

Let vϵL2((0,T)×Ωϵj)v_{\epsilon}\in L^{2}((0,T)\times\Omega_{\epsilon}^{j}) for j{1,2}j\in\{1,2\} and wϵL2((0,T)×Γϵ)w_{\epsilon}\in L^{2}((0,T)\times\Gamma_{\epsilon}). Then, for 0<h10<h\ll 1 , |z|<h|z|<h, and ϵ\epsilon small enough it holds that

𝒯ϵvϵ(t,x+z,y)𝒯ϵvϵL2(0,T)×Ω2h×Yj)2\displaystyle\big{\|}\mathcal{T}_{\epsilon}v_{\epsilon}(t,x+z,y)-\mathcal{T}_{\epsilon}v_{\epsilon}\big{\|}_{L^{2}(0,T)\times\Omega_{2h}\times Y_{j})}^{2} k{0,1}nδvϵL2((0,T)×Ωϵ,hj)2,\displaystyle\leq\sum_{k\in\{0,1\}^{n}}\|\delta v_{\epsilon}\|_{L^{2}((0,T)\times\Omega_{\epsilon,h}^{j})}^{2},
𝒯ϵwϵ(t,x+z,y)𝒯ϵwϵL2(0,T)×Ω2h×Γ)2\displaystyle\big{\|}\mathcal{T}_{\epsilon}w_{\epsilon}(t,x+z,y)-\mathcal{T}_{\epsilon}w_{\epsilon}\big{\|}_{L^{2}(0,T)\times\Omega_{2h}\times\Gamma)}^{2} ϵk{0,1}nδwϵL2((0,T)×Γϵ,h)2,\displaystyle\leq\epsilon\sum_{k\in\{0,1\}^{n}}\|\delta w_{\epsilon}\|_{L^{2}((0,T)\times\Gamma_{\epsilon,h})}^{2},

with l=l(ϵ,z,k)=k+[zϵ]l=l(\epsilon,z,k)=k+\left[\frac{z}{\epsilon}\right].

Proof.

The proof for a thin layer can be found in [27, p. 709] and can be easily extended to our setting. See also [19] for more details. ∎

Theorem 1.

Let vϵL2((0,T),2,ϵ)H1((0,T),2,ϵ)v_{\epsilon}\in L^{2}((0,T),\mathbb{H}_{2,\epsilon})\cap H^{1}((0,T),\mathbb{H}_{2,\epsilon}^{\prime}) with:

  1. (i)

    It holds the estimate

    vϵL2((0,T),2,ϵ)+tvϵL2((0,T),2,ϵ)C.\displaystyle\|v_{\epsilon}\|_{L^{2}((0,T),\mathbb{H}_{2,\epsilon})}+\|\partial_{t}v_{\epsilon}\|_{L^{2}((0,T),\mathbb{H}_{2,\epsilon}^{\prime})}\leq C.
  2. (ii)

    For 0<h10<h\ll 1 and lnl\in\mathbb{Z}^{n} with |lϵ|<h|l\epsilon|<h it holds that

    δvϵL2((0,T),L2(Ωϵ,h2))ϵl00.\displaystyle\|\delta v_{\epsilon}\|_{L^{2}((0,T),L^{2}(\Omega_{\epsilon,h}^{2}))}\overset{\epsilon l\to 0}{\longrightarrow}0.

Then, there exists v0L2((0,T),L2(Ω))v_{0}\in L^{2}((0,T),L^{2}(\Omega)), such that for β(12,1)\beta\in\left(\frac{1}{2},1\right) and p[1,2)p\in[1,2) it holds up to a subsequence that

𝒯ϵvϵv0 in Lp(Ω,L2((0,T),2β)).\displaystyle\mathcal{T}_{\epsilon}v_{\epsilon}\rightarrow v_{0}\quad\mbox{ in }L^{p}(\Omega,L^{2}((0,T),\mathbb{H}_{2}^{\beta})).

Especially, vϵv_{\epsilon} and vϵ|Γϵv_{\epsilon}|_{\Gamma_{\epsilon}} converge strongly in the two-scale sense to v0v_{0} (with respect to LpL^{p}).

Proof.

Our aim is to apply [17, Corollary 2.5] to (𝒯ϵvϵ)(\mathcal{T}_{\epsilon}v_{\epsilon}) as a sequence in Lp(Ω,L2((0,T),2β))L^{p}(\Omega,L^{2}((0,T),\mathbb{H}_{2}^{\beta})) for p[1,2)p\in[1,2) and β(12,1)\beta\in\left(\frac{1}{2},1\right). Hence, we have to check the following three conditions:

  1. (K1)

    For every measurable set AΩA\subset\Omega the sequence {Avϵ𝑑x}\left\{\int_{A}v_{\epsilon}dx\right\} is relatively compact in L2((0,T),2β)L^{2}((0,T),\mathbb{H}_{2}^{\beta}),

  2. (K2)

    for all 0<h10<h\ll 1 and |z|<h|z|<h it holds that

    supϵvϵ(+z)vϵLp(Ωh,B)0 for z0,\displaystyle\sup_{\epsilon}\|v_{\epsilon}(\cdot+z)-v_{\epsilon}\|_{L^{p}(\Omega_{h},B)}\rightarrow 0\quad\mbox{ for }z\to 0,
  3. (K3)

    for h>0h>0 it holds that supϵΩΩh|vϵ(x)|p𝑑x0 for h0\,\sup_{\epsilon}\int_{\Omega\setminus\Omega_{h}}|v_{\epsilon}(x)|^{p}dx\rightarrow 0\,\mbox{ for }h\to 0.

We start with the condition (K1). Let AΩA\subset\Omega be measurable and we define Vϵ(t,y):=A𝒯ϵvϵ(t,x,y)𝑑xV_{\epsilon}(t,y):=\int_{A}\mathcal{T}_{\epsilon}v_{\epsilon}(t,x,y)dx. To show the relative compactness of (Vϵ)(V_{\epsilon}), we use again [30, Theorem 1] as in the proof of Proposition 4. First of all, due to our assumptions on vϵv_{\epsilon} and the properties of the unfolding operator, for t1,t2(0,T)t_{1},t_{2}\in(0,T) it holds that

t1t2Vϵ𝑑t2𝒯ϵvϵL2((0,T)×Ω,2)C.\displaystyle\left\|\int_{t_{1}}^{t_{2}}V_{\epsilon}dt\right\|_{\mathbb{H}_{2}}\leq\|\mathcal{T}_{\epsilon}v_{\epsilon}\|_{L^{2}((0,T)\times\Omega,\mathbb{H}_{2})}\leq C.

Due to the compact embedding 22β\mathbb{H}_{2}\hookrightarrow\mathbb{H}_{2}^{\beta} we obtain that t1t2Vϵ𝑑t\int_{t_{1}}^{t_{2}}V_{\epsilon}dt is relatively compact in 2β\mathbb{H}_{2}^{\beta}. Further, for 0<s10<s\ll 1 we obtain with the estimates for vϵv_{\epsilon} and the trace inequality (4)\eqref{TraceInequality}

\displaystyle\big{\|} Vϵ(t+s,y)VϵL2((0,Ts),2)𝒯ϵvϵ(t+s,x,y)𝒯ϵvϵL2((0,Ts)×Ω,2)\displaystyle V_{\epsilon}(t+s,y)-V_{\epsilon}\big{\|}_{L^{2}((0,T-s),\mathbb{H}_{2})}\leq\big{\|}\mathcal{T}_{\epsilon}v_{\epsilon}(t+s,x,y)-\mathcal{T}_{\epsilon}v_{\epsilon}\big{\|}_{L^{2}((0,T-s)\times\Omega,\mathbb{H}_{2})}
Cvϵ(t+s,x)vϵL2((0,Ts)×Ωϵ2)+Cϵvϵ(t+s,x)vϵL2((0,Ts)×Ωϵ2)\displaystyle\leq C\|v_{\epsilon}(t+s,x)-v_{\epsilon}\|_{L^{2}((0,T-s)\times\Omega_{\epsilon}^{2})}+C\epsilon\|\nabla v_{\epsilon}(t+s,x)-\nabla v_{\epsilon}\|_{L^{2}((0,T-s)\times\Omega_{\epsilon}^{2})}
+Cϵ32Γϵvϵ(t+s,x)ΓϵvϵL2((0,Ts)×Γϵ)\displaystyle\hskip 30.00005pt+C\epsilon^{\frac{3}{2}}\|\nabla_{\Gamma_{\epsilon}}v_{\epsilon}(t+s,x)-\nabla_{\Gamma_{\epsilon}}v_{\epsilon}\|_{L^{2}((0,T-s)\times\Gamma_{\epsilon})}
C(s14+ϵ),\displaystyle\leq C\left(s^{\frac{1}{4}}+\epsilon\right),

where for the last inequality we used Lemma 3 to estimate the first term in the line before by using the same arguments as for the inequality (7)\eqref{AuxiliaryEstimateShiftTime} in the proof of Proposition 4. Hence, [30, Theorem 1] implies that (Vϵ)(V_{\epsilon}) is relatively compact in L2((0,T),2β)L^{2}((0,T),\mathbb{H}_{2}^{\beta}), i. e., condition (K1).

For (K2) we fix 0<h10<h\ll 1 and choose |z|<h|z|<h. Lemma 4 with l=k+[zϵ]l=k+\left[\frac{z}{\epsilon}\right] (see the definition of the difference δ\delta in (5)\eqref{DefinitionDifference}), the conditions (i) and (ii), as well as the trace inequality (4)\eqref{TraceInequality} imply

\displaystyle\big{\|} 𝒯ϵvϵ(t,x+z,y)𝒯ϵvϵL2(Ω2h,L2((0,T),2))\displaystyle\mathcal{T}_{\epsilon}v_{\epsilon}(t,x+z,y)-\mathcal{T}_{\epsilon}v_{\epsilon}\big{\|}_{L^{2}(\Omega_{2h},L^{2}((0,T),\mathbb{H}_{2}))}
Ck{0,1}n(δvϵL2((0,T)×Ωϵ,h2)+ϵδvϵL2((0,T)×Ωϵ,h2)+ϵ32δΓϵvϵL2((0,T)×Γϵ,h))\displaystyle\leq C\sum_{k\in\{0,1\}^{n}}\left(\|\delta v_{\epsilon}\|_{L^{2}((0,T)\times\Omega_{\epsilon,h}^{2})}+\epsilon\|\delta\nabla v_{\epsilon}\|_{L^{2}((0,T)\times\Omega_{\epsilon,h}^{2})}+\epsilon^{\frac{3}{2}}\|\delta\nabla_{\Gamma_{\epsilon}}v_{\epsilon}\|_{L^{2}((0,T)\times\Gamma_{\epsilon,h})}\right)
C(j{0,1}nδvϵL2((0,T)×Ωϵ,h2)+ϵ)ϵ,z00.\displaystyle\leq C\left(\sum_{j\in\{0,1\}^{n}}\|\delta v_{\epsilon}\|_{L^{2}((0,T)\times\Omega_{\epsilon,h}^{2})}+\epsilon\right)\overset{\epsilon,z\to 0}{\longrightarrow}0.

We show that this implies the uniform convergence in (K2) with respect to ϵ\epsilon, see also [27, p.710-711] or [14, p.1476-1477]. Let 0<ρ0<\rho. Due to our previous results there exist 0<ϵ0,δ00<\epsilon_{0},\,\delta_{0}, such that for all ϵϵ0\epsilon\leq\epsilon_{0} and |z|δ0|z|\leq\delta_{0} it holds that

𝒯ϵvϵ(t,x+z,y)𝒯ϵvϵL2(Ω2h,L2((0,T),2))ρ.\displaystyle\big{\|}\mathcal{T}_{\epsilon}v_{\epsilon}(t,x+z,y)-\mathcal{T}_{\epsilon}v_{\epsilon}\big{\|}_{L^{2}(\Omega_{2h},L^{2}((0,T),\mathbb{H}_{2}))}\leq\rho. (8)

Since ϵ1\epsilon^{-1}\in\mathbb{N}, there are only finitely many elements ϵi\epsilon_{i} with i=1,,Ni=1,\ldots,N, such that ϵ0<ϵi\epsilon_{0}<\epsilon_{i}. For every ϵi\epsilon_{i} there exists a 0<δi0<\delta_{i}, such that (8)\eqref{UniformEstimateShifts} is valid for ϵ=ϵi\epsilon=\epsilon_{i} and all |z|δi|z|\leq\delta_{i}. Choosing δ:=maxi=0,,N{δi}\delta:=\max_{i=0,\ldots,N}\{\delta_{i}\}, inequality (8)\eqref{UniformEstimateShifts} holds uniformly with respect to ϵ\epsilon for all |z|δ|z|\leq\delta. This implies (K2). For the last condition (K3) we use the Hölder-inequality to obtain for p[1,2)p\in[1,2) and 0<h10<h\ll 1

𝒯ϵvϵLp(ΩΩh,L2((0,T),2))|ΩΩh|2p2p𝒯ϵvϵL2((0,T)×ΩΩh,2)Ch2p2ph00,\displaystyle\big{\|}\mathcal{T}_{\epsilon}v_{\epsilon}\big{\|}_{L^{p}(\Omega\setminus\Omega_{h},L^{2}((0,T),\mathbb{H}_{2}))}\leq|\Omega\setminus\Omega_{h}|^{\frac{2-p}{2p}}\big{\|}\mathcal{T}_{\epsilon}v_{\epsilon}\big{\|}_{L^{2}((0,T)\times\Omega\setminus\Omega_{h},\mathbb{H}_{2})}\leq Ch^{\frac{2-p}{2p}}\overset{h\to 0}{\longrightarrow}0,

where we used again estimate (i). Now, [17, Corollary 2.5] implies the the strong convergence of 𝒯ϵvϵ\mathcal{T}_{\epsilon}v_{\epsilon} up to a subsequence in Lp(Ω,L2((0,T),2β))L^{p}(\Omega,L^{2}((0,T),\mathbb{H}_{2}^{\beta})) to a limit function v0v_{0}. Lemma 7 implies the strong two-scale convergence of vϵv_{\epsilon} to the same limit. The fact v0L2((0,T),L2(Ω))v_{0}\in L^{2}((0,T),L^{2}(\Omega)) follows from standard two-scale compactness results, see [2], based on the estimate (i). ∎

Remark 2.

Theorem 1 and its proof remain valid if we replace 2,ϵ\mathbb{H}_{2,\epsilon} and β\mathbb{H}^{\beta} by H1(Ωϵ2)H^{1}(\Omega_{\epsilon}^{2}) and Hβ(Y2)H^{\beta}(Y_{2}).

5 Derivation of the macroscopic model

The aim of this section is the derivation of the macroscopic model (12)\eqref{MacroscopicModel} from Theorem 2 for ϵ0\epsilon\to 0. Therefore we make use of compactness results from Section 4 and the a priori estimates from Section 3. In the following Proposition we collect the convergence results for the microscopic solution uϵ=(uϵ1,uϵ2)u_{\epsilon}=(u_{\epsilon}^{1},u_{\epsilon}^{2}):

Proposition 5.

Let uϵ=(uϵ1,uϵ2)u_{\epsilon}=(u_{\epsilon}^{1},u_{\epsilon}^{2}) be the microscopic solution of the problem (1)\eqref{MicroscopicModel}. There exist

u01L2((0,T),H1(Ω)),u11L2((0,T),1/),u02L2((0,T)×Ω),\displaystyle u_{0}^{1}\in L^{2}((0,T),H^{1}(\Omega)),\quad u_{1}^{1}\in L^{2}((0,T),\mathbb{H}_{1}/\mathbb{R}),\quad u_{0}^{2}\in L^{2}((0,T)\times\Omega),
such that up to a subsequence it holds for p[1,2)p\in[1,2)
uϵ1\displaystyle u_{\epsilon}^{1} u01\displaystyle\rightarrow u_{0}^{1} strongly in the two-scale sense,\displaystyle\mbox{ strongly in the two-scale sense}, (9a)
uϵ1\displaystyle\nabla u_{\epsilon}^{1} u01+yu11\displaystyle\rightarrow\nabla u_{0}^{1}+\nabla_{y}u_{1}^{1} in the two-scale sense,\displaystyle\mbox{ in the two-scale sense}, (9b)
uϵ1|Γϵ\displaystyle u_{\epsilon}^{1}|_{\Gamma_{\epsilon}} u01\displaystyle\rightarrow u_{0}^{1} strongly in the two-scale sense on Γϵ,\displaystyle\mbox{ strongly in the two-scale sense on }\Gamma_{\epsilon}, (9c)
Γϵuϵ1\displaystyle\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{1} PΓu01+Γ,yu11\displaystyle\rightarrow P_{\Gamma}\nabla u_{0}^{1}+\nabla_{\Gamma,y}u_{1}^{1} in the two-scale sense on Γϵ,\displaystyle\mbox{ in the two-scale sense on }\Gamma_{\epsilon}, (9d)
uϵ2\displaystyle u_{\epsilon}^{2} u02\displaystyle\rightarrow u_{0}^{2} strongly in the two-scale sense in Lp,\displaystyle\mbox{ strongly in the two-scale sense in }L^{p}, (9e)
uϵ2\displaystyle\nabla u_{\epsilon}^{2} 0\displaystyle\rightarrow 0 in the two-scale sense,\displaystyle\mbox{ in the two-scale sense}, (9f)
uϵ2|Γϵ\displaystyle u_{\epsilon}^{2}|_{\Gamma_{\epsilon}} u02\displaystyle\rightarrow u_{0}^{2} strongly in the two-scale sense in Lp on Γϵ,\displaystyle\mbox{ strongly in the two-scale sense in }L^{p}\mbox{ on }\Gamma_{\epsilon}, (9g)
Γϵuϵ2\displaystyle\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{2} 0\displaystyle\rightarrow 0 in the two-scale sense on Γϵ.\displaystyle\mbox{ in the two-scale sense on }\Gamma_{\epsilon}. (9h)
Proof.

The convergence results (9a)\eqref{TSConvMicSolConnectedDomainSol} - (9d)\eqref{TSConvMicSolConnectDomainSurfaceGradient} follow immediately from Proposition 4, Lemma 5, and the a priori estimates in Proposition 2.
For (9e)\eqref{TSConvMicSolDisconnectedDomainSol} - (9h)\eqref{TSConvMicSolDisconnectedDomainSurfaceGradient} we first notice that due to Lemma 5 there exists u02L2((0,T)×Ω)u_{0}^{2}\in L^{2}((0,T)\times\Omega) and u12L2((0,T)×Ω,2/)u_{1}^{2}\in L^{2}((0,T)\times\Omega,\mathbb{H}_{2}/\mathbb{R}), such that up to a subsequence

uϵ2\displaystyle u_{\epsilon}^{2} u02\displaystyle\rightarrow u_{0}^{2} in the two-scale sense ,
uϵ2\displaystyle\nabla u_{\epsilon}^{2} yu12\displaystyle\rightarrow\nabla_{y}u_{1}^{2} in the two-scale sense ,
uϵ2|Γϵ\displaystyle u_{\epsilon}^{2}|_{\Gamma_{\epsilon}} u02\displaystyle\rightarrow u_{0}^{2} in the two-scale sense on Γϵ,\displaystyle\Gamma_{\epsilon},
Γϵuϵ2|Γϵ\displaystyle\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{2}|_{\Gamma_{\epsilon}} Γu12\displaystyle\rightarrow\nabla_{\Gamma}u_{1}^{2} in the two-scale sense on Γϵ.\displaystyle\Gamma_{\epsilon}.

For the strong two-scale convergence of uϵ2u_{\epsilon}^{2} and uϵ2|Γϵu_{\epsilon}^{2}|_{\Gamma_{\epsilon}} we make use of Theorem 1, where we have to check the conditions (i) and (ii). The first one is just the a priori estimate from Proposition 2. For (ii) we use Proposition 3 to obtain for fixed 0<h10<h\ll 1 and lnl\in\mathbb{Z}^{n} with ϵ|l|<h\epsilon|l|<h

δuϵ2L2((0,T),L2(Ωϵ,h2))C(δuϵ1L2((0,T),L2(Ωϵ,h1))+δ(uϵ,i2,uϵ,i,Γϵ2)𝕃2,ϵ,h+ϵ).\displaystyle\|\delta u_{\epsilon}^{2}\|_{L^{2}((0,T),L^{2}(\Omega_{\epsilon,h}^{2}))}\leq C\left(\|\delta u_{\epsilon}^{1}\|_{L^{2}((0,T),L^{2}(\Omega_{\epsilon,h}^{1}))}+\big{\|}\delta\big{(}u_{\epsilon,i}^{2},u_{\epsilon,i,\Gamma_{\epsilon}}^{2}\big{)}\big{\|}_{\mathbb{L}_{2,\epsilon,h}}+\epsilon\right).

For the first term on the right-hand side we have

δuϵ1L2((0,T),L2(Ωϵ,h1))𝒯ϵuϵ1(x+lϵ,y)𝒯ϵuϵ1L2((0,T)×Ωh×Y1).\displaystyle\|\delta u_{\epsilon}^{1}\|_{L^{2}((0,T),L^{2}(\Omega_{\epsilon,h}^{1}))}\leq\big{\|}\mathcal{T}_{\epsilon}u_{\epsilon}^{1}(x+l\epsilon,y)-\mathcal{T}_{\epsilon}u_{\epsilon}^{1}\big{\|}_{L^{2}((0,T)\times\Omega_{h}\times Y_{1})}.

The right-hand side converges to zero, due to the strong two-scale convergence of uϵ1u_{\epsilon}^{1}, i. e., the strong convergence of 𝒯ϵuϵ1\mathcal{T}_{\epsilon}u_{\epsilon}^{1} in L2((0,T)×Ω×Y1)L^{2}((0,T)\times\Omega\times Y_{1}), and the standard Kolmogorov-compactness theorem. For the L2L^{2}-norm of δuϵ,i2\delta u_{\epsilon,i}^{2} in the second term we argue in a similar way, where we can use the strong two-scale convergence in the Assumption (A5). For the norm of δuϵ,i,Γϵ2\delta u_{\epsilon,i,\Gamma_{\epsilon}}^{2} we can use the Kolmogorov-Simon-compactness result from [17, Corollary 2.5], applied to the strong convergent sequence (𝒯ϵuϵ,i,Γϵ2)\big{(}\mathcal{T}_{\epsilon}u_{\epsilon,i,\Gamma_{\epsilon}}^{2}\big{)} in L2((0,T)×Ω,L2(Γ))L^{2}((0,T)\times\Omega,L^{2}(\Gamma)).

To prove (9f)\eqref{TSConvMicSolDisconnectedDomainGradient} and (9h)\eqref{TSConvMicSolDisconnectedDomainSurfaceGradient} we have to show that u12u_{1}^{2} is constant with respect to yy. Therefore, we choose ϕϵ(t,x):=ϵϕ(t,x,xϵ)\phi_{\epsilon}(t,x):=\epsilon\phi\left(t,x,\frac{x}{\epsilon}\right) with ϕC0((0,T)×Ω×Y2¯)\phi\in C^{\infty}_{0}((0,T)\times\Omega\times\overline{Y_{2}}) (periodically extended in the last variable) as a test-function in (3)\eqref{VariationalEquationMicroscopicProblem} for j=2j=2 and integrate with respect to time to obtain

0Ttuϵ2,ϕϵ2,ϵ,2,ϵdt+0TΩϵ2Dϵ2uϵ2(ϵxϕ(t,x,xϵ)+yϕ(t,x,xϵ))dxdt+ϵ0TΓϵDΓϵ2Γϵuϵ2(ϵPΓϵxϕ(t,x,xϵ)+PΓϵyϕ(t,x,xϵ))𝑑σ𝑑t=ϵ0TΩϵ2fϵ2(uϵ2)ϕ(t,x,xϵ)𝑑x𝑑t+ϵ20TΩϵ2hϵ2(uϵ1,uϵ2)ϕ(t,x,xϵ)𝑑x𝑑t.\displaystyle\begin{aligned} \int_{0}^{T}\langle&\partial_{t}u_{\epsilon}^{2},\phi_{\epsilon}\rangle_{\mathbb{H}_{2,\epsilon}^{\prime},\mathbb{H}_{2,\epsilon}}dt+\int_{0}^{T}\int_{\Omega_{\epsilon}^{2}}D_{\epsilon}^{2}\nabla u_{\epsilon}^{2}\cdot\left(\epsilon\nabla_{x}\phi\left(t,x,\frac{x}{\epsilon}\right)+\nabla_{y}\phi\left(t,x,\frac{x}{\epsilon}\right)\right)dxdt\\ &+\epsilon\int_{0}^{T}\int_{\Gamma_{\epsilon}}D_{\Gamma_{\epsilon}}^{2}\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{2}\cdot\left(\epsilon P_{\Gamma_{\epsilon}}\nabla_{x}\phi\left(t,x,\frac{x}{\epsilon}\right)+P_{\Gamma_{\epsilon}}\nabla_{y}\phi\left(t,x,\frac{x}{\epsilon}\right)\right)d\sigma dt\\ =&\epsilon\int_{0}^{T}\int_{\Omega_{\epsilon}^{2}}f_{\epsilon}^{2}(u_{\epsilon}^{2})\phi\left(t,x,\frac{x}{\epsilon}\right)dxdt+\epsilon^{2}\int_{0}^{T}\int_{\Omega_{\epsilon}^{2}}h_{\epsilon}^{2}(u_{\epsilon}^{1},u_{\epsilon}^{2})\phi\left(t,x,\frac{x}{\epsilon}\right)dxdt.\end{aligned} (10)

For the first term on the left-hand side including the time-derivative we get by integration by parts in time

0Ttuϵ2,ϕϵ2,ϵ,2,ϵ𝑑t=ϵ\displaystyle\int_{0}^{T}\langle\partial_{t}u_{\epsilon}^{2},\phi_{\epsilon}\rangle_{\mathbb{H}_{2,\epsilon}^{\prime},\mathbb{H}_{2,\epsilon}}dt=-\epsilon 0TΩϵ2tϕ(t,x,xϵ)uϵ2dxdt\displaystyle\int_{0}^{T}\int_{\Omega_{\epsilon}^{2}}\partial_{t}\phi\left(t,x,\frac{x}{\epsilon}\right)u_{\epsilon}^{2}dxdt
ϵ20TΓϵtϕ(t,x,xϵ)uϵ2dσdt.\displaystyle-\epsilon^{2}\int_{0}^{T}\int_{\Gamma_{\epsilon}}\partial_{t}\phi\left(t,x,\frac{x}{\epsilon}\right)u_{\epsilon}^{2}d\sigma dt.

The right-hand side is of order ϵ\epsilon, due to the estimates in Proposition 2. Hence, all terms in (10)\eqref{HilfsgleichungGrenzwertNullGradientUepsTwo} except the terms including the y\nabla_{y} are of order ϵ\epsilon (again because of Proposition 2) and we obtain for ϵ0\epsilon\to 0

0TΩY2\displaystyle\int_{0}^{T}\int_{\Omega}\int_{Y_{2}} D2(y)yu12(t,x,y)yϕ(t,x,y)dydxdt\displaystyle D^{2}(y)\nabla_{y}u_{1}^{2}(t,x,y)\cdot\nabla_{y}\phi(t,x,y)dydxdt
+0TΩΓDΓ2(y)Γ,yu12(t,x,y)Γ,yϕ(t,x,y)𝑑σy𝑑x𝑑t=0.\displaystyle+\int_{0}^{T}\int_{\Omega}\int_{\Gamma}D_{\Gamma}^{2}(y)\nabla_{\Gamma,y}u_{1}^{2}(t,x,y)\cdot\nabla_{\Gamma,y}\phi(t,x,y)d\sigma_{y}dxdt=0.

Due to the density of C(Y2¯)C^{\infty}(\overline{Y_{2}}) in 2\mathbb{H}_{2}, see [16, Lemma 2.1], the equation above holds for all ϕL2((0,T)×Ω,2)\phi\in L^{2}((0,T)\times\Omega,\mathbb{H}_{2}). This implies u12=0u_{1}^{2}=0. The Proposition is proved.

We have the following representation of u11u_{1}^{1}:

Corollary 1.

Almost everywhere in (0,T)×Ω×Y1(0,T)\times\Omega\times Y_{1} it holds that

u11(t,x,y)=i=1nxiu01(t,x)wi1(y),\displaystyle u_{1}^{1}(t,x,y)=\sum_{i=1}^{n}\partial_{x_{i}}u_{0}^{1}(t,x)w_{i}^{1}(y), (11)

where wi11/w_{i}^{1}\in\mathbb{H}_{1}/\mathbb{R} with YY-periodic boundary conditions is the unique weak solution of the following cell problem (i=1,,ni=1,\ldots,n)

y(D1(ywi1+ei))\displaystyle-\nabla_{y}\cdot\big{(}D^{1}(\nabla_{y}w_{i}^{1}+e_{i})\big{)} =0\displaystyle=0 in Y1,\displaystyle\mbox{ in }Y_{1},
D1(ywi1+ei)ν\displaystyle-D^{1}(\nabla_{y}w_{i}^{1}+e_{i})\cdot\nu =Γ,y(DΓ1(Γ,ywi1+Γ,yyi))\displaystyle=-\nabla_{\Gamma,y}\cdot\big{(}D_{\Gamma}^{1}(\nabla_{\Gamma,y}w_{i}^{1}+\nabla_{\Gamma,y}y_{i})\big{)} on Γ,\displaystyle\mbox{ on }\Gamma,
wi1 is\displaystyle w_{i}^{1}\mbox{ is } Y-periodic and Γwi1𝑑σ=0.\displaystyle Y\mbox{-periodic and }\int_{\Gamma}w_{i}^{1}d\sigma=0.
Proof.

The procedure is quite standard, see e. g.,  [2], but for the sake of completeness we give the main steps. We choose ϕϵ(t,x)=ϵϕ(t,x,xϵ)\phi_{\epsilon}(t,x)=\epsilon\phi\left(t,x,\frac{x}{\epsilon}\right) with ϕC0((0,T)×Ω,Cper(Y1¯))\phi\in C^{\infty}_{0}((0,T)\times\Omega,C_{\mathrm{per}}^{\infty}(\overline{Y_{1}})) as a test-function in (3)\eqref{VariationalEquationMicroscopicProblem} and integrate with respect to time to obtain (10)\eqref{HilfsgleichungGrenzwertNullGradientUepsTwo} if we replace j=2j=2 by j=1j=1. From Proposition 5 we get for ϵ0\epsilon\to 0

0=\displaystyle 0= 0TΩY1D1(y)[xu01(t,x)+yu11(t,x,y)]yϕ(t,x,y)𝑑y𝑑x𝑑t\displaystyle\int_{0}^{T}\int_{\Omega}\int_{Y_{1}}D^{1}(y)\left[\nabla_{x}u_{0}^{1}(t,x)+\nabla_{y}u_{1}^{1}(t,x,y)\right]\cdot\nabla_{y}\phi(t,x,y)dydxdt
+0TΩΓDΓ1(y)[PΓ(y)xu01(t,x)+Γ,yu11(t,x,y)]Γ,yϕ(t,x,y)𝑑σy𝑑x𝑑t.\displaystyle+\int_{0}^{T}\int_{\Omega}\int_{\Gamma}D_{\Gamma}^{1}(y)\left[P_{\Gamma}(y)\nabla_{x}u_{0}^{1}(t,x)+\nabla_{\Gamma,y}u_{1}^{1}(t,x,y)\right]\cdot\nabla_{\Gamma,y}\phi(t,x,y)d\sigma_{y}dxdt.

Due to the Lax-Milgram Lemma this problem has a unique solution u11u_{1}^{1} and due to its linearity we easily obtain the representation (11)\eqref{RepresentationUoo}. ∎

Now, we are able to formulate the macroscopic model. We show that u0=(u01,u02)u_{0}=(u_{0}^{1},u_{0}^{2}) from Proposition 5 is the unique weak solution (the definition of a weak solution is given below) of the macro-model

(|Y1|+|Γ|)tu01(D^1u01)=Y1f1(t,y,u01)𝑑y+Γh1(t,y,u01,u02)𝑑σy in (0,T)×Ω,(|Y2|+|Γ|)tu02=Y2f2(t,y,u02)𝑑y+Γh2(t,y,u01,u02)𝑑σy in (0,T)×Ω,D^1u01ν=0 on (0,T)×Ω,u0j(0)=|Yj|u0,ij+|Γ|u0,i,Γj|Yj|+|Γ| in Ω,\displaystyle\begin{aligned} (|Y_{1}|+|\Gamma|)\partial_{t}u_{0}^{1}-\nabla\cdot\big{(}\widehat{D}^{1}\nabla u_{0}^{1}\big{)}&=\int_{Y_{1}}f^{1}(t,y,u_{0}^{1})dy+\int_{\Gamma}h^{1}(t,y,u_{0}^{1},u_{0}^{2})d\sigma_{y}&\mbox{ in }&(0,T)\times\Omega,\\ (|Y_{2}|+|\Gamma|)\partial_{t}u_{0}^{2}&=\int_{Y_{2}}f^{2}(t,y,u_{0}^{2})dy+\int_{\Gamma}h^{2}(t,y,u_{0}^{1},u_{0}^{2})d\sigma_{y}&\mbox{ in }&(0,T)\times\Omega,\\ -\widehat{D}^{1}\nabla u_{0}^{1}\cdot\nu&=0&\mbox{ on }&(0,T)\times\partial\Omega,\\ u_{0}^{j}(0)&=\frac{|Y_{j}|u_{0,i}^{j}+|\Gamma|u_{0,i,\Gamma}^{j}}{|Y_{j}|+|\Gamma|}&\mbox{ in }&\Omega,\end{aligned} (12)

where the homogenized diffusion coefficient D^1n×n\widehat{D}^{1}\in\mathbb{R}^{n\times n} is defined by (i,l=1,,ni,l=1,\ldots,n)

(D^1)il:=Y1D1(\displaystyle\big{(}\widehat{D}^{1}\big{)}_{il}:=\int_{Y_{1}}D^{1}( ywi1+ei)(ywl1+el)dy\displaystyle\nabla_{y}w_{i}^{1}+e_{i})\cdot(\nabla_{y}w_{l}^{1}+e_{l})dy
+ΓDΓ1(Γ,ywi1+Γ,yyi)(Γ,ywl1+Γ,yyl)𝑑σ,\displaystyle+\int_{\Gamma}D_{\Gamma}^{1}(\nabla_{\Gamma,y}w_{i}^{1}+\nabla_{\Gamma,y}y_{i})\cdot(\nabla_{\Gamma,y}w_{l}^{1}+\nabla_{\Gamma,y}y_{l})d\sigma,

and wi11/w_{i}^{1}\in\mathbb{H}_{1}/\mathbb{R} (see Section 3.1 for the definition of this space) for i=1,,ni=1,\ldots,n are the solutions of the cell problems

y(D1(ywi1+ei))=0 in Y1,D1(ywi1+ei)ν=Γ,y(DΓ1(Γ,ywi1+Γ,yyi)) on Γ,wi1 is Y-periodic and Γwi1𝑑σ=0.\displaystyle\begin{aligned} -\nabla_{y}\cdot\big{(}D^{1}(\nabla_{y}w_{i}^{1}+e_{i})\big{)}&=0&\mbox{ in }Y_{1},\\ -D^{1}(\nabla_{y}w_{i}^{1}+e_{i})\cdot\nu&=-\nabla_{\Gamma,y}\cdot\big{(}D_{\Gamma}^{1}(\nabla_{\Gamma,y}w_{i}^{1}+\nabla_{\Gamma,y}y_{i})\big{)}&\mbox{ on }\Gamma,\\ w_{i}^{1}\mbox{ is }&Y\mbox{-periodic and }\int_{\Gamma}w_{i}^{1}d\sigma=0.\end{aligned} (13)

We say that u0=(u01,u02)u_{0}=(u_{0}^{1},u_{0}^{2}) is a weak solution of the macroscopic model, if

u01\displaystyle u_{0}^{1} L2((0,T),H1(Ω))H1((0,T),H1(Ω)),\displaystyle\in L^{2}((0,T),H^{1}(\Omega))\cap H^{1}((0,T),H^{1}(\Omega)^{\prime}),
u02\displaystyle u_{0}^{2} L2((0,T)×Ω)H1((0,T),L2(Ω)),\displaystyle\in L^{2}((0,T)\times\Omega)\cap H^{1}((0,T),L^{2}(\Omega)),

the equation for tu02\partial_{t}u_{0}^{2} in (12)\eqref{MacroscopicModel} is valid in L2((0,T)×Ω)L^{2}((0,T)\times\Omega), and for all ϕH1(Ω)\phi\in H^{1}(\Omega) it holds almost everywhere in (0,T)(0,T)

(|Y1|+|Γ|)tu01,ϕ\displaystyle(|Y_{1}|+|\Gamma|)\big{\langle}\partial_{t}u_{0}^{1},\phi H1(Ω),H1(Ω)+ΩD^1u01ϕdx\displaystyle\big{\rangle}_{H^{1}(\Omega)^{\prime},H^{1}(\Omega)}+\int_{\Omega}\widehat{D}^{1}\nabla u_{0}^{1}\cdot\nabla\phi dx
=ΩY1f1(y,u01)ϕ𝑑y𝑑x+ΩΓh1(y,u01,u02)ϕ𝑑σy𝑑x,\displaystyle=\int_{\Omega}\int_{Y_{1}}f^{1}(y,u_{0}^{1})\phi dydx+\int_{\Omega}\int_{\Gamma}h^{1}(y,u_{0}^{1},u_{0}^{2})\phi d\sigma_{y}dx,

together with the initial conditions from (12)\eqref{MacroscopicModel}.

Theorem 2.

The limit function u0=(u01,u02)u_{0}=(u_{0}^{1},u_{0}^{2}) from Proposition 5 is the unique solution of the macroscopic problem (12)\eqref{MacroscopicModel}.

Proof.

We illustrate the procedure for j=1j=1 (the case j=2j=2 follows by similar arguments, where the diffusion terms vanishes in the limit). As a test-function in (3)\eqref{VariationalEquationMicroscopicProblem} for j=1j=1 we choose ϕC0([0,T)×Ω¯)\phi\in C^{\infty}_{0}([0,T)\times\overline{\Omega}) and integrate with respect to time. By integration by parts in time we obtain

0T\displaystyle-\int_{0}^{T} Ωϵ1uϵ1tϕdxdtϵ0TΓϵuϵ1tϕdσdt\displaystyle\int_{\Omega_{\epsilon}^{1}}u_{\epsilon}^{1}\partial_{t}\phi dxdt-\epsilon\int_{0}^{T}\int_{\Gamma_{\epsilon}}u_{\epsilon}^{1}\partial_{t}\phi d\sigma dt
+0TΩϵ1Dϵ1uϵ1ϕdxdt+ϵ0TΓϵDΓϵ1Γϵuϵ1Γϵϕdσdt\displaystyle+\int_{0}^{T}\int_{\Omega_{\epsilon}^{1}}D_{\epsilon}^{1}\nabla u_{\epsilon}^{1}\cdot\nabla\phi dxdt+\epsilon\int_{0}^{T}\int_{\Gamma_{\epsilon}}D_{\Gamma_{\epsilon}}^{1}\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{1}\cdot\nabla_{\Gamma_{\epsilon}}\phi d\sigma dt
=\displaystyle= 0TΩϵ1fϵ1(uϵ1)ϕ𝑑x𝑑t+ϵ0TΓϵhϵ1(uϵ1,uϵ2)ϕ𝑑σ𝑑t\displaystyle\int_{0}^{T}\int_{\Omega_{\epsilon}^{1}}f_{\epsilon}^{1}(u_{\epsilon}^{1})\phi dxdt+\epsilon\int_{0}^{T}\int_{\Gamma_{\epsilon}}h_{\epsilon}^{1}(u_{\epsilon}^{1},u_{\epsilon}^{2})\phi d\sigma dt
+Ωϵ1uϵ,i1ϕ𝑑x+ϵΓϵuϵ,i,Γϵ1ϕ𝑑σ.\displaystyle+\int_{\Omega_{\epsilon}^{1}}u_{\epsilon,i}^{1}\phi dx+\epsilon\int_{\Gamma_{\epsilon}}u_{\epsilon,i,\Gamma_{\epsilon}}^{1}\phi d\sigma.

Using the convergence results from Proposition 5, Corollary 1, and Lemma 1, as well as the Assumption (A5) on the initial conditions, we obtain for ϵ0\epsilon\to 0

(|Y1|+|Γ|)\displaystyle-\big{(}|Y_{1}|+|\Gamma|\big{)} 0TΩu01tϕdxdt+0TΩD^1u01ϕdxdt\displaystyle\int_{0}^{T}\int_{\Omega}u_{0}^{1}\partial_{t}\phi dxdt+\int_{0}^{T}\int_{\Omega}\widehat{D}^{1}\nabla u_{0}^{1}\cdot\nabla\phi dxdt
=0TΩY1f1(u01)ϕ𝑑y𝑑x𝑑t+0TΩΓh1(u01,u02)ϕ𝑑σy𝑑x𝑑t\displaystyle=\int_{0}^{T}\int_{\Omega}\int_{Y_{1}}f^{1}(u_{0}^{1})\phi dydxdt+\int_{0}^{T}\int_{\Omega}\int_{\Gamma}h^{1}(u_{0}^{1},u_{0}^{2})\phi d\sigma_{y}dxdt
+Ω|Y1|u0,i1ϕ(0)𝑑x+Ω|Γ|u0,i,Γ1ϕ(0)𝑑x\displaystyle\hskip 20.00003pt+\int_{\Omega}|Y_{1}|u_{0,i}^{1}\phi(0)dx+\int_{\Omega}|\Gamma|u_{0,i,\Gamma}^{1}\phi(0)dx

Choosing ϕ\phi with compact support in (0,T)(0,T) we get tu01L2((0,T),H1(Ω))\partial_{t}u_{0}^{1}\in L^{2}((0,T),H^{1}(\Omega)^{\prime}) (see also Remark 3) with u01(0)=|Y1|u0,i1+|Γ|u0,i,Γ1|Y1|+|Γ|u_{0}^{1}(0)=\frac{|Y_{1}|u_{0,i}^{1}+|\Gamma|u_{0,i,\Gamma}^{1}}{|Y_{1}|+|\Gamma|}, and by density we obtain that u01u_{0}^{1} is a weak solution of the macroscopic equation for j=1j=1 in (12)\eqref{MacroscopicModel}. Uniqueness follows by standard energy estimates. ∎

Remark 3.

  1. (i)

    The regularity of the time-derivative is also directly obtained from the a priori estimates in Proposition 2. In fact, define for 0<h10<h\ll 1 and v:(0,T)Xv:(0,T)\rightarrow X for a Banach space XX the difference quotient for t(0,Th)t\in(0,T-h)

    thv(t):=v(t+h)v(t)h.\displaystyle\partial_{t}^{h}v(t):=\frac{v(t+h)-v(t)}{h}.

    Then for all ϕC0((0,T),C(Ω¯))\phi\in C^{\infty}_{0}((0,T),C^{\infty}(\overline{\Omega})) it holds, due to Proposition 5 and the a priori estimates for the time-derivative in Proposition 2:

    thu01,ϕ\displaystyle\langle\partial_{t}^{h}u_{0}^{1},\phi L2((0,Th),H1(Ω)),L2((0,Th),H1(Ω))=0ThΩthu01ϕdxdt\displaystyle\rangle_{L^{2}((0,T-h),H^{1}(\Omega)^{\prime}),L^{2}((0,T-h),H^{1}(\Omega))}=\int_{0}^{T-h}\int_{\Omega}\partial_{t}^{h}u_{0}^{1}\phi dxdt
    =limϵ01|Y1|+|Γ|(0ThΩϵ1thuϵ1ϕdx+ϵΓϵthuϵ1ϕdσdt)\displaystyle=\lim_{\epsilon\to 0}\frac{1}{|Y_{1}|+|\Gamma|}\left(\int_{0}^{T-h}\int_{\Omega_{\epsilon}^{1}}\partial_{t}^{h}u_{\epsilon}^{1}\phi dx+\epsilon\int_{\Gamma_{\epsilon}}\partial_{t}^{h}u_{\epsilon}^{1}\phi d\sigma dt\right)
    =limϵ01|Y1|+|Γ|0Tthuϵ1,ϕ1,ϵ,1,ϵ𝑑t.\displaystyle=\lim_{\epsilon\to 0}\frac{1}{|Y_{1}|+|\Gamma|}\int_{0}^{T}\langle\partial_{t}^{h}u_{\epsilon}^{1},\phi\rangle_{\mathbb{H}_{1,\epsilon}^{\prime},\mathbb{H}_{1,\epsilon}}dt.
    limϵ01|Y1|+|Γ|thuϵ1L2((0,T),1,ϵ)ϕL2((0,T),1,ϵ)\displaystyle\leq\lim_{\epsilon\to 0}\frac{1}{|Y_{1}|+|\Gamma|}\|\partial_{t}^{h}u_{\epsilon}^{1}\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon}^{\prime})}\|\phi\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon})}
    Climϵ0tuϵ1L2((0,T),1,ϵ)ϕL2((0,T),1,ϵ)\displaystyle\leq C\lim_{\epsilon\to 0}\|\partial_{t}u_{\epsilon}^{1}\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon}^{\prime})}\|\phi\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon})}
    Climϵ0ϕL2((0,T),1,ϵ)CϕL2((0,T),H1(Ω)),\displaystyle\leq C\lim_{\epsilon\to 0}\|\phi\|_{L^{2}((0,T),\mathbb{H}_{1,\epsilon})}\leq C\|\phi\|_{L^{2}((0,T),H^{1}(\Omega))},

    where at the end we used that PΓP_{\Gamma} is an orthogonal projection. By density and the reflexivity of L2((0,Th),H1(Ω))L^{2}((0,T-h),H^{1}(\Omega)) we obtain the boundedness

    thu01L2((0,Th),H1(Ω))C,\displaystyle\|\partial_{t}^{h}u_{0}^{1}\|_{L^{2}((0,T-h),H^{1}(\Omega)^{\prime})}\leq C,

    for a constant CC independent of hh. This implies tu01L2((0,T),H1(Ω))\partial_{t}u_{0}^{1}\in L^{2}((0,T),H^{1}(\Omega)^{\prime}). A similar argument implies tu02L2((0,T),H1(Ω))\partial_{t}u_{0}^{2}\in L^{2}((0,T),H^{1}(\Omega)^{\prime}). However, the limit equation for u02u_{0}^{2} even improves the regularity of tu02\partial_{t}u_{0}^{2}.

  2. (ii)

    We can also consider the case of a connected-connected porous medium (for n3n\geq 3 and a domain Ω\Omega which can be decomposed in microscopic cells, for example a rectangle with integer side length, and an additional boundary condition on Γϵ\partial\Gamma_{\epsilon} is needed). In this case both macroscopic solutions are described by a reaction-diffusion equation as for u01u_{0}^{1} in Theorem 2. The derivation of the macroscopic model for the connected-connected case even gets simpler, because we only need the a priori estimates from Proposition 2 and the convergence results for the connected domain in Section 4.1. The estimates for the shifts in Proposition 3 are no longer necessary.

  3. (iii)

    The results can be easily extended to systems, see [15] for more details.

6 Discussion

By the methods of two-scale convergence and the unfolding operator we derived a macroscopic model for a reaction-diffusion equation in a connected-disconnected porous medium with a nonlinear dynamic Wentzell-interface condition across the interface. The crucial point was to pass to the limit in the nonlinear terms, especially on the interface. Therefore, we established strong two-scale compactness results just depending on a priori estimates for the sequence of solutions. For the proof we used the unfolding operator and a Banach-valued Kolmogorov-Simon-compactness argument, which was necessarily for the disconnected domain. In fact, while the solutions in the connected domain Ωϵ1\Omega_{\epsilon}^{1} can be extended to the whole domain Ω\Omega preserving the a priori estimates, this is not possible anymore for the disconnected domain.

We emphasize that the strong compactness result in Theorem 1 is not restricted to our specific problem, but on the a priori estimates and the estimates for the shifts for the sequence. Therefore it can be easily applied to other problems. Especially, the results above can be extended to systems in an obvious way.

The time-derivative in the Wentzell-boundary condition on the interface Γϵ\Gamma_{\epsilon} regularizes the problem and leads to a simple variational structure with respect to the Gelfand-triple (j,ϵ,𝕃j,ϵ,j,ϵ)(\mathbb{H}_{j,\epsilon},\mathbb{L}_{j,\epsilon},\mathbb{H}_{j,\epsilon}^{\prime}), see (3)\eqref{VariationalEquationMicroscopicProblem}. Hence, the problem seems to be more complex regarding stationary interface conditions (neglecting the time-derivative). On the other hand, neglecting the diffusion term on the surface leads to an ordinary differential equation on the surface. Hence, we loose spatial regularity on the surface and therefore we have to replace the space j,ϵ\mathbb{H}_{j,\epsilon} by the function space {(uϵ,vϵ)H1(Ωϵj)×L2(Γϵ):uϵ|Γϵ=vϵ}\left\{(u_{\epsilon},v_{\epsilon})\in H^{1}(\Omega_{\epsilon}^{j})\times L^{2}(\Gamma_{\epsilon})\,:\,u_{\epsilon}|_{\Gamma_{\epsilon}}=v_{\epsilon}\right\} with norm uϵH1(Ωϵj)2+ϵvϵL2(Γϵ)2\sqrt{\|u_{\epsilon}\|_{H^{1}(\Omega_{\epsilon}^{j})}^{2}+\epsilon\|v_{\epsilon}\|_{L^{2}(\Gamma_{\epsilon})}^{2}}. For this choice it could be expected that the methods in the paper can be adapted to the case without surface diffusion. Nevertheless, both cases should be considered in more detail and are part of my ongoing work.

Acknowledgements

The author was supported by the Odysseus program of the Research Foundation - Flanders FWO (Project-Nr. G0G1316N) and the project SCIDATOS (Scientific Computing for Improved Detection and Therapy of Sepsis), which was funded by the Klaus Tschira Foundation, Germany (Grant number 00.0277.2015).

Appendix A Two-scale convergence and unfolding operator

We repeat the definition of the two-scale convergence and the unfolding operator and summarize some well known properties and compactness results.

A.1 Two-scale convergence

In the following, unless stated otherwise, we assume that p(1,)p\in(1,\infty) and pp^{\prime} is the dual exponent of pp. We start with the definition of the two-scale convergence, see [2, 28].

Definition 1.

We say the sequence uϵLp((0,T)×Ω)u_{\epsilon}\in L^{p}((0,T)\times\Omega) converges in the two-scale sense (in LpL^{p}) to a limit function u0Lp((0,T)×Y)u_{0}\in L^{p}((0,T)\times Y), if for all ϕLp((0,T)×Ω,Cper0(Y))\phi\in L^{p^{\prime}}((0,T)\times\Omega,C^{0}_{\mathrm{per}}(Y)) it holds that

limϵ00TΩuϵ(t,x)ϕ(t,x,xϵ)𝑑x𝑑t=0TΩYu0(t,x,y)ϕ(t,x,y)𝑑x𝑑y𝑑t.\displaystyle\lim_{\epsilon\to 0}\int_{0}^{T}\int_{\Omega}u_{\epsilon}(t,x)\phi\left(t,x,\frac{x}{\epsilon}\right)dxdt=\int_{0}^{T}\int_{\Omega}\int_{Y}u_{0}(t,x,y)\phi(t,x,y)dxdydt.

We say the sequence converges strongly in the two-scale sense (in LpL^{p}), if it holds that

limϵ0uϵLp((0,T)×Ω)=u0Lp((0,T)×Ω×Y).\displaystyle\lim_{\epsilon\to 0}\|u_{\epsilon}\|_{L^{p}((0,T)\times\Omega)}=\|u_{0}\|_{L^{p}((0,T)\times\Omega\times Y)}.
Remark 4.

  1. (i)

    For sequences in Lp((0,T)×Ωϵj)L^{p}((0,T)\times\Omega_{\epsilon}^{j}) on the perforated domain we also use the designation ”two-scale convergence”. The definition is also valid for such functions by extension by zero (or with the extension operator from [1]), and considering suitable test-functions.

  2. (ii)

    The two scale convergence introduced above should actually be referred to as ”weak two scale convergence”. However, in accordance with the definition in [2] we neglect the word ”weak” and only use ”strong” to highlight the ”strong two-scale convergence”.

  3. (iii)

    For the ”two-scale convergence in L2L^{2}” we just write ”two-scale convergence”.

Next, we give the definition of the two-scale convergence on oscillating surfaces, see [3, 26].

Definition 2.

We say the sequence uϵLp((0,T)×Γϵ)u_{\epsilon}\in L^{p}((0,T)\times\Gamma_{\epsilon}) converges in the two-scale sense (in LpL^{p}) to a limit function u0Lp((0,T)×Ω×Γ)u_{0}\in L^{p}((0,T)\times\Omega\times\Gamma), if for all ϕC0([0,T]×Ω¯,Cper0(Γ))\phi\in C^{0}([0,T]\times\overline{\Omega},C_{\mathrm{per}}^{0}(\Gamma)) it holds that

limϵ0ϵ0TΓϵuϵ(t,x)ϕ(t,x,xϵ)𝑑σx𝑑t=0TΩΓu0(t,x,y)ϕ(t,x,y)𝑑σy𝑑x𝑑t.\displaystyle\lim_{\epsilon\to 0}\epsilon\int_{0}^{T}\int_{\Gamma_{\epsilon}}u_{\epsilon}(t,x)\phi\left(t,x,\frac{x}{\epsilon}\right)d\sigma_{x}dt=\int_{0}^{T}\int_{\Omega}\int_{\Gamma}u_{0}(t,x,y)\phi(t,x,y)d\sigma_{y}dxdt.

We say the sequence converges strongly in the two-scale sense, if it holds that

limϵ0ϵ1puϵLp((0,T)×Γϵ)=u0Lp((0,T)×Ω×Γ).\displaystyle\lim_{\epsilon\to 0}\epsilon^{\frac{1}{p}}\|u_{\epsilon}\|_{L^{p}((0,T)\times\Gamma_{\epsilon})}=\|u_{0}\|_{L^{p}((0,T)\times\Omega\times\Gamma)}.

In accordance with Remark 4, we proceed analogously for the two-scale convergence on Γϵ\Gamma_{\epsilon} and neglect the word ”weak” and the addition ”L2L^{2}”.

To pass to the limit ϵ0\epsilon\to 0 in the diffusion terms in the bulk domain Ωϵj\Omega_{\epsilon}^{j} and the surface Γϵ\Gamma_{\epsilon} in the microscopic equation (3)\eqref{VariationalEquationMicroscopicProblem} we need compactness results for the spaces j,ϵ\mathbb{H}_{j,\epsilon}. In the following Lemma we summarize some weak two-scale compactness results for such functions, which can be found in [16]:

Lemma 5.

For j{1,2}j\in\{1,2\} let uϵjL2((0,T),j,ϵ)u_{\epsilon}^{j}\in L^{2}((0,T),\mathbb{H}_{j,\epsilon}) be a sequence with

uϵjL2((0,T),j,ϵ)C.\displaystyle\|u_{\epsilon}^{j}\|_{L^{2}((0,T),\mathbb{H}_{j,\epsilon})}\leq C.

Then it holds:

  1. (i)

    For j=1j=1 there exist u01L2((0,T),H1(Ω))u_{0}^{1}\in L^{2}((0,T),H^{1}(\Omega)) and a YY-periodic function u11L2((0,T)×Ω,1/)u_{1}^{1}\in L^{2}((0,T)\times\Omega,\mathbb{H}_{1}/\mathbb{R}), such that up to a subsequence

    uϵ1\displaystyle u_{\epsilon}^{1} u01\displaystyle\rightarrow u_{0}^{1} in the two-scale sense ,
    uϵ1\displaystyle\nabla u_{\epsilon}^{1} xu01+yu11\displaystyle\rightarrow\nabla_{x}u_{0}^{1}+\nabla_{y}u_{1}^{1} in the two-scale sense ,
    uϵ1|Γϵ\displaystyle u_{\epsilon}^{1}|_{\Gamma_{\epsilon}} u01\displaystyle\rightarrow u_{0}^{1} in the two-scale sense on Γϵ,\displaystyle\Gamma_{\epsilon},
    Γϵuϵ1|Γϵ\displaystyle\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{1}|_{\Gamma_{\epsilon}} PΓu01+Γu11|Γ\displaystyle\rightarrow P_{\Gamma}\nabla u_{0}^{1}+\nabla_{\Gamma}u_{1}^{1}|_{\Gamma} in the two-scale sense on Γϵ.\displaystyle\Gamma_{\epsilon}.
  2. (ii)

    For j=2j=2 there exist u02L2((0,T)×Ω)u_{0}^{2}\in L^{2}((0,T)\times\Omega) and u12L2((0,T)×Ω,2/)u_{1}^{2}\in L^{2}((0,T)\times\Omega,\mathbb{H}_{2}/\mathbb{R}) such that up to a subsequence

    uϵ2\displaystyle u_{\epsilon}^{2} u02\displaystyle\rightarrow u_{0}^{2} in the two-scale sense ,
    uϵ2\displaystyle\nabla u_{\epsilon}^{2} yu12\displaystyle\rightarrow\nabla_{y}u_{1}^{2} in the two-scale sense ,
    uϵ2|Γϵ\displaystyle u_{\epsilon}^{2}|_{\Gamma_{\epsilon}} u02\displaystyle\rightarrow u_{0}^{2} in the two-scale sense on Γϵ,\displaystyle\Gamma_{\epsilon},
    Γϵuϵ2|Γϵ\displaystyle\nabla_{\Gamma_{\epsilon}}u_{\epsilon}^{2}|_{\Gamma_{\epsilon}} Γu12\displaystyle\rightarrow\nabla_{\Gamma}u_{1}^{2} in the two-scale sense on Γϵ.\displaystyle\Gamma_{\epsilon}.

A.2 The unfolding operator

In the following we give the definition of the unfolding operator and summarize some well known properties, see the monograph [11] for an overview about this topic, and also [7, 8, 9, 10, 31]. In the following we consider the tuple (Gϵ,G){(Ω,Y),(Ωϵ1,Y1),(Ωϵ2,Y2),(Γϵ,Γ)}(G_{\epsilon},G)\in\{(\Omega,Y),(\Omega_{\epsilon}^{1},Y_{1}),(\Omega_{\epsilon}^{2},Y_{2}),(\Gamma_{\epsilon},\Gamma)\} and we define

G^ϵ:=intkKϵϵ(G¯+k),Λϵ:=ΩG^ϵ¯.\displaystyle\widehat{G}_{\epsilon}:=\mathrm{int}\bigcup_{k\in K_{\epsilon}}\epsilon\left(\overline{G}+k\right),\quad\quad\Lambda_{\epsilon}:=\Omega\setminus\overline{\widehat{G}_{\epsilon}}.

Then, for p(1,)p\in(1,\infty) we define the unfolding operator

𝒯ϵ:Lp((0,T)×Gϵ)Lp((0,T)×Ω×G),\displaystyle\mathcal{T}_{\epsilon}:L^{p}((0,T)\times G_{\epsilon})\rightarrow L^{p}((0,T)\times\Omega\times G),

with

𝒯ϵ(ϕϵ)(t,x,y):={ϕϵ(t,ϵ[xϵ]+ϵy) for xG^ϵ,0 for xΛϵ.\displaystyle\mathcal{T}_{\epsilon}(\phi_{\epsilon})(t,x,y):=\begin{cases}\phi_{\epsilon}\left(t,\epsilon\left[\frac{x}{\epsilon}\right]+\epsilon y\right)&\mbox{ for }x\in\widehat{G}_{\epsilon},\\ 0&\mbox{ for }x\in\Lambda_{\epsilon}.\end{cases}

We emphasize that we use the same notation for the unfolding operator for the different choices of the tuple (Gϵ,G)(G_{\epsilon},G). It should be clear from the context in which sense it has to be understood. Further, we mention that unfolding operator commutes with the trace operator in the following sense: For ϕϵLp((0,T),W1,p(Ωϵj))\phi_{\epsilon}\in L^{p}((0,T),W^{1,p}(\Omega_{\epsilon}^{j})) for j{1,2}j\in\{1,2\} it holds that

𝒯ϵ(ϕϵ|Γϵ)=(𝒯ϵ(ϕϵ))|Γ.\displaystyle\mathcal{T}_{\epsilon}\big{(}\phi_{\epsilon}|_{\Gamma_{\epsilon}}\big{)}=\big{(}\mathcal{T}_{\epsilon}(\phi_{\epsilon})\big{)}|_{\Gamma}.
Lemma 6.

  1. (a)

    For (Gϵ,G){(Ω,Y),(Ωϵ1,Y1),(Ωϵ2,Y2)}(G_{\epsilon},G)\in\{(\Omega,Y),(\Omega_{\epsilon}^{1},Y_{1}),(\Omega_{\epsilon}^{2},Y_{2})\} we have:

    1. (i)

      For ϕϵLp((0,T)×Gϵ)\phi_{\epsilon}\in L^{p}((0,T)\times G_{\epsilon}) it holds that

      𝒯ϵϕϵLp((0,T)×Ω×G)=ϕϵLp((0,T)×G^ϵ).\displaystyle\|\mathcal{T}_{\epsilon}\phi_{\epsilon}\|_{L^{p}((0,T)\times\Omega\times G)}=\|\phi_{\epsilon}\|_{L^{p}((0,T)\times\widehat{G}_{\epsilon})}.
    2. (ii)

      For ϕϵLp((0,T),W1,p(Gϵ))\phi_{\epsilon}\in L^{p}((0,T),W^{1,p}(G_{\epsilon})) it holds that

      y𝒯ϵϕϵ=ϵ𝒯ϵxϕϵ.\displaystyle\nabla_{y}\mathcal{T}_{\epsilon}\phi_{\epsilon}=\epsilon\mathcal{T}_{\epsilon}\nabla_{x}\phi_{\epsilon}.
  2. (b)

    For the unfolding operator on the surface we have:

    1. (i)

      For ϕϵLp((0,T)×Γϵ)\phi_{\epsilon}\in L^{p}((0,T)\times\Gamma_{\epsilon}) it holds that

      𝒯ϵϕϵLp((0,T)×Ω×Γ)=ϵ1pϕϵLp((0,T)×Γϵ).\displaystyle\|\mathcal{T}_{\epsilon}\phi_{\epsilon}\|_{L^{p}((0,T)\times\Omega\times\Gamma)}=\epsilon^{\frac{1}{p}}\|\phi_{\epsilon}\|_{L^{p}((0,T)\times\Gamma_{\epsilon})}.
    2. (ii)

      For ϕϵLp((0,T),W1,p(Γϵ))\phi_{\epsilon}\in L^{p}((0,T),W^{1,p}(\Gamma_{\epsilon})) it holds that

      Γ,y𝒯ϵϕϵ=ϵ𝒯ϵΓϵϕϵ.\displaystyle\nabla_{\Gamma,y}\mathcal{T}_{\epsilon}\phi_{\epsilon}=\epsilon\mathcal{T}_{\epsilon}\nabla_{\Gamma_{\epsilon}}\phi_{\epsilon}.
Proof.

For (a) and (b)(b)(i) see [11]. A proof for (b)(b)(ii) can be found in [22]. ∎

In the following Lemma we give an equivalent relation between the unfolding operator and the two-scale convergence. For a proof see for example [8, 9, 11].

Lemma 7.

Let p(1,)p\in(1,\infty).

  1. (a)

    For (Gϵ,G){(Ω,Y),(Ωϵ1,Y1),(Ωϵ2,Y2)}(G_{\epsilon},G)\in\{(\Omega,Y),(\Omega_{\epsilon}^{1},Y_{1}),(\Omega_{\epsilon}^{2},Y_{2})\} and a sequence uϵLp((0,T)×Gϵ)u_{\epsilon}\in L^{p}((0,T)\times G_{\epsilon}), the following statements are equivalent:

    1. (a)

      uϵu0u_{\epsilon}\rightarrow u_{0} weakly/strongly in the two-scale sense in LpL^{p},

    2. (b)

      𝒯ϵuϵu0\mathcal{T}_{\epsilon}u_{\epsilon}\rightarrow u_{0} weakly/strongly in Lp((0,T)×Ω×G)L^{p}((0,T)\times\Omega\times G).

  2. (b)

    For a sequence uϵLp((0,T)×Γϵ)u_{\epsilon}\in L^{p}((0,T)\times\Gamma_{\epsilon}) with ϵ1puϵLp((0,T)×Γϵ)C\epsilon^{\frac{1}{p}}\|u_{\epsilon}\|_{L^{p}((0,T)\times\Gamma_{\epsilon})}\leq C, the following statements are equivalent:

    1. (a)

      uϵu0u_{\epsilon}\rightarrow u_{0} weakly/strongly in the two-scale sense on Γϵ\Gamma_{\epsilon} in LpL^{p},

    2. (b)

      𝒯ϵuϵu0\mathcal{T}_{\epsilon}u_{\epsilon}\rightarrow u_{0} weakly/strongly in Lp((0,T)×Ω×Γ)L^{p}((0,T)\times\Omega\times\Gamma).

References

  • [1] E. Acerbi, V. Chiadò, G. D. Maso, and D. Percivale. An extension theorem from connected sets, and homogenization in general periodic domains. Nonlinear Analysis, Theory, Methods & Applications, 18(5):481–496, 1992.
  • [2] G. Allaire. Homogenization and two-scale convergence. SIAM J. Math. Anal., 23:1482–1518, 1992.
  • [3] G. Allaire, A. Damlamian, and U. Hornung. Two-scale convergence on periodic surfaces and applications. in Proceedings of the International Conference on Mathematical Modelling of Flow Through Porous Media, A. Bourgeat et al., eds., World Scientific, Singapore, pages 15–25, 1996.
  • [4] G. Allaire and H. Hutridurga. Homogenization of reactive flows in porous media and competition between bulk and surface diffusion. IMA Journal of Applied Mathematics, 77:788–215, 2012.
  • [5] M. Amar and R. Gianni. Laplace-Beltrami operator for the heat conduction in polymer coating of electronic devices. Discrete Cont. Dyn. Sys. B, 23(4):1739–1756, 2018.
  • [6] M. Anguiano. Existence, uniqueness and homogenization of nonlinear parabolic problems with dynamical boundary conditions in perforated media. Mediterr. J. Math., 2020.
  • [7] T. Arbogast, J. Douglas, and U. Hornung. Derivation of the double porosity model of single phase flow via homogenization theory. SIAM J. Math. Anal., 27:823–836, 1990.
  • [8] A. Bourgeat, S. Luckhaus, and A. Mikelić. Convergence of the homogenization process for a double-porosity model of immiscible two-phase flow. SIAM J. Math. Anal., 27:1520–1543, 1996.
  • [9] D. Cioranescu, A. Damlamian, P. Donato, G. Griso, and R. Zaki. The periodic unfolding method in domains with holes. SIAM J. Math. Anal., 44(2):718–760, 2012.
  • [10] D. Cioranescu, A. Damlamian, and G. Griso. The periodic unfolding method in homogenization. SIAM J. Math. Anal., 40:1585–1620, 2008.
  • [11] D. Cioranescu, G. Griso, and A. Damlamian. The periodic unfolding method. Springer, 2018.
  • [12] D. Cioranescu and J. S. J. Paulin. Homogenization in open sets with holes. J. Math. Pures et Appl., 71:590–607, 1979.
  • [13] P. Donato and K. H. L. Nguyen. Homogenization of diffusion problems with a nonlinear interfacial resistance. NoDEA Nonlinear Differential Equations and Appl., 22(5):1345–1380, 2015.
  • [14] G. Friesecke, R. D. James, and S. Müller. A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 55(11):1461–1506, 2002.
  • [15] M. Gahn. Derivation of effective models for reaction-diffusion processes in multi-component media. PhD thesis (Friedrich-Alexander-Universität Erlangen-Nürnberg); Shaker Verlag, 2017.
  • [16] M. Gahn. Multi-scale modeling of processes in porous media - coupling reaction-diffusion processes in the solid and the fluid phase and on the separating interfaces. Discrete & Continuous Dynamical Systems Series B, 24(12):6511–6531, 2019.
  • [17] M. Gahn and M. Neuss-Radu. A characterization of relatively compact sets in Lp(Ω,B)L^{p}(\Omega,B). Stud. Univ. Babeş-Bolyai Math., 61(3):279–290, 2016.
  • [18] M. Gahn, M. Neuss-Radu, and P. Knabner. Derivation of an effective model for metabolic processes in living cells including substrate channeling. Vietnam J. Math., 45(1–2):265–293, 2016.
  • [19] M. Gahn, M. Neuss-Radu, and P. Knabner. Homogenization of reaction–diffusion processes in a two-component porous medium with nonlinear flux conditions at the interface. SIAM J. Appl. Math., 76(5):1819–1843, 2016.
  • [20] G. P. Galdi. An Introduction to the Mathematical Theory of the Navier–Stokes Equations. Springer Monogr. in Math., Springer-Verlag, New York, 2011.
  • [21] I. Graf and M. A. Peter. A convergence result for the periodic unfolding method related to fast diffusion on manifolds. C. R. Acad. Sci. Paris, Ser. I, 352(6):485–490, 2014.
  • [22] I. Graf and M. A. Peter. Diffusion on surfaces and the boundary periodic unfolding operator with an application to carcinogenesis in human cells. SIAM J. Math. Anal., 46(4):3025–3049, 2014.
  • [23] E. Hewitt and K. Stromberg. Real and Abstract Analysis. Springer-Verlag, New York, 1975.
  • [24] U. Hornung, W. Jäger, and A. Mikelić. Reactive transport through an array of cells with semi-permeable membranes. ESAIM Mathematical Model. and Numer. Anal., 28:59–94, 1994.
  • [25] A. Meirmanov and R. Zimin. Compactness result for periodic structures and its application to the homogenization of a diffusion-convection equation. Elec. J. of Diff. Equat., 115:1–11, 2011.
  • [26] M. Neuss-Radu. Some extensions of two-scale convergence. C. R. Acad. Sci. Paris Sér. I Math., 322:899–904, 1996.
  • [27] M. Neuss-Radu and W. Jäger. Effective transmission conditions for reaction-diffusion processes in domains separated by an interface. SIAM J. Math. Anal., 39:687–720, 2007.
  • [28] G. Nguetseng. A general convergence result for a functional related to the theory of homogenization. SIAM J. Math. Anal., 20:608–623, 1989.
  • [29] M. Ptashnyk and T. Roose. Derivation of a macroscopic model for transport of strongly sorbed solutes in the soil using homogenization theory. SIAM J. Appl. Math., 70:2097–2118, 2010.
  • [30] J. Simon. Compact sets in the space Lp(0,T;B){L}^{p}(0,{T};{B}). Ann. Mat. Pura Appl., 146:65–96, 1987.
  • [31] C. Vogt. A homogenization theorem leading to a volterra integro-differential equation for permeation chromatography. SFB 123, University of Heidelberg, Preprint 155 and Diploma-thesis, 1982.
  • [32] B. S. Winkel. Metabolic channeling in plants. Annu. Rev. Plant Biol., 55:85–107, 2004.