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

[intoc]

Elliptic Equations in Weak Oscillatory Thin Domains: Beyond Periodicity with Boundary-Concentrated Reaction Terms

Pricila S. Barbosa e-mail: [email protected] Departamento de Matemática, Universidade Tecnológica Federal do Paraná, Londrina, Brazil Manuel Villanueva-Pesqueira e-mail: [email protected]. Partially supported by PID2022-137074NB-I00 funded by MCIN Grupo de Dinámica No Lineal, Universidad Pontificia Comillas, ICAI, Alberto Aguilera 25 28015 Madrid, Spain
Abstract

In this paper we analyze the limit behavior of a family of solutions of the Laplace operator with homogeneous Neumann boundary conditions, set in a two-dimensional thin domain which presents weak oscillations on both boundaries and with terms concentrated in a narrow oscillating neighborhood of the top boundary. The aim of this problem is to study the behavior of the solutions as the thin domain presents oscillatory behaviors beyond the classical periodic assumptions,including scenarios like quasi-periodic or almost-periodic oscillations. We then prove that the family of solutions converges to the solution of a 11-dimensional limit equation capturing the geometry and oscillatory behavior of boundary of the domain and the narrow strip where the concentration terms take place. In addition, we include a series of numerical experiments illustrating the theoretical results obtained in the quasi-periodic context.

Keywords: Thin domains; oscillatory boundary; homogenization; quasi-periodic; almost periodic; concentrating terms.

1 Introduction

We are keen on examining the behavior of the solutions to certain Elliptic Partial Differential Equations which are posed in a oscillating thin domain RϵR^{\epsilon}. This domain is a slender region in 2\mathbb{R}^{2} displaying oscillations along its boundary. It is described as the region between two oscillatory functions, that is,

(1) Rϵ={(x,y)2|xI,ϵk1(x,ϵ)<y<ϵk2(x,ϵ)},R^{\epsilon}=\Big{\{}(x,y)\in\mathbb{R}^{2}\;|\;x\in I\subset\mathbb{R},\;-\epsilon k^{1}(x,\epsilon)<y<\epsilon k^{2}(x,\epsilon)\Big{\}},

where I=[a,b]I=[a,b] is any interval in \mathbb{R} and the functions k1,k2k^{1},k^{2} satisfy certain hypotheses, see (H) in the next section. These two functions may present periodic oscillation, see Figure 1 for a example, but also more intricate situations beyond the classical periodic setting, see Figure 2 for an example with quasi-periodic oscillating boundaries.

Refer to caption
Figure 1: Thin domain RϵR^{\epsilon} with doubly periodic weak oscillatory boundary.
Refer to caption
Figure 2: Thin domain RϵR^{\epsilon} with doubly quasi-periodic oscillatory boundary.

We are interested in analyzing the following linear elliptic equation with reaction terms concentrated in a very narrow oscillating neighborhood of the upper oscillating boundary

(2) {Δwϵ+wϵ=1ϵγχθϵfϵ in Rϵ,wϵνϵ=0 on Rϵ,\left\{\begin{gathered}-\Delta w^{\epsilon}+w^{\epsilon}=\frac{1}{\epsilon^{\gamma}}\chi_{\theta^{\epsilon}}f^{\epsilon}\quad\textrm{ in }R^{\epsilon},\\ \frac{\partial w^{\epsilon}}{\partial\nu^{\epsilon}}=0\quad\textrm{ on }\partial R^{\epsilon},\end{gathered}\right.

where fϵL2(Rϵ)f^{\epsilon}\in L^{2}(R^{\epsilon}), and νϵ\nu^{\epsilon} is the unit outward normal to Rϵ\partial R^{\epsilon}. Let be χθϵ\chi_{\theta^{\epsilon}} the characteristic function of the narrow strip defined by

(3) θϵ={(x,y)2|xI,ϵ[k2(x,ϵ)ϵγHϵ(x)]<y<ϵk2(x,ϵ)},\theta^{\epsilon}=\Big{\{}(x,y)\in\mathbb{R}^{2}\;|\;x\in I\subset\mathbb{R},\;\epsilon[k^{2}(x,\epsilon)-\epsilon^{\gamma}H_{\epsilon}(x)]<y<\epsilon k^{2}(x,\epsilon)\Big{\}},

where γ>0\gamma>0 and Hϵ:IH_{\epsilon}:I\subset\mathbb{R}\to\mathbb{R} is a C1C^{1} quasi-periodic and bounded function.

It is important to highlight that we will obtain the limiting problem explicitly, assuming very general conditions on oscillatory functions. Afterward, we will study this limiting problem in interesting specific cases, such as quasi-periodic and even almost periodic functions. The authors firmly believe that analyzing problems in homogenization beyond the classical periodic setting is compelling from both an applied and a purely mathematical standpoint. In the real world, many materials and structures defy perfect periodicity and many real world problems involve multiple scales of heterogeneity or complexity.

Homogenization, a powerful mathematical tool in the realm of materials science and engineering, has traditionally been confined to the study of materials with perfect periodic structures, see [1, 13, 18, 19]. However, a growing body of research suggests that venturing beyond this classical periodic setting offers a rich landscape of possibilities, both in terms of practical applications and as a playground for mathematical exploration. Therefore, over the years, different homogenization techniques have been adapted and created to study problems beyond the periodic setting. In this context, the almost periodic framework offers an appropriate answer to certain limitations presented by periodic models. To illustrate how this topic is at the forefront of research, we can cite several articles, knowing that there are many more, arranged from the most classic to the most recent, see [2, 16, 23, 25, 28].

In the literature, there exists a multitude of studies dedicated to exploring the impact that oscillatory boundaries and thickness of the domain have on the solution behaviors of partial differential equations posed in thin domains. For further insight into this subject, one may refer to the following selected references and the studies cited within them [4, 5, 10, 20, 27]. However, the volume of research that avoids the consideration of periodic boundaries is significantly less. In this regard, it is worth mentioning some articles have dealt with bounded domains with locally periodic boundaries, see [9, 12]. Equations set in thin domains with the upper and lower boundaries oscillating at different periods have also been studied. This means that the requirements for traditional homogenization are not met, as there is no cell that describes the domain. In fact, the behavior of the two boundaries is reflected in the limit. While [11] focuses on boundaries with strong oscillations, [3] explores weak oscillations, yielding results characteristic of quasi-periodic functions

On the other hand, numerous studies have also focused on singular elliptic and parabolic problems, characterized by potential and reactive terms that are localized within a narrow region close of the boundary. First, equations in fixed and bounded domains were studied, see [6, 21, 22]. Recently, problems combining both singular situations have been studied. That is, thin domains with oscillating boundaries with terms concentrated on the boundary, [7, 8, 24]. However, all of these fall within the assumptions of periodicity.

The paper is organized as follows: In Section 2, we establish the necessary hypotheses, define the notation, and present the central result of the article, which reveals the explicit limit problem under general hypotheses about the oscillating boundaries. In Section 3, adapting the technique introduced in [3], we will show that after appropiate change of variables the solutions of the problem (2) can be aproximated by a one-dimensional problem with oscillating coefficients and with a force term reflecting the reaction terms on the thin neighborhood of the upper boundary. In Section 4, we proceed to analyze the limit of the one-dimensional equation derived in the previous section, thereby proving the main result of the paper. In Section 5, we derive the limit problem for specific cases of oscillating functions, including quasi-periodic and almost-periodic types. Finally, Section 6, we show some numerical evidences about the proved results.


2 Assumptions, notations and main result

Let kik^{i} be a function, i=1,2i=1,2,

ki:I×(0,1)+(x,ϵ)ki(x,ϵ)=kϵi(x),\begin{array}[]{rccl}k^{i}:&I\times(0,1)&\longrightarrow&\mathbb{R}^{+}\\ &(x,\epsilon)&\longrightarrow&k^{i}(x,\epsilon)=k_{\epsilon}^{i}(x),\end{array}

such that

  • (H.1)

    kϵik_{\epsilon}^{i} is a C1C^{1} function and

    (4) ϵ|ddxkϵi(x)|ϵ00 uniformly in I.\epsilon\Big{|}\frac{d}{dx}k^{i}_{\epsilon}(x)\Big{|}\buildrel\epsilon\to 0\over{\longrightarrow}0\hbox{ uniformly in }I.
  • (H.2)

    There exist two positive constants independent of ϵ\epsilon such that

    (5) 0<C1ikϵi()C2i.0<C^{i}_{1}\leq k_{\epsilon}^{i}(\cdot)\leq C^{i}_{2}.
  • (H.3)

    There exists a function KiK^{i} in L2(I)L^{2}(I) such that

    kϵiϵ0Ki wL2(I).k^{i}_{\epsilon}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}K^{i}\quad\hbox{ w}-L^{2}(I).
  • (H.4)

    There exists a function PP in L2(I)L^{2}(I) such that

    1kϵ1+kϵ2ϵ0P wL2(I).\frac{1}{k^{1}_{\epsilon}+k^{2}_{\epsilon}}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}P\quad\hbox{ w}-L^{2}(I).

Indeed, the thickness of the domain has order ϵ\epsilon and we say that the domain presents weak oscillations due to the convergence (4).

We will assume that the function which defines the narrow strip is a quasi-periodic smooth function. There are constants H0,H10H_{0},H_{1}\geq 0 such that

(6) H0Hϵ(x)H1for allxIandϵ>0.H_{0}\leq H_{\epsilon}(x)\leq H_{1}\,\,\mbox{for all}\,\,x\in I\,\,\mbox{and}\,\,\epsilon>0.

First, we present the homogenized problem for the general case of functions satisfying the hypotheses (H). Subsequently, we direct our attention to several intriguing examples that provide insights into this matter. In particular, our overarching framework allows us consider quasi-periodic or almost periodic functions. See Figure 2 where quasi-periodic functions are considered. That is the case where

(7) kϵ1(x)=h(x/ϵα),kϵ2(x)=g(x/ϵβ),k^{1}_{\epsilon}(x)=h(x/\epsilon^{\alpha}),\quad k^{2}_{\epsilon}(x)=g(x/\epsilon^{\beta}),

with 0<α,β<10<\alpha,\beta<1 and the functions g,h:g,h\,:\mathbb{R}\to\mathbb{R} are C1C^{1} quasi-periodic functions. Note that the extensively studied structure known as the periodic setting is included within this framework.

The variational formulation of (2) is the following: find wϵH1(Rϵ)w^{\epsilon}\in H^{1}(R^{\epsilon}) such that

Rϵ{wϵxφx+wϵyφy+wϵφ}𝑑x𝑑y=1ϵγθϵfϵφ𝑑x𝑑y,φH1(Rϵ).\int_{R^{\epsilon}}\Big{\{}\displaystyle\frac{\partial w^{\epsilon}}{\partial x}\displaystyle\frac{\partial\varphi}{\partial x}+\displaystyle\frac{\partial w^{\epsilon}}{\partial y}\displaystyle\frac{\partial\varphi}{\partial y}+w^{\epsilon}\varphi\Big{\}}dxdy=\frac{1}{\epsilon^{\gamma}}\int_{\theta^{\epsilon}}f^{\epsilon}\varphi dxdy,\,\forall\varphi\in H^{1}(R^{\epsilon}).

Observe that, for fixed ϵ>0\epsilon>0, the existence and uniqueness of solution to problem (2) is guaranteed by Lax-Milgram Theorem. Then, we will analyze the behavior of the solutions as the parameter ϵ\epsilon tends to zero. Particularly, by adapting the procedure demonstrated in [3], we initially implement a suitable change of variables. This enables us to substitute, in a certain sense, the original problem (2) posed in a 2-oscillating thin domain into a simpler problem with oscillating coefficients posed in an interval of \mathbb{R}. Notice that, this fact is in agreement with the intuitive idea that the family of solutions wϵw^{\epsilon} should converge to a function of just one variable as ϵ\epsilon goes to zero since the domain shrinks in the vertical direction. Subsequently, by using the previous results and adapting well-known techniques in homogenization we obtain explicitly the homogenized limit problem for the general case.

Due to the order of the height of the thin domains RϵR^{\epsilon} it makes sense to consider the following measure in thin domains

ρϵ(𝒪)=1ϵ|𝒪|,𝒪Rϵ.\rho_{\epsilon}(\mathcal{O})=\frac{1}{\epsilon}|\mathcal{O}|,\;\forall\,\mathcal{O}\subset R^{\epsilon}.

The rescaled Lebesgue measure ρϵ\rho_{\epsilon} allows us to preserve the relative capacity of a measurable subset 𝒪Rϵ\mathcal{O}\subset R^{\epsilon}. Moreover, using the previous measure we introduce the spaces Lp(Rϵ,ρϵ)L^{p}(R^{\epsilon},\rho_{\epsilon}) and W1,p(Rϵ,ρϵ)W^{1,p}(R^{\epsilon},\rho_{\epsilon}), for 1p<1\leq p<\infty endowed with the norms obtained rescaling the usual norms by the factor 1ϵ\frac{1}{\epsilon}, that is,

|φ|Lp(Rϵ)\displaystyle|||\varphi|||_{L^{p}(R^{\epsilon})} =ϵ1/pφLp(Rϵ),φLp(Rϵ),\displaystyle=\epsilon^{-1/p}||\varphi||_{L^{p}(R^{\epsilon})},\quad\forall\varphi\in L^{p}(R^{\epsilon}),
|φ|W1,p(Rϵ)\displaystyle|||\varphi|||_{W^{1,p}(R^{\epsilon})} =ϵ1/pφW1,p(Rϵ),φW1,p(Rϵ).\displaystyle=\epsilon^{-1/p}||\varphi||_{W^{1,p}(R^{\epsilon})},\quad\forall\varphi\in W^{1,p}(R^{\epsilon}).

It is very common to consider this kind of norms in works involving thin domains, see e.g. [20, 27, 26].

Then, assuming that

(8) 1ϵγ+1ϵkϵ1(x)ϵkϵ2(x)χθϵ(x,y)fϵ(x,y)𝑑yϵ0f0(x)wL2(I),\frac{1}{\epsilon^{\gamma+1}}\int_{-\epsilon k^{1}_{\epsilon}(x)}^{\epsilon k^{2}_{\epsilon}(x)}\chi_{\theta^{\epsilon}}(x,y)f^{\epsilon}(x,y)dy\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}f_{0}(x)\quad\hbox{w}-L^{2}(I),

the main result obtained is the following:

Theorem 2.1.

Let wϵw^{\epsilon} be the solution of problem (2). Given the function PP introduced in hypothesis (H.4), the definition of f0f_{0} as provided in (8), and denoting f^=f0K1+K2\hat{f}=\displaystyle\frac{f_{0}}{K^{1}+K^{2}}, then we have

wϵw^, wH1(I),w^{\epsilon}\to\hat{w},\hbox{ w}-H^{1}(I),
|wϵw^|L2(Rϵ)0,|||w^{\epsilon}-\hat{w}|||_{L^{2}(R^{\epsilon})}\to 0,

where w^H1(I)\hat{w}\in H^{1}(I) is the weak solution of the following Neumann problem

{1P(K1+K2)w^xx+w^=f^,xI,w^x(a)=w^x(b)=0.\left\{\begin{gathered}-\frac{1}{P(K^{1}+K^{2})}{\hat{w}}_{xx}+\hat{w}=\hat{f},\quad x\in I,\\ \hat{w}_{x}(a)=\hat{w}_{x}(b)=0.\end{gathered}\right.

Finally, we will get the boundary limiting problem for interesting particular cases such as quasi-periodicity or almost periodic framework.

3 Reduction to an one-dimensional problem with oscillating coefficients

In this section, we focus on the study of the elliptic problem (2) and begin by performing a change of variables that simplifies the domain in which the original problem is posed. As a matter of fact, we will be able to reduce the study of (2) in the thin domain RϵR^{\epsilon} to the study of an elliptic problem with oscillating coefficients capturing the geometry and oscillatory behavior of the open sets where the concentrations take place in the lower dimensional fixed domain II. This dimension reduction will be the key point to obtain the correct limiting equation. In order to state the main result of this section, let us first make some definitions. We will denote by

(9) ηi(ϵ)=supxI|ϵddxkϵi(x)|>0,i=1,2 and η(ϵ)=η1(ϵ)+η2(ϵ).\displaystyle\eta^{i}(\epsilon)=\sup_{x\in I}\Big{|}\epsilon\frac{d}{dx}k^{i}_{\epsilon}(x)\Big{|}>0,\quad i=1,2\quad\hbox{ and }\quad\eta(\epsilon)=\eta^{1}(\epsilon)+\eta^{2}(\epsilon).

Observe that from hypothesis (H.1) we have η(ϵ)ϵ00\eta(\epsilon)\buildrel\epsilon\to 0\over{\longrightarrow}0

Also, we denote by Kϵ(x)=kϵ1(x)+kϵ2(x)K_{\epsilon}(x)=k^{1}_{\epsilon}(x)+k_{\epsilon}^{2}(x) (that is ϵKϵ(x)\epsilon K_{\epsilon}(x) is the thickness of the thin domain RϵR^{\epsilon} at the point xIx\in I), so from (5) there exist constants K0,K1>0K_{0},K_{1}>0 independent of ϵ>0\epsilon>0 such that

(10) 0<K0Kϵ(x)K1for allI.0<K_{0}\leq K_{\epsilon}(x)\leq K_{1}\,\,\mbox{for all}\,\,I\subset\mathbb{R}.

Now we will consider the following one dimensional problem

(11) {1Kϵ(x)(Kϵ(x)w^xϵ)x+w^ϵ=1ϵγf^ϵ in I,w^xϵ(a)=w^xϵ(b)=0\left\{\begin{gathered}-\frac{1}{K_{\epsilon}(x)}(K_{\epsilon}(x)\hat{w}^{\epsilon}_{x})_{x}+\hat{w}^{\epsilon}=\frac{1}{\epsilon^{\gamma}}\hat{f}^{\epsilon}\quad\textrm{ in }I,\\ \hat{w}^{\epsilon}_{x}(a)=\hat{w}^{\epsilon}_{x}(b)=0\end{gathered}\right.

where

(12) f^ϵ(x)=1ϵKϵ(x)ϵkϵ1(x)ϵkϵ2(x)χθϵ(x,y)fϵ(x,y)𝑑y.\displaystyle\hat{f}^{\epsilon}(x)=\frac{1}{\epsilon K_{\epsilon}(x)}\int_{-\epsilon k^{1}_{\epsilon}(x)}^{\epsilon k^{2}_{\epsilon}(x)}\chi_{\theta^{\epsilon}}(x,y)f^{\epsilon}(x,y)dy.

The key result of this section is the following

Theorem 3.1.

Let wϵw^{\epsilon} and w^ϵ\hat{w}^{\epsilon} be the solutions of problems (2) and (11) respectively. Then, we have

|wϵw^ϵ|H1(Rϵ)ϵ00.|||w^{\epsilon}-\hat{w}^{\epsilon}|||_{H^{1}(R^{\epsilon})}\buildrel\epsilon\to 0\over{\longrightarrow}0.

In order to prove this result, we will need to obtain first some preliminary lemmas. We start transforming equation (2) into an equation in the modified thin domain

(13) Raϵ={(x¯,y¯)2|x¯I, 0<y¯<ϵKϵ(x¯)}.R_{a}^{\epsilon}=\Big{\{}(\bar{x},\bar{y})\in\mathbb{R}^{2}\;|\;\bar{x}\in I,\;0<\bar{y}<\epsilon\,K_{\epsilon}(\bar{x})\Big{\}}.

(see Figure 3) where it can be seen that we have transformed the oscillations of both boundaries into oscillations of just the boundary at the top. For this, we considering the following family of diffeomorphisms

Lϵ:RaϵRϵ(x¯,y¯)(x,y):=(x¯,y¯ϵkϵ1(x¯)).\begin{array}[]{rl}L^{\epsilon}:R_{a}^{\epsilon}&\longrightarrow R^{\epsilon}\\ (\bar{x},\bar{y})&\longrightarrow(x,y):=(\bar{x},\,\bar{y}\,-\,\epsilon\,k^{1}_{\epsilon}(\bar{x})).\end{array}

Notice that the inverse of this diffeomorphism is (Lϵ)1(x,y)=(x,y+ϵkϵ1(x))(L^{\epsilon})^{-1}(x,y)=(x,\,y\,+\,\epsilon\,k^{1}_{\epsilon}(x)). Moreover, from the structure of these diffeomorphisms and hypothesis (H.1) we easily get that there exists a constant CC such that the Jacobian Matrix of LϵL^{\epsilon} and (Lϵ)1(L^{\epsilon})^{-1} satisfy

(14) JLϵL,J(Lϵ)1LC.\|JL^{\epsilon}\|_{L^{\infty}},\|J(L^{\epsilon})^{-1}\|_{L^{\infty}}\leq C.

Moreover, we also have det(JLϵ)(x¯,y¯)=det(J(Lϵ)1)(x,y)=1det(JL^{\epsilon})(\bar{x},\bar{y})=det(J(L^{\epsilon})^{-1})(x,y)=1.

We will show that the study of the limit behavior of the solutions of (2) is equivalent to analyze the behavior of the solutions of the following problem

(15) {Δvϵ+vϵ=1ϵγχθ1ϵf1ϵ in Raϵ,vϵνϵ=0 on Raϵ,\left\{\begin{gathered}-\Delta v^{\epsilon}+v^{\epsilon}=\frac{1}{\epsilon^{\gamma}}\chi_{\theta^{\epsilon}_{1}}f_{1}^{\epsilon}\quad\textrm{ in }R_{a}^{\epsilon},\\ \frac{\partial v^{\epsilon}}{\partial\nu^{\epsilon}}=0\quad\textrm{ on }\partial R_{a}^{\epsilon},\end{gathered}\right.

where the function χθ1ϵ:2\chi_{\theta^{\epsilon}_{1}}:\mathbb{R}^{2}\rightarrow\mathbb{R} is the characteristic function of the narrow strip θ1ϵ\theta^{\epsilon}_{1} given by

(16) θ1ϵ={(x¯,y¯)2|x¯I,ϵ[Kϵ(x¯)ϵγHϵ(x¯)]<y¯<ϵKϵ(x¯)}.\theta^{\epsilon}_{1}=\Big{\{}(\bar{x},\bar{y})\in\mathbb{R}^{2}\;|\;\bar{x}\in I\subset\mathbb{R},\;\epsilon[K_{\epsilon}(\bar{x})-\epsilon^{\gamma}H_{\epsilon}(\bar{x})]<\bar{y}<\epsilon K_{\epsilon}(\bar{x})\Big{\}}.

The vector νϵ\nu^{\epsilon} is the outward unit normal to Raϵ\partial R_{a}^{\epsilon} and

(17) f1ϵ=fϵLϵ.f_{1}^{\epsilon}=f^{\epsilon}\circ L^{\epsilon}.
Refer to caption
Figure 3: Thin domain RaϵR_{a}^{\epsilon} obtained from RϵR^{\epsilon}.
Remark 1.

Notice that

(18) f1ϵL2(θ1ϵ)2=θ1ϵ|fϵLϵ(X¯)|2dX¯=θϵ|fϵ(X)|2|det(J(Lϵ)1))(X)|dX=fϵL2(θϵ)2,\|f_{1}^{\epsilon}\|_{L^{2}(\theta^{\epsilon}_{1})}^{2}=\int_{\theta^{\epsilon}_{1}}|f^{\epsilon}\circ L^{\epsilon}(\bar{X})|^{2}d\bar{X}=\int_{\theta^{\epsilon}}|f^{\epsilon}(X)|^{2}|det(J(L^{\epsilon})^{-1}))(X)|dX=\|f^{\epsilon}\|_{L^{2}(\theta^{\epsilon})}^{2},

where X¯=(x¯,y¯)\bar{X}=(\bar{x},\bar{y}) and X=(x,y)X=(x,y). Therefore, |f1ϵ|L2(θ1ϵ)=|fϵ|L2(θϵ)|||f_{1}^{\epsilon}|||_{L^{2}(\theta^{\epsilon}_{1})}=|||f^{\epsilon}|||_{L^{2}(\theta^{\epsilon})} for all fϵL2(Rϵ)f^{\epsilon}\in L^{2}(R^{\epsilon}).

To achieve a better understanding of the behavior from the nonlinear concentrated term, we will analize the concentrating integral. We are now interested in analyzing concentrated integrals on the narrow strips θ1ϵ\theta_{1}^{\epsilon}, to this end we will adapt results from [8] at Section 3. The next Lemma is about concentrating integrals, some estimates will be given in different Lebesgue and Sobolev-Bochner spaces.

Lemma 3.2.

Let RaϵR_{a}^{\epsilon} and θ1ϵ\theta^{\epsilon}_{1} defined by (13) and (16) respectively. If vϵW1,p(Raϵ)v^{\epsilon}\in W^{1,p}(R_{a}^{\epsilon}), there exists C>0C>0 independent of ϵ\epsilon such that, for 11p<s11-\frac{1}{p}<s\leq 1,

(19) 1ϵγθ1ϵ|vϵ|q𝑑X¯CvϵLq(I;Ws,p(0,ϵKϵ(x¯)))q,q1.\frac{1}{\epsilon^{\gamma}}\int_{\theta_{1}^{\epsilon}}|v^{\epsilon}|^{q}d\bar{X}\leq C\|v^{\epsilon}\|^{q}_{L^{q}(I;W^{s,p}(0,\epsilon K_{\epsilon}(\bar{x})))},\,\,\,\forall q\geq 1.

In particular, if p2p\geq 2, we have

(20) 1ϵγθ1ϵ|vϵ|p𝑑X¯CvϵW1,p(Raϵ)q,qp.\frac{1}{\epsilon^{\gamma}}\int_{\theta_{1}^{\epsilon}}|v^{\epsilon}|^{p}d\bar{X}\leq C\|v^{\epsilon}\|^{q}_{W^{1,p}(R^{\epsilon}_{a})},\,\,\forall q\leq p.
Proof.

The estimate (19) is true by changing the HsH^{s} space by Ws,pW^{s,p}, for 11p<s11-\frac{1}{p}<s\leq 1, in [8, Theorem 3.7], which implies that exists C>0C>0 independent of ϵ>0\epsilon>0 such that

1ϵγθ1ϵ|vϵ|q𝑑X¯CvϵLq(I;Ws,p(0,ϵKϵ(x¯)))q,q1.\frac{1}{\epsilon^{\gamma}}\int_{\theta_{1}^{\epsilon}}|v^{\epsilon}|^{q}d\bar{X}\leq C\|v^{\epsilon}\|^{q}_{L^{q}(I;W^{s,p}(0,\epsilon K_{\epsilon}(\bar{x})))},\,\,\,\forall q\geq 1.

Also it is not difficult to see that W1,p(Raϵ)W^{1,p}(R^{\epsilon}_{a}) is included in Lp(I;W1,p(0,ϵKϵ(x¯)))L^{p}(I;W^{1,p}(0,\epsilon K_{\epsilon}(\bar{x}))) and, consequently, in Lp(I;Ws,p(0,ϵKϵ(x¯)))L^{p}(I;W^{s,p}(0,\epsilon K_{\epsilon}(\bar{x}))), with constant independent of ϵ>0\epsilon>0, analogously as [8, Preposition 3.6].

On the other hand, if p>2p>2, we have that pq>1\frac{p}{q}>1 for all q<pq<p, and then, if we define K¯ϵ(x¯)=ϵKϵ(x¯){\bar{K}_{\epsilon}}(\bar{x})=\epsilon K_{\epsilon}(\bar{x}),

vϵLq(I;W1,p(0,K¯ϵ(x¯))q\displaystyle\|v^{\epsilon}\|^{q}_{L^{q}(I;W^{1,p}(0,\bar{K}_{\epsilon}(\bar{x}))} =I(0K¯ϵ(x¯)|vϵ(x¯,y¯)|p𝑑y¯)qp𝑑x¯\displaystyle=\int_{I}\Big{(}\int_{0}^{\bar{K}_{\epsilon}(\bar{x})}|v^{\epsilon}(\bar{x},\bar{y})|^{p}d\bar{y}\Big{)}^{\frac{q}{p}}d\bar{x}
+I(0K¯ϵ(x¯)|y¯vϵ(x¯,y¯)|p𝑑y¯)qp𝑑x¯\displaystyle+\int_{I}\Big{(}\int_{0}^{\bar{K}_{\epsilon}(\bar{x})}|\partial_{\bar{y}}v^{\epsilon}(\bar{x},\bar{y})|^{p}d\bar{y}\Big{)}^{\frac{q}{p}}d\bar{x}
K1pqp(I0K¯ϵ(x¯)|vϵ(x¯,y¯)|p𝑑y¯𝑑x¯)qp\displaystyle\leq K_{1}^{\frac{p-q}{p}}\Big{(}\int_{I}\int_{0}^{\bar{K}_{\epsilon}(\bar{x})}|v^{\epsilon}(\bar{x},\bar{y})|^{p}d\bar{y}d\bar{x}\Big{)}^{\frac{q}{p}}
+K1pqp(I0K¯ϵ(x¯)|y¯vϵ(x¯,y¯)|p𝑑y¯𝑑x¯)qp\displaystyle+K_{1}^{\frac{p-q}{p}}\Big{(}\int_{I}\int_{0}^{\bar{K}_{\epsilon}(\bar{x})}|\partial_{\bar{y}}v^{\epsilon}(\bar{x},\bar{y})|^{p}d\bar{y}d\bar{x}\Big{)}^{\frac{q}{p}}
CvϵW1,p(Raϵ)q\displaystyle\leq C\|v^{\epsilon}\|^{q}_{W^{1,p}(R^{\epsilon}_{a})}

We would like to obtain a uniform boundedness for the solutions of the problem (15) for any ϵ>0\epsilon>0 and γ>0\gamma>0, so from now on we will assume that fϵf^{\epsilon} satisfies the following hypothesis

(21) 1ϵγ|fϵ|L2(θϵ)2c,\frac{1}{\epsilon^{\gamma}}|||f^{\epsilon}|||^{2}_{L^{2}(\theta^{\epsilon})}\leq c,

for some positive constant cc independent of ϵ>0\epsilon>0.

Lemma 3.3.

Consider the variational formulation of (15). Then, there exists c0>0c_{0}>0, also independent of ϵ\epsilon, such that

|vϵ|H1(Raϵ)c0.|||v^{\epsilon}|||_{H^{1}(R^{\epsilon}_{a})}\leq c_{0}.
Proof.

Taking the test function vϵv^{\epsilon} in the variational formulation of (15) we have

|vϵ|H1(Raϵ)2\displaystyle|||v^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon}_{a})} =1ϵvϵH1(Raϵ)2(1ϵγ+1θ1ϵ|f1ϵ(x¯,y¯)|2𝑑X¯)12(1ϵγ+1θ1ϵ|vϵ(x¯,y¯)|2𝑑X¯)12\displaystyle=\frac{1}{\epsilon}\|v^{\epsilon}\|^{2}_{H^{1}(R^{\epsilon}_{a})}\leq\Big{(}\frac{1}{\epsilon^{\gamma+1}}\int_{\theta^{\epsilon}_{1}}|f^{\epsilon}_{1}(\bar{x},\bar{y})|^{2}d\bar{X}\Big{)}^{\frac{1}{2}}\Big{(}\frac{1}{\epsilon^{\gamma+1}}\int_{\theta^{\epsilon}_{1}}|v^{\epsilon}(\bar{x},\bar{y})|^{2}d\bar{X}\Big{)}^{\frac{1}{2}}
C(1ϵγ|f1ϵ|L2(θ1ϵ)2)12|vϵ|H1(Raϵ)\displaystyle\leq C\Big{(}\frac{1}{\epsilon^{\gamma}}|||f_{1}^{\epsilon}|||_{L^{2}(\theta^{\epsilon}_{1})}^{2}\Big{)}^{\frac{1}{2}}|||v^{\epsilon}|||_{H^{1}(R^{\epsilon}_{a})}

using (18), (21) and Lemma 3.2 we obtain

|vϵ|H1(Raϵ)c0,|||v^{\epsilon}|||_{H^{1}(R^{\epsilon}_{a})}\leq c_{0},

where c0c_{0} is a positive constant independent of ϵ>0\epsilon>0. ∎

The next result discusses the relationship between the solutions of problems (2) and (15).

Lemma 3.4.

Let wϵw^{\epsilon} and vϵv^{\epsilon} be the solutions of problems (2) and (15) respectively, assuming (21), we have

|wϵLϵvϵ|H1(Raϵ)ϵ00.|||w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon}|||_{H^{1}(R_{a}^{\epsilon})}\buildrel\epsilon\to 0\over{\longrightarrow}0.
Proof.

From the definition of LϵL^{\epsilon} we have

(wϵLϵ)x¯=wϵxϵ(ddxkϵ1(x))wϵy,\displaystyle\frac{\partial(w^{\epsilon}\circ L^{\epsilon})}{\partial\bar{x}}=\frac{\partial w^{\epsilon}}{\partial x}-\epsilon\Big{(}\frac{d}{dx}k^{1}_{\epsilon}(x)\Big{)}\frac{\partial w^{\epsilon}}{\partial y},
(wϵLϵ)y¯=wϵy.\displaystyle\frac{\partial(w^{\epsilon}\circ L^{\epsilon})}{\partial\bar{y}}=\frac{\partial w^{\epsilon}}{\partial y}.

In the new system of variables (x=x¯x=\bar{x} and y¯=y+ϵkϵ1(x)\bar{y}=y\,+\,\epsilon\,k^{1}_{\epsilon}(x)) the variational formulation of (2) is given by

(22) Raϵ{(wϵLϵ)xφx+(wϵLϵ)y¯φy¯+(wϵLϵ)φ}𝑑X¯+Raϵ{ϵ(ddxkϵ1(x))((wϵLϵ)y¯φx+(wϵLϵ)xφy¯)}𝑑X¯+Raϵ{(ϵddxkϵ1(x))2(wϵLϵ)y¯φy¯}𝑑X¯=1ϵγθ1ϵf1ϵφ𝑑X¯,φH1(Raϵ).\begin{split}&\int_{R_{a}^{\epsilon}}\Big{\{}\frac{\partial(w^{\epsilon}\circ L^{\epsilon})}{\partial x}\frac{\partial\varphi}{\partial x}+\frac{\partial(w^{\epsilon}\circ L^{\epsilon})}{\partial\bar{y}}\frac{\partial\varphi}{\partial\bar{y}}+(w^{\epsilon}\circ L^{\epsilon})\varphi\Big{\}}d\bar{X}\\ &\quad+\int_{R_{a}^{\epsilon}}\Big{\{}\epsilon\Big{(}\frac{d}{dx}k^{1}_{\epsilon}(x)\Big{)}\Big{(}\frac{\partial(w^{\epsilon}\circ L^{\epsilon})}{\partial\bar{y}}\frac{\partial\varphi}{\partial x}+\frac{\partial(w^{\epsilon}\circ L^{\epsilon})}{\partial x}\frac{\partial\varphi}{\partial\bar{y}}\Big{)}\Big{\}}d\bar{X}\\ &\quad+\int_{R_{a}^{\epsilon}}\Big{\{}\Big{(}\epsilon\frac{d}{dx}k^{1}_{\epsilon}(x)\Big{)}^{2}\frac{\partial(w^{\epsilon}\circ L^{\epsilon})}{\partial\bar{y}}\frac{\partial\varphi}{\partial\bar{y}}\Big{\}}d\bar{X}\\ &\quad=\frac{1}{\epsilon^{\gamma}}\int_{\theta^{\epsilon}_{1}}f^{\epsilon}_{1}\varphi\,d\bar{X},\quad\forall\varphi\in H^{1}(R_{a}^{\epsilon}).\end{split}

On the other hand, the weak formulation of (15) is: find vϵH1(Raϵ)v^{\epsilon}\in H^{1}(R_{a}^{\epsilon}) such that

(23) Raϵ{vϵxφx+vϵy¯φy¯+vϵφ}𝑑X¯=1ϵγθ1ϵf1ϵφ𝑑X¯,φH1(Raϵ).\int_{R_{a}^{\epsilon}}\Big{\{}\frac{\partial v^{\epsilon}}{\partial x}\frac{\partial\varphi}{\partial x}+\frac{\partial v^{\epsilon}}{\partial\bar{y}}\frac{\partial\varphi}{\partial\bar{y}}+v^{\epsilon}\varphi\Big{\}}\,d\bar{X}=\frac{1}{\epsilon^{\gamma}}\int_{\theta^{\epsilon}_{1}}f^{\epsilon}_{1}\varphi\,d\bar{X},\quad\forall\varphi\in H^{1}(R_{a}^{\epsilon}).

Therefore, subtracting (23) from (22), taking (wϵLϵvϵ)(w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon}) as a test function and after some computations and simplifications, we obtain

wϵLϵvϵH1(Raϵ)2cη1(ϵ)(wϵLϵ)L2(Raϵ)(wϵLϵvϵ)L2(Raϵ),\displaystyle\|w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon}\|_{H^{1}(R_{a}^{\epsilon})}^{2}\leq c\,\eta^{1}(\epsilon)\|\nabla(w^{\epsilon}\circ L^{\epsilon})\|_{L^{2}(R_{a}^{\epsilon})}\|\nabla(w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon})\|_{L^{2}(R_{a}^{\epsilon})},

where c>0c>0 is a constant independent of ϵ\epsilon and η1(ϵ)\eta^{1}(\epsilon) is given by (9).

This implies

wϵLϵvϵH1(Raϵ)\displaystyle\|w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon}\|_{H^{1}(R_{a}^{\epsilon})} cη1(ϵ)wϵLϵH1(Raϵ)\displaystyle\leq c\,\eta^{1}(\epsilon)\|w^{\epsilon}\circ L^{\epsilon}\|_{H^{1}(R_{a}^{\epsilon})}
cη1(ϵ)wϵLϵvϵH1(Raϵ)+cη1(ϵ)vϵH1(Raϵ),\displaystyle\leq c\,\eta^{1}(\epsilon)\|w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon}\|_{H^{1}(R_{a}^{\epsilon})}+c\,\eta^{1}(\epsilon)\|v^{\epsilon}\|_{H^{1}(R_{a}^{\epsilon})},

and therefore using Lemma 3.3 since vϵv_{\epsilon} is the solution of (15) we have

|wϵLϵvϵ|H1(Raϵ)cη1(ϵ)1cη1(ϵ)|vϵ|H1(Raϵ)cη1(ϵ)1cη1(ϵ)c0,\displaystyle|||w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon}|||_{H^{1}(R_{a}^{\epsilon})}\leq\frac{c\,\eta^{1}(\epsilon)}{1-c\,\eta^{1}(\epsilon)}|||v^{\epsilon}|||_{H^{1}(R_{a}^{\epsilon})}\leq\frac{c\,\eta^{1}(\epsilon)}{1-c\,\eta^{1}(\epsilon)}c_{0},

the result follows from the definition of η1(ϵ)\eta^{1}(\epsilon). ∎

Remark 2.

Notice that from the point of view of the limit behavior of the solutions it is the same to study problem (2) defined in the doubly oscillating thin domain as to analyze problem (15) posed in a thin domain with just one oscillating boundary. It is important to note that this is true because of (H.1). If that assumption is not satisfied, at least for kϵ1k_{\epsilon}^{1}, then the simplification it is not possible. For instance, if we have a domain with oscillations with the same period in both boundaries like this one

Rϵ={(x,y)2|x(0,1),ϵ(2g(x/ϵ))<y<ϵg(x/ϵ)},R^{\epsilon}=\Big{\{}(x,y)\in\mathbb{R}^{2}\;|\;x\in(0,1),\;-\epsilon(2-g(x/\epsilon))<y<\epsilon g(x/\epsilon)\Big{\}},

where g:g:\mathbb{R}\to\mathbb{R} is a smooth LL-periodic function. Observe that in this particular case we have

kϵ1(x)=2g(x/ϵ),kϵ2(x)=g(x/ϵ).k^{1}_{\epsilon}(x)=2-g(x/\epsilon),\quad k^{2}_{\epsilon}(x)=g(x/\epsilon).

Therefore, it follows straightforward that condition (4) is not satisfied. Observe that the original problem (2) for this particular thin domain is in the framework of the classical periodic homogenization while the converted problem (15) is posed in a rectangle of height ϵ\epsilon where homogenization theory is not necessary to analyze the behavior of the solutions. Indeed, if kϵ1k_{\epsilon}^{1} does not satisfy (4) the solutions of problems (2) and (15) are not comparable in general.

Now we define a transformation on the thin domain RaϵR^{\epsilon}_{a}, which will map RaϵR^{\epsilon}_{a} into the fixed rectangle Q=I×(0,1)Q=I\times(0,1). This transformation is given by

Sϵ:QRaϵ(x,y)(x¯,y¯):=(x,yϵKϵ(x)).\begin{array}[]{rl}S^{\epsilon}:Q&\longrightarrow R_{a}^{\epsilon}\\ (x,y)&\longrightarrow(\bar{x},\bar{y}):=(x\,,\,y\,\epsilon K_{\epsilon}(x)).\end{array}

We recall that Kϵ(x)=kϵ2(x)+kϵ1(x)K_{\epsilon}(x)=k_{\epsilon}^{2}(x)+k_{\epsilon}^{1}(x).

Using the chain rule and standard computations it is not difficult to see that there exist c,C>0c,C>0 such that if vϵH1(Raϵ)v^{\epsilon}\in H^{1}(R_{a}^{\epsilon}) and uϵ=vϵSϵH1(Q)u^{\epsilon}=v^{\epsilon}\circ S^{\epsilon}\in H^{1}(Q) then the following estimates hold

(24) c|vϵ|L2(Raϵ)2uϵL2(Q)2C|vϵ|L2(Raϵ)2,c|||v^{\epsilon}|||^{2}_{L^{2}(R^{\epsilon}_{a})}\leq\|u^{\epsilon}\|^{2}_{L^{2}(Q)}\leq C|||v^{\epsilon}|||^{2}_{L^{2}(R^{\epsilon}_{a})},
(25) c|vϵy¯|L2(Raϵ)21ϵ2uϵyL2(Q)2C|vϵy¯|L2(Raϵ)2,c\Big{|}\Big{|}\Big{|}\frac{\partial v^{\epsilon}}{\partial\bar{y}}\Big{|}\Big{|}\Big{|}^{2}_{L^{2}(R^{\epsilon}_{a})}\leq\frac{1}{\epsilon^{2}}\Big{\|}\frac{\partial u^{\epsilon}}{\partial y}\Big{\|}^{2}_{L^{2}(Q)}\leq C\Big{|}\Big{|}\Big{|}\frac{\partial v^{\epsilon}}{\partial\bar{y}}\Big{|}\Big{|}\Big{|}^{2}_{L^{2}(R^{\epsilon}_{a})},
(26) c|vϵ|L2(Raϵ)2uϵL2(Q)2C|vϵ|L2(Raϵ)2.c|||\nabla v^{\epsilon}|||^{2}_{L^{2}(R^{\epsilon}_{a})}\leq\|\nabla u^{\epsilon}\|^{2}_{L^{2}(Q)}\leq C|||\nabla v^{\epsilon}|||^{2}_{L^{2}(R^{\epsilon}_{a})}.

Now, under this change of variables and defining uϵ=vϵSϵu^{\epsilon}=v^{\epsilon}\circ S^{\epsilon} where vϵv^{\epsilon} satisfies (15) and f2ϵ=f1ϵSϵf^{\epsilon}_{2}=f^{\epsilon}_{1}\circ S^{\epsilon}, where f1ϵf_{1}^{\epsilon} is defined in (17), problem (15) becomes

(27) {1Kϵdiv(Bϵ(uϵ))+uϵ=1ϵγχθ2ϵf2ϵ in Q,B(uϵ)η=0 on Q,uϵ=vϵSϵ in Q,\left\{\begin{gathered}-\frac{1}{K_{\epsilon}}\hbox{div}\big{(}B^{\epsilon}(u^{\epsilon})\big{)}+u^{\epsilon}=\frac{1}{\epsilon^{\gamma}}\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon}\quad\textrm{ in }Q,\\ B(u^{\epsilon})\cdot\eta=0\quad\textrm{ on }\partial Q,\\ u^{\epsilon}=v^{\epsilon}\circ S^{\epsilon}\textrm{ in }Q,\end{gathered}\right.

where the function χθ2ϵ:2\chi_{\theta^{\epsilon}_{2}}:\mathbb{R}^{2}\rightarrow\mathbb{R} is the characteristic function of the narrow strip θ2ϵ\theta^{\epsilon}_{2} given by

(28) θ2ϵ={(x,y)2|xI, 1ϵγHϵ(x)Kϵ(x)<y<1}.\theta^{\epsilon}_{2}=\Big{\{}(x,y)\in\mathbb{R}^{2}\;|\;x\in I,\;1-\epsilon^{\gamma}\frac{H_{\epsilon}(x)}{K_{\epsilon}(x)}<y<1\Big{\}}.

The vector η\eta denotes the outward unit normal vector field to Q\partial Q and

Bϵ(uϵ)=(KϵuϵxydKϵdxuϵy,ydKϵdxuϵx+(y2Kϵ(dKϵdx)2+1ϵ2Kϵ)uϵy).B^{\epsilon}(u^{\epsilon})=\Big{(}K_{\epsilon}\frac{\partial u^{\epsilon}}{\partial x}-y\displaystyle\frac{dK_{\epsilon}}{dx}\frac{\partial u^{\epsilon}}{\partial y}\,,\,-y\displaystyle\frac{dK_{\epsilon}}{dx}\frac{\partial u^{\epsilon}}{\partial x}+\Big{(}\frac{y^{2}}{K_{\epsilon}}\Big{(}\displaystyle\frac{dK_{\epsilon}}{dx}\Big{)}^{2}+\frac{1}{\epsilon^{2}K_{\epsilon}}\Big{)}\frac{\partial u^{\epsilon}}{\partial y}\Big{)}.

Notice that

(29) c|f1ϵ|L2(θ1ϵ)2f2ϵL2(θ2ϵ)2C|f1ϵ|L2(θ1ϵ)2.c|||f^{\epsilon}_{1}|||^{2}_{L^{2}(\theta^{\epsilon}_{1})}\leq\|f^{\epsilon}_{2}\|^{2}_{L^{2}(\theta_{2}^{\epsilon})}\leq C|||f^{\epsilon}_{1}|||^{2}_{L^{2}(\theta^{\epsilon}_{1})}.

Therefore from (18) and (29) we have

(30) c|fϵ|L2(θϵ)2f2ϵL2(θ2ϵ)2C|fϵ|L2(θϵ)2.c|||f^{\epsilon}|||^{2}_{L^{2}(\theta^{\epsilon})}\leq\|f^{\epsilon}_{2}\|^{2}_{L^{2}(\theta_{2}^{\epsilon})}\leq C|||f^{\epsilon}|||^{2}_{L^{2}(\theta^{\epsilon})}.

In the new system of coordinates we obtain a domain which is neither thin nor oscillating anymore. In some sense, we have rescaled the neighborhood (16) into the strip θ2ϵQ\theta_{2}^{\epsilon}\subset Q and substituted the thin domain RaϵR_{a}^{\epsilon} for a domain QQ independent on ϵ\epsilon, at a cost of replacing the oscillating thin domain by oscillating coefficients in the differential operator.

In order to analyze the limit behavior of the solutions of (27) we establish the relation to the solutions of the following easier problem

(31) {1Kϵ[x(Kϵw1ϵx)+1ϵ2Kϵ2w1ϵy2]+w1ϵ=1ϵγχθ2ϵf2ϵ in Q,Kϵw1ϵxη1+1ϵ2Kϵw1ϵyη2=0 on Q,\left\{\begin{gathered}-\frac{1}{K_{\epsilon}}\Big{[}\frac{\partial}{\partial x}\Big{(}K_{\epsilon}\frac{\partial w_{1}^{\epsilon}}{\partial x}\Big{)}+\frac{1}{\epsilon^{2}K_{\epsilon}}\frac{\partial^{2}w_{1}^{\epsilon}}{\partial y^{2}}\Big{]}+w_{1}^{\epsilon}=\frac{1}{\epsilon^{\gamma}}\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon}\quad\textrm{ in }Q,\\ K_{\epsilon}\frac{\partial w_{1}^{\epsilon}}{\partial x}\,\eta_{1}+\frac{1}{\epsilon^{2}K_{\epsilon}}\frac{\partial w_{1}^{\epsilon}}{\partial y}\,\eta_{2}=0\quad\textrm{ on }\partial Q,\end{gathered}\right.

where η=(η1,η2)\eta=(\eta_{1},\eta_{2}) is the outward unit normal to Q\partial Q.

Before obtaining a priori estimates for the solutions of (31), it is necessary to prove the following result

Lemma 3.5.

Let Q=I×(0,1)Q=I\times(0,1) and θ2ϵ\theta_{2}^{\epsilon} like defined in (28), suppose that w1ϵHs(Q)w_{1}^{\epsilon}\in H^{s}(Q) with 12<s 1\frac{1}{2}<s\leq\ 1 and s11qs-1\geq-\frac{1}{q}. Then, for small ϵ0>0\epsilon_{0}>0, there exist a constant C>0C>0 independent of ϵ\epsilon and w1ϵw_{1}^{\epsilon}, such that for any 0<ϵϵ00<\epsilon\leq\epsilon_{0}, we have

1ϵγθ2ϵ|w1ϵ|q𝑑x𝑑yCw1ϵHs(Q)q.\frac{1}{\epsilon^{\gamma}}\int_{\theta_{2}^{\epsilon}}|w_{1}^{\epsilon}|^{q}dxdy\leq C\|w_{1}^{\epsilon}\|^{q}_{H^{s}(Q)}.
Proof.

First we observe that

1ϵγθ2ϵ|w1ϵ(x,y)|q𝑑x𝑑y=1ϵγI0ϵγHϵ(x)Kϵ(x)|w1ϵ(x,1y)|q𝑑y𝑑xrϵ|w1ϵ(x,1y)|q𝑑x𝑑y,\frac{1}{\epsilon^{\gamma}}\int_{\theta_{2}^{\epsilon}}|w_{1}^{\epsilon}(x,y)|^{q}dxdy=\frac{1}{\epsilon^{\gamma}}\int_{I}\int_{0}^{\epsilon^{\gamma}\frac{H_{\epsilon}(x)}{K_{\epsilon}(x)}}|w_{1}^{\epsilon}(x,1-y)|^{q}dydx\leq\int_{r_{\epsilon}}|w_{1}^{\epsilon}(x,1-y)|^{q}dxdy,

where rϵr_{\epsilon} is the strip without oscillatory behavior given by

rϵ={(x,y)2|xI,0<y<ϵγH1K0}.r_{\epsilon}=\Big{\{}(x,y)\in\mathbb{R}^{2}\,|\,x\in I,0<y<\epsilon^{\gamma}\frac{H_{1}}{K_{0}}\Big{\}}.

Using [6, Lemma 2.1] we have that there exists ϵ0>0\epsilon_{0}>0 and C>0C>0 independent of ϵ\epsilon and vϵ=w1ϵτv^{\epsilon}=w^{\epsilon}_{1}\circ\tau such that

1ϵγθ2ϵ|w1ϵ(x,y)|q𝑑x𝑑yvϵHs(τ1(Q))q,ϵ(0,ϵ0),\frac{1}{\epsilon^{\gamma}}\int_{\theta_{2}^{\epsilon}}|w_{1}^{\epsilon}(x,y)|^{q}dxdy\leq\|v^{\epsilon}\|^{q}_{H^{s}(\tau^{-1}(Q))},\,\,\,\forall\epsilon\in(0,\epsilon_{0}),

where we are taking τ:22\tau:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} given by τ(x,y)=(x,1y)\tau(x,y)=(x,1-y). From [6, Section 2] we have that the norms vϵHs(τ1(Q))\|v^{\epsilon}\|_{H^{s}(\tau^{-1}(Q))} and w1ϵHs(Q)\|w_{1}^{\epsilon}\|_{H^{s}(Q)} are equivalents, so we can conclude the proof. ∎

Observe that assuming (21), estimates (30), under the assumptions on the functions kϵ1k^{1}_{\epsilon} and kϵ2k^{2}_{\epsilon}, Lemma 3.5, so the equation (31) admits a unique solution w1ϵH1(Q)w_{1}^{\epsilon}\in H^{1}(Q), which satisfies the priori estimates

(32) w1ϵL2(Q),w1ϵxL2(Q),1ϵw1ϵyL2(Q)C,\begin{gathered}\|w_{1}^{\epsilon}\|_{L^{2}(Q)},\,\,\,\Big{\|}\frac{\partial w_{1}^{\epsilon}}{\partial x}\Big{\|}_{L^{2}(Q)},\,\,\,\frac{1}{\epsilon}\Big{\|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{\|}_{L^{2}(Q)}\leq C,\end{gathered}

where the positive constant CC is independet of ϵ>0\epsilon>0.

Lemma 3.6.

Let uϵu^{\epsilon} and w1ϵw_{1}^{\epsilon} be the solution of problems (27) and (31) respectively and suppose that fϵf^{\epsilon} satisfying (21). Then, we have

(uϵw1ϵ)xL2(Q)2+1ϵ2(uϵw1ϵ)yL2(Q)2+uϵw1ϵL2(Q)2ϵ00.\displaystyle\Big{|}\Big{|}\frac{\partial(u^{\epsilon}-w_{1}^{\epsilon})}{\partial x}\Big{|}\Big{|}_{L^{2}(Q)}^{2}+\frac{1}{\epsilon^{2}}\Big{|}\Big{|}\frac{\partial(u^{\epsilon}-w_{1}^{\epsilon})}{\partial y}\Big{|}\Big{|}_{L^{2}(Q)}^{2}+||u^{\epsilon}-w_{1}^{\epsilon}||_{L^{2}(Q)}^{2}\buildrel\epsilon\to 0\over{\longrightarrow}0.
Proof.

Subtracting the weak formulation of (31) from the weak formulation of (27) and choosing uϵw1ϵu^{\epsilon}-w_{1}^{\epsilon} as test function we get

Q{Kϵ((uϵw1ϵ)x)2+1ϵ2Kϵ((uϵw1ϵ)y)2+Kϵ(uϵw1ϵ)2}dxdy=Q{yuϵydKϵdx(uϵw1ϵ)x+ydKϵdxuϵx(uϵw1ϵ)yy2Kϵ(dKϵdx)2uϵy(uϵw1ϵ)y}𝑑x𝑑yCη(ϵ)Q{1ϵ|uϵy|(|uϵx|+|w1ϵx|)+|uϵx|1ϵ(|uϵy|+|w1ϵy|)}𝑑x𝑑y+C0(η(ϵ))2Q(1ϵ2|uϵy|2+1ϵ|uϵy|1ϵ|w1ϵy|)𝑑x𝑑y,\begin{split}\int_{Q}&\Big{\{}K_{\epsilon}\Big{(}\frac{\partial(u^{\epsilon}-w_{1}^{\epsilon})}{\partial x}\Big{)}^{2}+\frac{1}{\epsilon^{2}K_{\epsilon}}\Big{(}\frac{\partial(u^{\epsilon}-w_{1}^{\epsilon})}{\partial y}\Big{)}^{2}+K_{\epsilon}\big{(}u^{\epsilon}-w_{1}^{\epsilon}\big{)}^{2}\Big{\}}dxdy\\ =&\int_{Q}\Big{\{}y\frac{\partial u^{\epsilon}}{\partial y}\displaystyle\frac{dK_{\epsilon}}{dx}\frac{\partial(u^{\epsilon}-w^{\epsilon}_{1})}{\partial x}+y\displaystyle\frac{dK_{\epsilon}}{dx}\frac{\partial u^{\epsilon}}{\partial x}\frac{\partial(u^{\epsilon}-w_{1}^{\epsilon})}{\partial y}-\frac{y^{2}}{K_{\epsilon}}\Big{(}\displaystyle\frac{dK_{\epsilon}}{dx}\Big{)}^{2}\frac{\partial u^{\epsilon}}{\partial y}\frac{\partial(u^{\epsilon}-w_{1}^{\epsilon})}{\partial y}\Big{\}}dxdy\\ \leq&\,C\eta(\epsilon)\int_{Q}\Big{\{}\frac{1}{\epsilon}\Big{|}\frac{\partial u^{\epsilon}}{\partial y}\Big{|}\Big{(}\Big{|}\frac{\partial u^{\epsilon}}{\partial x}\Big{|}+\Big{|}\frac{\partial w^{\epsilon}_{1}}{\partial x}\Big{|}\Big{)}+\Big{|}\frac{\partial u^{\epsilon}}{\partial x}\Big{|}\frac{1}{\epsilon}\Big{(}\Big{|}\frac{\partial u^{\epsilon}}{\partial y}\Big{|}+\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{|}\Big{)}\Big{\}}dxdy\\ +&\,C_{0}(\eta(\epsilon))^{2}\int_{Q}\Big{(}\frac{1}{\epsilon^{2}}\Big{|}\frac{\partial u^{\epsilon}}{\partial y}\Big{|}^{2}+\frac{1}{\epsilon}\Big{|}\frac{\partial u^{\epsilon}}{\partial y}\Big{|}\frac{1}{\epsilon}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{|}\Big{)}dxdy,\end{split}

where CC and C0C_{0} are positive constants which does not depend on ϵ\epsilon. Taking into account that ϵ|dKϵdx|η(ϵ)\epsilon\Big{|}\displaystyle\frac{dK_{\epsilon}}{dx}\Big{|}\leq\eta(\epsilon), convergence (4), estimates (25) and(26) applied to uϵu^{\epsilon} and vϵv^{\epsilon}, Lemma 3.3, the priori estimate of w1ϵw_{1}^{\epsilon} (see (32)) and following standard computations we obtain the result. ∎

Finally, we compare the behavior of the solution of (31) to the solutions of the following problem posed in II\subset\mathbb{R}

(33) {1Kϵx(Kϵu1ϵx)+u1ϵ=1ϵγf3ϵ in I,(u1ϵ)x(a)=(u1ϵ)x(b)=0 on I,\left\{\begin{gathered}-\frac{1}{K_{\epsilon}}\frac{\partial}{\partial x}\Big{(}K_{\epsilon}\frac{\partial u_{1}^{\epsilon}}{\partial x}\Big{)}+u_{1}^{\epsilon}=\frac{1}{\epsilon^{\gamma}}f^{\epsilon}_{3}\quad\textrm{ in }I,\\ (u_{1}^{\epsilon})_{x}(a)=(u_{1}^{\epsilon})_{x}(b)=0\quad\textrm{ on }\partial I,\end{gathered}\right.

where f3ϵ(x)=1ϵγHϵ(x)Kϵ(x)1f2ϵ(x,y)𝑑yf^{\epsilon}_{3}(x)=\displaystyle\int_{1-\epsilon^{\gamma}\frac{H_{\epsilon}(x)}{K_{\epsilon}(x)}}^{1}f_{2}^{\epsilon}(x,y)dy for a.e. xIx\in I, which is a function depending only on the xx variable.

Then, considering u1ϵ(x)u_{1}^{\epsilon}(x) as a function defined in QQ (extending it in a constant way in the yy direction) we prove the following lemma.

Lemma 3.7.

Let u1ϵu^{\epsilon}_{1} and w1ϵw_{1}^{\epsilon} be the solution of problems (33) and (31) respectively and suppose that fϵf^{\epsilon} satisfying (21). Then, we have

(w1ϵu1ϵ)xL2(Q)2+1ϵ2w1ϵyL2(Q)2+w1ϵu1ϵL2(Q)2ϵ00.\Big{|}\Big{|}\frac{\partial(w^{\epsilon}_{1}-u_{1}^{\epsilon})}{\partial x}\Big{|}\Big{|}_{L^{2}(Q)}^{2}+\frac{1}{\epsilon^{2}}\Big{|}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{|}\Big{|}_{L^{2}(Q)}^{2}+||w^{\epsilon}_{1}-u_{1}^{\epsilon}||^{2}_{L^{2}(Q)}\buildrel\epsilon\to 0\over{\longrightarrow}0.
Proof.

Taking w1ϵu1ϵw_{1}^{\epsilon}-u_{1}^{\epsilon} as test function in the variational formulation of (31) and 01w1ϵ𝑑yu1ϵ\int_{0}^{1}w_{1}^{\epsilon}dy-u_{1}^{\epsilon} as test function in the variational formulation of (33) and subtracting both weak formulations we obtain

(34) Q{Kϵ((w1ϵu1ϵ)x)2+1ϵ2Kϵ(w1ϵy)2+Kϵ(w1ϵu1ϵ)2}dxdy=1ϵγQKϵχθ2ϵf2ϵ(w1ϵu1ϵ)𝑑x𝑑y1ϵγQKϵf3ϵ(w1ϵu1ϵ)𝑑x𝑑y.\begin{split}\int_{Q}&\Big{\{}K_{\epsilon}\Big{(}\frac{\partial(w^{\epsilon}_{1}-u_{1}^{\epsilon})}{\partial x}\Big{)}^{2}+\frac{1}{\epsilon^{2}K_{\epsilon}}\Big{(}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{)}^{2}+K_{\epsilon}\big{(}w_{1}^{\epsilon}-u_{1}^{\epsilon}\big{)}^{2}\Big{\}}dxdy\\ &=\frac{1}{\epsilon^{\gamma}}\int_{Q}K_{\epsilon}\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon}(w_{1}^{\epsilon}-u_{1}^{\epsilon})dxdy-\frac{1}{\epsilon^{\gamma}}\int_{Q}K_{\epsilon}f_{3}^{\epsilon}(w_{1}^{\epsilon}-u_{1}^{\epsilon})dxdy.\end{split}

Now we analyze the two terms in the right hand-side. First, taking into account the definition of f3ϵf_{3}^{\epsilon} we have that for any function φ\varphi defined in II, that is, φ=φ(x)\varphi=\varphi(x)

QKϵ(f3ϵχθ2ϵf2ϵ)φ𝑑x𝑑y=IφKϵf3ϵ𝑑xθ2ϵφKϵf2ϵ𝑑y=0.\int_{Q}K_{\epsilon}(f_{3}^{\epsilon}-\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon})\varphi dxdy=\int_{I}\varphi K_{\epsilon}f_{3}^{\epsilon}dx-\int_{\theta^{\epsilon}_{2}}\varphi K_{\epsilon}f_{2}^{\epsilon}dy=0.

In particular QKϵ(f3ϵχθ2ϵf2ϵ)u1ϵ𝑑x𝑑y=0\displaystyle\int_{Q}K_{\epsilon}(f_{3}^{\epsilon}-\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon})u_{1}^{\epsilon}dxdy=0 and QKϵ(f3ϵχθ2ϵf2ϵ)w1ϵ(x,0)𝑑x𝑑y=0\displaystyle\int_{Q}K_{\epsilon}(f_{3}^{\epsilon}-\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon})w_{1}^{\epsilon}(x,0)dxdy=0. Hence, using Holder inequality, (6), (10), (21), (30) and (32) we get

1ϵγ|QKϵ(χθ2ϵf2ϵf3ϵ)w1ϵ𝑑x𝑑y|=1ϵγ|QKϵ(χθ2ϵf2ϵf3ϵ)(w1ϵ(x,y)w1ϵ(x,0))𝑑x𝑑y|\displaystyle\frac{1}{\epsilon^{\gamma}}\Big{|}\int_{Q}K_{\epsilon}(\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon}-f_{3}^{\epsilon})w_{1}^{\epsilon}dxdy\Big{|}=\frac{1}{\epsilon^{\gamma}}\Big{|}\int_{Q}K_{\epsilon}(\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon}-f_{3}^{\epsilon})(w_{1}^{\epsilon}(x,y)-w_{1}^{\epsilon}(x,0))dxdy\Big{|}
K1ϵγQ(|χθ2ϵf2ϵf3ϵ||0yw1ϵs(x,s)𝑑s|)𝑑x𝑑y\displaystyle\leq\frac{K_{1}}{\epsilon^{\gamma}}\int_{Q}\Big{(}\Big{|}\chi_{\theta^{\epsilon}_{2}}f_{2}^{\epsilon}-f_{3}^{\epsilon}\Big{|}\Big{|}\int_{0}^{y}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)ds\Big{|}\Big{)}dxdy
K1ϵγ[θ2ϵ|f2ϵ||(01|w1ϵs(x,s)|ds)dxdy+I|f3ϵ|(01|w1ϵs(x,s)|ds)dx]\displaystyle\leq\frac{K_{1}}{\epsilon^{\gamma}}\Big{[}\int_{\theta_{2}^{\epsilon}}|f_{2}^{\epsilon}|\Big{|}\Big{(}\int_{0}^{1}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)\Big{|}ds\Big{)}dxdy+\int_{I}|f_{3}^{\epsilon}|\Big{(}\int_{0}^{1}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)\Big{|}ds\Big{)}dx\Big{]}
K1ϵγ[f2ϵL2(θ2ϵ)01|w1ϵs(x,s)|dsL2(θ2ϵ)+I|1ϵγHϵ(x)Kϵ(x)1f2ϵ(x,y)𝑑y|(01|w1ϵs(x,s)|𝑑s)𝑑x]\displaystyle\leq\frac{K_{1}}{\epsilon^{\gamma}}\Big{[}\|f_{2}^{\epsilon}\|_{L^{2}(\theta_{2}^{\epsilon})}\Big{\|}\int_{0}^{1}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)\Big{|}ds\Big{\|}_{L^{2}(\theta_{2}^{\epsilon})}+\int_{I}\Big{|}\displaystyle\int_{1-\epsilon^{\gamma}\frac{H_{\epsilon}(x)}{K_{\epsilon}(x)}}^{1}f_{2}^{\epsilon}(x,y)dy\Big{|}\Big{(}\int_{0}^{1}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)\Big{|}ds\Big{)}dx\Big{]}
K1ϵγ[f2ϵL2(θ2ϵ)(ϵγH1K0)12w1ϵyL2(Q)+I1ϵγHϵ(x)Kϵ(x)1|f2ϵ|(01|w1ϵs(x,s)|𝑑s)𝑑y𝑑x]\displaystyle\leq\frac{K_{1}}{\epsilon^{\gamma}}\Big{[}\|f_{2}^{\epsilon}\|_{L^{2}(\theta_{2}^{\epsilon})}\Big{(}\epsilon^{\gamma}\frac{H_{1}}{K_{0}}\Big{)}^{\frac{1}{2}}\Big{\|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{\|}_{L^{2}(Q)}+\int_{I}\displaystyle\int_{1-\epsilon^{\gamma}\frac{H_{\epsilon}(x)}{K_{\epsilon}(x)}}^{1}|f_{2}^{\epsilon}|\Big{(}\int_{0}^{1}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)\Big{|}ds\Big{)}dydx\Big{]}
K1ϵγ[f2ϵL2(θ2ϵ)(ϵγH1K0)12w1ϵyL2(Q)+θ2ϵ|f2ϵ|(01|w1ϵs(x,s)|𝑑s)𝑑x𝑑y]\displaystyle\leq\frac{K_{1}}{\epsilon^{\gamma}}\Big{[}\|f_{2}^{\epsilon}\|_{L^{2}(\theta_{2}^{\epsilon})}\Big{(}\epsilon^{\gamma}\frac{H_{1}}{K_{0}}\Big{)}^{\frac{1}{2}}\Big{\|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{\|}_{L^{2}(Q)}+\displaystyle\int_{\theta_{2}^{\epsilon}}|f_{2}^{\epsilon}|\Big{(}\int_{0}^{1}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)\Big{|}ds\Big{)}dxdy\Big{]}
K1ϵγ[f2ϵL2(θ2ϵ)(ϵγH1K0)12w1ϵyL2(Q)+f2ϵL2(θ2ϵ)01|w1ϵs(x,s)|dsL2(θ2ϵ)]\displaystyle\leq\frac{K_{1}}{\epsilon^{\gamma}}\Big{[}\|f_{2}^{\epsilon}\|_{L^{2}(\theta_{2}^{\epsilon})}\Big{(}\epsilon^{\gamma}\frac{H_{1}}{K_{0}}\Big{)}^{\frac{1}{2}}\Big{\|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{\|}_{L^{2}(Q)}+\|f_{2}^{\epsilon}\|_{L^{2}(\theta_{2}^{\epsilon})}\Big{\|}\int_{0}^{1}\Big{|}\frac{\partial w_{1}^{\epsilon}}{\partial s}(x,s)\Big{|}ds\Big{\|}_{L^{2}(\theta_{2}^{\epsilon})}\Big{]}
Cϵγ2f2ϵL2(θ2ϵ)w1ϵyL2(Q)C0ϵ\displaystyle\leq C\epsilon^{-\frac{\gamma}{2}}\|f_{2}^{\epsilon}\|_{L^{2}(\theta_{2}^{\epsilon})}\Big{\|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{\|}_{L^{2}(Q)}\leq C_{0}\epsilon

whrere C0C_{0} is a positive constant wich does not depend on ϵ\epsilon. Therefore, from (34) the lemma is proved. ∎

Remark 3.

Notice that u1ϵu^{\epsilon}_{1}, the solution of (33) coincides with w^\hat{w}, the solution of (11). Indeed, by the definition of f3ϵf^{\epsilon}_{3}, f2ϵf^{\epsilon}_{2} and f1ϵf^{\epsilon}_{1}, we get

f3ϵ(x)\displaystyle f^{\epsilon}_{3}(x) =1ϵγHϵ(x)Kϵ(x)1f1ϵSϵ(x,y)𝑑y=1ϵKϵ(x)ϵ[Kϵ(x)ϵγHϵ(x)]ϵKϵ(x)f1ϵ(x,y)𝑑y\displaystyle=\int_{1-\epsilon^{\gamma}\frac{H_{\epsilon}(x)}{K_{\epsilon}(x)}}^{1}f_{1}^{\epsilon}\circ S^{\epsilon}(x,y)dy=\frac{1}{\epsilon K_{\epsilon}(x)}\int_{\epsilon[K_{\epsilon}(x)-\epsilon^{\gamma}H_{\epsilon}(x)]}^{\epsilon K_{\epsilon}(x)}f^{\epsilon}_{1}(x,y)dy
=1ϵKϵ(x)ϵ[Kϵ(x)ϵγHϵ(x)]ϵKϵ(x)fϵLϵ(x,y)𝑑y=1ϵKϵ(x)ϵ[kϵ2(x)ϵγHϵ(x)]ϵkϵ2(x)fϵ(x,y)𝑑y\displaystyle=\frac{1}{\epsilon K_{\epsilon}(x)}\int_{\epsilon[K_{\epsilon}(x)-\epsilon^{\gamma}H_{\epsilon}(x)]}^{\epsilon K_{\epsilon}(x)}f^{\epsilon}\circ L^{\epsilon}(x,y)dy=\frac{1}{\epsilon K_{\epsilon}(x)}\int_{\epsilon[k^{2}_{\epsilon}(x)-\epsilon^{\gamma}H_{\epsilon}(x)]}^{\epsilon k^{2}_{\epsilon}(x)}f^{\epsilon}(x,y)dy
=f^ϵ(x)\displaystyle=\hat{f}^{\epsilon}(x)

After these lemmas we can provide a proof of the main result of this section.


Proof of Therorem 3.1. Let wϵw^{\epsilon} and w^ϵ\hat{w}^{\epsilon} the solution of (2) and (11) respectively. Then, from (14) and the fact that u1ϵu_{1}^{\epsilon} does not depend on the yy variable, we get

|wϵw^ϵ|H1(Rϵ)2=|wϵu1ϵ|H1(Rϵ)2\displaystyle|||w^{\epsilon}-\hat{w}^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon})}=|||w^{\epsilon}-u_{1}^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon})}
C|wϵLϵu1ϵLϵ|H1(Raϵ)2=C|wϵLϵu1ϵ|H1(Raϵ)2,\displaystyle\leq C|||w^{\epsilon}\circ L^{\epsilon}-u_{1}^{\epsilon}\circ L^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon}_{a})}=C|||w^{\epsilon}\circ L^{\epsilon}-u_{1}^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon}_{a})},

where CC is a positive constant independent of ϵ>0\epsilon>0. But,

|wϵLϵu1ϵ|H1(Raϵ)22|wϵLϵvϵ|H1(Raϵ)2+2|vϵu1ϵ|H1(Raϵ)2,\displaystyle|||w^{\epsilon}\circ L^{\epsilon}-u_{1}^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon}_{a})}\leq 2|||w^{\epsilon}\circ L^{\epsilon}-v^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon}_{a})}+2|||v^{\epsilon}-u_{1}^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon}_{a})},

where we have used the inequality (a+b)22a2+2b2(a+b)^{2}\leq 2a^{2}+2b^{2}.

Now, using that uϵ=vϵSϵu^{\epsilon}=v^{\epsilon}\circ S^{\epsilon} and (24), (25), (26) and that u1ϵu_{1}^{\epsilon} is independent of yy and therefore u1ϵSϵ=u1ϵu_{1}^{\epsilon}\circ S^{\epsilon}=u_{1}^{\epsilon}, we get

|vϵu1ϵ|H1(Raϵ)2C(uϵu1ϵ||H1(Q)2+1ϵ2uϵyL2(Q)2)|||v^{\epsilon}-u_{1}^{\epsilon}|||^{2}_{H^{1}(R^{\epsilon}_{a})}\leq C\Big{(}\|u^{\epsilon}-u_{1}^{\epsilon}||^{2}_{H^{1}(Q)}+\frac{1}{\epsilon^{2}}\Big{\|}\frac{\partial u^{\epsilon}}{\partial y}\Big{\|}^{2}_{L^{2}(Q)}\Big{)}

and applying now the triangular inequality

C(2uϵw1ϵH1(Q)2+2ϵ2(uϵw1ϵ)yL2(Q)2+2w1ϵu1ϵH1(Q)2+2ϵ2w1ϵyL2(Q)2).\leq C\Big{(}2\|u^{\epsilon}-w_{1}^{\epsilon}\|^{2}_{H^{1}(Q)}+\frac{2}{\epsilon^{2}}\Big{\|}\frac{\partial(u^{\epsilon}-w_{1}^{\epsilon})}{\partial y}\Big{\|}^{2}_{L^{2}(Q)}+2\|w_{1}^{\epsilon}-u_{1}^{\epsilon}\|^{2}_{H^{1}(Q)}+\frac{2}{\epsilon^{2}}\Big{\|}\frac{\partial w_{1}^{\epsilon}}{\partial y}\Big{\|}^{2}_{L^{2}(Q)}\Big{)}.

Putting all this inequalities together and using Lemmas 3.4, 3.6 and 3.7 we prove the result.

4 Limit problem for the elliptic equation

In view of Theorem 3.1, the homogenized limit problem of (2) will be obtained passing to the limit in the reduced problem (11).

As briefly explained in the introduction, we will explicitly derive the homogenized limit problem of (11) for oscillating functions satisfying (H). It is important to note that this situation encompasses the classical scenario where both the top and bottom boundaries are represented by the graphs of two periodic functions.

Recall that we assume that f^ϵ\hat{f}^{\epsilon} defined in (12) satisfies the following convergence

(35) 1ϵγKϵ()f^ϵ()=1ϵγ+1ϵkϵ1(x)ϵkϵ2(x)χθϵ(x,y)fϵ(x,y)𝑑yϵ0f0()w-L2(I),\frac{1}{\epsilon^{\gamma}}K_{\epsilon}(\cdot)\hat{f}^{\epsilon}(\cdot)=\frac{1}{\epsilon^{\gamma+1}}\int_{-\epsilon k^{1}_{\epsilon}(x)}^{\epsilon k^{2}_{\epsilon}(x)}\chi_{\theta^{\epsilon}}(x,y)f^{\epsilon}(x,y)dy\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}f_{0}(\cdot)\quad\hbox{w-}L^{2}(I),

for certain f0L2(I)f_{0}\in L^{2}(I).

To begin, we establish a priori estimates of w^ϵ{\hat{w}}^{\epsilon} that are independent of the specific functions kϵ1(x)k^{1}_{\epsilon}(x) and kϵ2(x)k^{2}_{\epsilon}(x). We leverage the fact that both oscillatory functions are uniformly bounded.

Lemma 4.1.

Consider the variational formulation of (11). Then, there exists c0>0c_{0}>0, independent of ϵ\epsilon, such that

w^ϵH1(I)c0.\|\hat{w}^{\epsilon}\|_{H^{1}(I)}\leq c_{0}.
Proof.

The weak formulation of (11) is given by

(36) I{Kϵw^xϵϕx+Kϵw^ϵϕ}𝑑x=1ϵγIKϵf^ϵϕ𝑑x, for all ϕH1(I).\int_{I}\Big{\{}K_{\epsilon}\hat{w}^{\epsilon}_{x}\phi_{x}+K_{\epsilon}{\hat{w}}^{\epsilon}\phi\big{\}}\,dx=\frac{1}{\epsilon^{\gamma}}\int_{I}K_{\epsilon}\hat{f}^{\epsilon}\phi\,dx,\quad\hbox{ for all }\phi\in H^{1}(I).

Considering w^ϵ{\hat{w}}^{\epsilon} as a test function in (36) and using (5), (6) and (21), we get

w^ϵH1(I)2\displaystyle\|\hat{w}^{\epsilon}\|^{2}_{H^{1}(I)} C1ϵγ+1Iϵkϵ1(x))ϵkϵ2(x))χθϵ(x,y)fϵ(x,y)w^ϵ(x)𝑑y𝑑x\displaystyle\leq C\frac{1}{\epsilon^{\gamma+1}}\int_{I}\int_{-\epsilon k^{1}_{\epsilon}(x))}^{\epsilon\,k^{2}_{\epsilon}(x))}\chi_{\theta^{\epsilon}}(x,y)f^{\epsilon}(x,y)\hat{w}^{\epsilon}(x)dydx
C(1ϵγ+1θϵ|fϵ|2𝑑x𝑑y)12(1ϵγ+1θϵ|w^ϵ|2𝑑x𝑑y)12\displaystyle\leq C\Big{(}\frac{1}{\epsilon^{\gamma+1}}\int_{\theta^{\epsilon}}|f^{\epsilon}|^{2}dxdy\Big{)}^{\frac{1}{2}}\Big{(}\frac{1}{\epsilon^{\gamma+1}}\int_{\theta^{\epsilon}}|\hat{w}^{\epsilon}|^{2}dxdy\Big{)}^{\frac{1}{2}}
C(1ϵγ|fϵ|L2(θϵ)2)12(IHϵ|w^ϵ|2𝑑x)12\displaystyle\leq C\Big{(}\frac{1}{\epsilon^{\gamma}}|||f^{\epsilon}|||_{L^{2}(\theta^{\epsilon})}^{2}\Big{)}^{\frac{1}{2}}\Big{(}\int_{I}H_{\epsilon}|\hat{w}^{\epsilon}|^{2}dx\Big{)}^{\frac{1}{2}}
C1w^ϵL2(I),\displaystyle\leq C_{1}\|\hat{w}^{\epsilon}\|_{L^{2}(I)},

where CC and C1C_{1} are positive constants independents of ϵ>0\epsilon>0, so we obtain

w^ϵH1(I)c0,\|\hat{w}^{\epsilon}\|_{H^{1}(I)}\leq c_{0},

where c0c_{0} is a positive constant independent of ϵ>0\epsilon>0. ∎

Thus, by weak compactness there exists w^H1(I)\hat{w}\in H^{1}(I) such that, up to subsequences

(37) w^ϵϵ0w^ wH1(I).{\hat{w}}^{\epsilon}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}\hat{w}\quad\hbox{ w}-H^{1}(I).

Now, we are poised to derive the homogenized limit problem using the classical approach for homogenization of variable coefficients in one dimensional problems. Observe that Kϵw^xϵK_{\epsilon}\hat{w}^{\epsilon}_{x} is uniformly bounded in L2(I)L^{2}(I) since

w^xϵL2(I)C and 0<Kϵ(x)<C21+C22, for each xI.||\hat{w}^{\epsilon}_{x}||_{L^{2}(I)}\leq C\quad\hbox{ and }\quad 0<K_{\epsilon}(x)<C^{1}_{2}+C^{2}_{2},\hbox{ for each }x\in I.

Moreover, taking into account that

(Kϵw^xϵ)x=1ϵγKϵf^ϵ+Kϵw^ϵ,(K_{\epsilon}\hat{w}^{\epsilon}_{x})_{x}=-\frac{1}{\epsilon^{\gamma}}K_{\epsilon}\hat{f}^{\epsilon}+K_{\epsilon}{\hat{w}}^{\epsilon},

we deduce that Kϵw^xϵK_{\epsilon}\hat{w}^{\epsilon}_{x} is uniformly bounded in H1(I).H^{1}(I). Then, it follows that there exists a function σ\sigma such that, up to subsequences,

Kϵw^xϵσstrongly in L2(I).K_{\epsilon}\hat{w}^{\epsilon}_{x}\longrightarrow\sigma\quad\hbox{strongly in }L^{2}(I).

Thus, using (H.4) we have

w^xϵ=1Kϵ(Kϵw^xϵ)ϵ0Pσ wL2(I).\hat{w}^{\epsilon}_{x}=\frac{1}{K_{\epsilon}}\Big{(}K_{\epsilon}\hat{w}^{\epsilon}_{x}\Big{)}\,{\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}}\,P\sigma\quad\hbox{ w}-L^{2}(I).

Consequently, due to convergence (37) we have

w^x=Pσ,\hat{w}_{x}=P\sigma,

or equivalently,

(38) Kϵw^xϵϵ01Pw^xstrongly in L2(I).K_{\epsilon}\hat{w}^{\epsilon}_{x}\buildrel\epsilon\to 0\over{\longrightarrow}\frac{1}{P}\hat{w}_{x}\quad\hbox{strongly in }L^{2}(I).

Therefore, in view of (35), (37) and (38) we can pass to the limit in (36) and we obtain the following weak formulation

I{1Pw^xϕx+(K1+K2)w^ϕ}𝑑x=If0ϕ𝑑x.\int_{I}\Big{\{}\frac{1}{P}\hat{w}_{x}\phi_{x}+(K^{1}+K^{2})\hat{w}\phi\Big{\}}\,dx=\int_{I}f_{0}\phi\,dx.

With this, Theorem 2.1 is proven.

5 Quasi-periodic and almost periodic setting

In this section, we will analyze specific and interesting cases that satisfy the hypotheses (H). We will begin by examining the scenario where the oscillating boundary is given by quasi-periodic functions. In particular, we consider

(39) kϵ1(x)=h(x/ϵα),kϵ2(x)=g(x/ϵβ),k^{1}_{\epsilon}(x)=h(x/\epsilon^{\alpha}),\quad k^{2}_{\epsilon}(x)=g(x/\epsilon^{\beta}),

where 0<α,β<10<\alpha,\beta<1 and the functions g,h:g,h\,:\mathbb{R}\to\mathbb{R} are C1C^{1} quasi-periodic functions verifying

(40) 0h0h()h1,\displaystyle 0\leq h_{0}\leq h(\cdot)\leq h_{1},
(41) 0<g0g()g1.\displaystyle 0<g_{0}\leq g(\cdot)\leq g_{1}.

Therefore g,h:g,h\,:\mathbb{R}\to\mathbb{R} present a combination of multiple frequencies that are rationally independent. That is, there exist two C1C^{1} periodic functions h¯:n\bar{h}\,:\mathbb{R}^{n}\to\mathbb{R} and g¯:m\bar{g}\,:\mathbb{R}^{m}\to\mathbb{R} , n,mn,m\in\mathbb{N}, such that

(42) h(x)=h¯(x,,xn),g(x)=g¯(x,,xm).h(x)=\bar{h}(\underbrace{x,\cdots,x}_{n}),\quad g(x)=\bar{g}(\underbrace{x,\cdots,x}_{m}).

Notice that h¯\bar{h} and g¯\bar{g} are periodic with respect to each of their arguments. For example in the case of h¯\bar{h}, for all 1jn1\leq j\leq n, there exists Lj>0L_{j}>0 such that h¯(x1,,xj+Lj,,xn)=h¯(x1,,xj,,xn)\bar{h}(x_{1},\cdots,x_{j}+L_{j},\cdots,x_{n})=\bar{h}(x_{1},\cdots,x_{j},\cdots,x_{n}) for all xnx\in\mathbb{R}^{n}. Analogously, this holds for g¯\bar{g}.

We define QP(L)QP(L) as the set of quasi-periodic functions associated with L=(L1,,Ln)L=(L_{1},\cdots,L_{n}). The positive numbers L1,,LnL_{1},\cdots,L_{n} are referred to as quasi-periods. Therefore we will assume that hQP(Lh)h\in QP(L^{h}) and gQP(Lg)g\in QP(L^{g}), where LhL^{h} and LgL^{g} are vectors in n\mathbb{R}^{n} and m\mathbb{R}^{m}, respectively.

It is noteworthy that when n=1n=1 and m=1m=1, the scenario reverts to the conventional periodic case.Furthermore, to denote periodic functions associated with quasi-periodic functions, we will use letters with bars.

It is not restrictive to assume that the associated frequencies to the quasi-periods L=(L1,,Ln)L=(L_{1},\cdots,L_{n}) are linearly independet on \mathbb{Z}. Under this assumption, Kronecker’s Lemma (see Appendix of [17]) guarantees that h¯\bar{h} and g¯\bar{g} are uniquely determined by hh and gg respectively.

Using properties of quasi-periodic functions we obtain explicitly the homogenized limit problem of (11) when the oscillating boundaries are given by (39). We just need to prove that these particular functions kϵ1k^{1}_{\epsilon} and kϵ2k^{2}_{\epsilon} satisfy hypothesis (H).

First, notice that

ϵkϵ1x(x)=ϵ1αhx(xϵα),ϵkϵ2x(x)=ϵ1βgx(xϵβ).\epsilon\frac{\partial k^{1}_{\epsilon}}{\partial x}(x)=\epsilon^{1-\alpha}\frac{\partial h}{\partial x}\Big{(}\frac{x}{\epsilon^{\alpha}}\Big{)},\quad\epsilon\frac{\partial k^{2}_{\epsilon}}{\partial x}(x)=\epsilon^{1-\beta}\frac{\partial g}{\partial x}\Big{(}\frac{x}{\epsilon^{\beta}}\Big{)}.

Then, since 0<α,β<10<\alpha,\beta<1 we directly have (H.1).

As (H.2) is immediately verified by hypothesis, we will focus on proving weak convergences (H.3) and (H.4) for quasi-periodic functions with multiple scales.

Taking into account the definition of gg and hh we can assume that K(x)=g(x)+h(x)K(x)=g(x)+h(x) is a quasi-periodic function with quasi-periods L=(L1,L2,,Ln+m)L=(L_{1},L_{2},\cdots,L_{n+m}) where the quasi-periods are not necessarily rationally independent. Therefore, there exists a LL-periodic function K¯:n+m\bar{K}:\mathbb{R}^{n+m}\rightarrow\mathbb{R} such that KK is the trace of K¯\bar{K} in the sense of

K(x)=K¯(x,x,,x),x.K(x)=\bar{K}(x,x,\cdots,x),\quad\forall x\in\mathbb{R}.

In addition, KK has an average, see Proposition 1.2 in [15], which is defined as follows

μ(K)=limT12TTT(g(y)+h(y))𝑑y=1|I(L)|I(L)K¯(x1,,xn+m)𝑑x1𝑑xn+m,{\displaystyle\mu(K)=\lim_{T\to\infty}\frac{1}{2T}\int_{-T}^{T}\big{(}g(y)+h(y)\big{)}\,dy=\frac{1}{|I(L)|}\int_{I(L)}\bar{K}(x_{1},\cdots,x_{n+m})\,dx_{1}\cdots dx_{n+m},}

where I(L)=(0,L1)×(0,L2)××(0,Ln+m)I(L)=(0,L_{1})\times(0,L_{2})\times\cdots\times(0,L_{n+m}).

Since K()>0K(\cdot)>0 we can conclude that 1K\frac{1}{K} belongs to QP(L)QP(L) and its average is given by

μ(1K)=limT12TTT1g(y)+h(y)𝑑y=1|I(L)|I(L)1K¯(x1,,xn+m)𝑑x1𝑑xn+m.{\displaystyle\mu\Big{(}\frac{1}{K}\Big{)}=\lim_{T\to\infty}\frac{1}{2T}\int_{-T}^{T}\frac{1}{g(y)+h(y)}\,dy=\frac{1}{|I(L)|}\int_{I(L)}\frac{1}{\bar{K}(x_{1},\cdots,x_{n+m})}\,dx_{1}\cdots dx_{n+m}.}

Below we show the convergence obtained for quasi-periodic functions with different oscillation scales.

Proposition 1.

Let Fϵ:F_{\epsilon}:\mathbb{R}\rightarrow\mathbb{R} be a quasi-periodic function defined by

Fϵ(x)=F¯(xϵα,xϵα,xϵα,xϵβ,xϵβ),F_{\epsilon}(x)=\bar{F}\Big{(}\frac{x}{\epsilon^{\alpha}},\frac{x}{\epsilon^{\alpha}},\cdots\frac{x}{\epsilon^{\alpha}},\frac{x}{\epsilon^{\beta}}\cdots,\frac{x}{\epsilon^{\beta}}\Big{)},

where F¯n+m\bar{F}\in\mathbb{R}^{n+m} is a LL-periodic function. The following weak convergence holds

(43) Fϵϵ0μ(F¯)wL2(I).F_{\epsilon}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}\mu(\bar{F})\quad w-L^{2}(I).
Proof.

We just prove the result for FϵF_{\epsilon} a trigonometric polinomial of quasi-periods LL since the set of trigonometric polynomials of quasi-periods LL is dense en QP(L)QP(L) for the uniform norm, see [15]. Therefore, we have

Fϵ(x)=kn+make2πij=1nkjxLjϵαe2πij=n+1n+mkjxLjϵβ,F_{\epsilon}(x)=\sum_{k\in\mathbb{Z}^{n+m}}a_{k}\,e^{2\pi i\sum_{j=1}^{n}\frac{k_{j}x}{L_{j}\epsilon^{\alpha}}}e^{2\pi i\sum_{j=n+1}^{n+m}\frac{k_{j}x}{L_{j}\epsilon^{\beta}}},

where the sequence of (ak)kN(a_{k})_{k\in\mathbb{Z}^{N}} vanishes except for a finite number of values of kk. Notice that, for any interval I=(a,b)I=(a,b) we have

abFϵ(x)𝑑x\displaystyle\int_{a}^{b}F_{\epsilon}(x)dx =(ba)a0+ϵα+βk0ake2πij=1nkjbLjϵαe2πij=n+1n+mkjbLjϵβ2iπ(ϵβλk+ϵαμk)\displaystyle=(b-a)a_{0}+\epsilon^{\alpha+\beta}\sum_{k\neq 0}\frac{a_{k}e^{2\pi i\sum_{j=1}^{n}\frac{k_{j}b}{L_{j}\epsilon^{\alpha}}}e^{2\pi i\sum_{j=n+1}^{n+m}\frac{k_{j}b}{L_{j}\epsilon^{\beta}}}}{2i\pi(\epsilon^{\beta}\lambda_{k}+\epsilon^{\alpha}\mu_{k})}
ϵα+βk0ake2πij=1nkjaLjϵαe2πij=n+1n+mkjaLjϵβ2iπ(ϵβλk+ϵαμk)ϵ0(ba)a0=(ba)μ(F),\displaystyle-\epsilon^{\alpha+\beta}\sum_{k\neq 0}\frac{a_{k}e^{2\pi i\sum_{j=1}^{n}\frac{k_{j}a}{L_{j}\epsilon^{\alpha}}}e^{2\pi i\sum_{j=n+1}^{n+m}\frac{k_{j}a}{L_{j}\epsilon^{\beta}}}}{2i\pi(\epsilon^{\beta}\lambda_{k}+\epsilon^{\alpha}\mu_{k})}\buildrel\epsilon\to 0\over{\longrightarrow}(b-a)a_{0}=(b-a)\mu(F),

where λk=j=1nkjLj\lambda_{k}=\sum_{j=1}^{n}\frac{k_{j}}{L_{j}} y μk=j=n+1n+mkjLj\mu_{k}=\sum_{j=n+1}^{n}+m\frac{k_{j}}{L_{j}}.

Thus, for any piecewise constant compactly supported function φ\varphi we obtain

Fϵφϵ0φμ(F).\int F_{\epsilon}\varphi\buildrel\epsilon\to 0\over{\longrightarrow}\varphi\mu(F).

Using that piecewise constant functions are dense in L2()L^{2}(\mathbb{R}) we have the result. ∎

Consequently, we obtain that (H.3) and (H.4) are satisfied for the particular case where both oscillating boundaries are given by quasi-periodic functions.

Corollary 1.

Let Kϵ(x)=h(xϵα)+g(xϵβ)K_{\epsilon}(x)=h\big{(}\frac{x}{\epsilon^{\alpha}}\big{)}+g\big{(}\frac{x}{\epsilon^{\beta}}\big{)} be a quasi-periodic function with two different scales, we have

Kϵϵ0μ(K)wL2(I),1Kϵ=1g(ϵβ)+h(ϵα)ϵ0μ(1K)wL2(I).K_{\epsilon}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}\mu(K)\quad\text{w}-L^{2}(I),\quad\frac{1}{K_{\epsilon}}=\frac{1}{g\big{(}\frac{\cdot}{\epsilon^{\beta}}\big{)}+h\big{(}\frac{\cdot}{\epsilon^{\alpha}}\big{)}}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}\mu\Big{(}\frac{1}{K}\Big{)}\quad\hbox{w}-L^{2}(I).
Proof.

Trivial from the previous Theorem. ∎

Therefore, we are in conditions of Theorem 2.1 and with the definition of f0f_{0} given by (35) and denoting by f^=f0μ(g)+μ(h)\hat{f}=\frac{f_{0}}{\mu(g)+\mu(h)}, then we have we obtain the following convergence result for the particular case of quasi-periodic functions:

w^ϵw^, wH1(I),\hat{w}^{\epsilon}\to\hat{w},\hbox{ w}-H^{1}(I),
|wϵw^|L2(Rϵ)0,|||w^{\epsilon}-\hat{w}|||_{L^{2}(R^{\epsilon})}\to 0,

where w^H1(I)\hat{w}\in H^{1}(I) is the weak solution of the following Neumann problem

{1P(μ(g)+μ(h))w^xx+w^=f^,xI,w^(0)=w^(1)=0,\left\{\begin{gathered}-\frac{1}{P{\big{(}\mu}(g)+{\mu}(h)\big{)}}{\hat{w}}_{xx}+\hat{w}=\hat{f},\quad x\in I,\\ \hat{w}^{\prime}(0)=\hat{w}^{\prime}(1)=0,\end{gathered}\right.

where the constant PP is such that

1h(xϵα)+g(xϵβ)ϵ0PwL2(I).\frac{1}{h\Big{(}\frac{x}{\epsilon^{\alpha}}\Big{)}+g\Big{(}\frac{x}{\epsilon^{\beta}}\Big{)}}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}P\quad\text{w}-L^{2}(I).

Therefore PP is given by

P=1|I(L)|I(L)1F(x1,,xm+n)𝑑x1𝑑xm+n.P=\frac{1}{|I(L)|}\int_{I(L)}\frac{1}{F(x_{1},\cdots,x_{m+n})}\,dx_{1}\cdots dx_{m+n}\,.
Remark 4.

Notice what happend in the particular case where gg and hh are L1L_{1}-periodic and L2L_{2}-periodic respectively. In [3] the authors have the following convergence

1Gϵ=1g(ϵβ)+h(ϵα)ϵ01p0,\frac{1}{G_{\epsilon}}=\frac{1}{g(\frac{\cdot}{\epsilon^{\beta}})+h(\frac{\cdot}{\epsilon^{\alpha}})}\stackrel{{\scriptstyle\epsilon\to 0}}{{\rightharpoonup}}\frac{1}{p_{0}},

where p0p_{0} is defined as follows

1p0={limT1T0T1g(y)+h(y)𝑑y,if α=β,1L1L20L10L21g(y)+h(z)𝑑z𝑑y,if αβ.\displaystyle\frac{1}{p_{0}}=\left\{\begin{array}[]{ll}{\displaystyle\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\frac{1}{g(y)+h(y)}\,dy,\quad\hbox{if }\alpha=\beta,}\\ {\displaystyle\frac{1}{L_{1}L_{2}}\int_{0}^{L_{1}}\int_{0}^{L_{2}}\frac{1}{g(y)+h(z)}\,dzdy,\quad\hbox{if }\alpha\neq\beta.}\end{array}\right.

Notice that this result is in complete accordance with the previous proposition. For this particular case, it is found that P=1p0P=\frac{1}{p_{0}}. In fact, for the case αβ\alpha\neq\beta and for α=β\alpha=\beta we have:

P\displaystyle\displaystyle P =μ(1g+h)=limT1T0T1g(y)+h(y)𝑑y\displaystyle=\mu\Big{(}\frac{1}{g+h}\Big{)}=\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\frac{1}{g(y)+h(y)}\,dy
=1L1L20L10L21g(y)+h(z)𝑑z𝑑y=1p0.\displaystyle=\frac{1}{L_{1}L_{2}}\int_{0}^{L_{1}}\int_{0}^{L_{2}}\frac{1}{g(y)+h(z)}\,dzdy=\frac{1}{p_{0}}.

We can also write Theorem 2.1 for almost periodic functions in the sense of Besicovitch, see [14]. We just have to take into account that the set of almost periodic functions in \mathbb{R} is the closure of the set of trigonometric polynomial for the the mean square norm (or Besicovitch norm), defined for a PnP_{n} trigonometric polynomial as follows:

Pn2=(lim supT12TTT|Pn(t)|2𝑑t)1/2.\|P_{n}\|_{2}=\left(\limsup_{T\to\infty}\frac{1}{2T}\int_{-T}^{T}|P_{n}(t)|^{2}\,dt\right)^{1/2}.

Therefore, it is obvious, by means of the previous approximation, that any almost periodic oscillatory functions satisfy hipothesis (H). Therefore, if the thin domain is given by two almost periodic functions gg and hh as follows:

Rϵ={(x,y)2|xI,ϵg(x/ϵα)<y<ϵh(x/ϵβ)}.R^{\epsilon}=\Big{\{}(x,y)\in\mathbb{R}^{2}\;|\;x\in I\subset\mathbb{R},\;-\epsilon g(x/\epsilon^{\alpha})<y<\epsilon h(x/\epsilon^{\beta})\Big{\}}.

Then, in condictions of Theorem 2.1 and denoting by f^=f0μ(g)+μ(h)\hat{f}=\frac{f_{0}}{\mu(g)+\mu(h)}, we can guarantee:

w^ϵw^, wH1(I),\hat{w}^{\epsilon}\to\hat{w},\hbox{ w}-H^{1}(I),
|wϵw^|L2(Rϵ)0,|||w^{\epsilon}-\hat{w}|||_{L^{2}(R^{\epsilon})}\to 0,

where w^H1(I)\hat{w}\in H^{1}(I) is the weak solution of the following Neumann problem

{1P(μ(g)+μ(h))w^xx+w^=f^,xI,w^(0)=w^(1)=0,\left\{\begin{gathered}-\frac{1}{P{\big{(}\mu}(g)+{\mu}(h)\big{)}}{\hat{w}}_{xx}+\hat{w}=\hat{f},\quad x\in I,\\ \hat{w}^{\prime}(0)=\hat{w}^{\prime}(1)=0,\end{gathered}\right.

where the constant P=lim supT12TTT1g(t)+h(t)𝑑t.P=\limsup_{T\to\infty}\frac{1}{2T}\int_{-T}^{T}\frac{1}{g(t)+h(t)}\,dt.

6 Numerical evidences

In this section, we numerically investigate the behavior of the solutions to equation (2) as ε\varepsilon approaches zero, where the thin domain is defined by the graph of a quasi-periodic function, see Figure 4.

Then, we consider the following particular thin domain:

(45) Rϵ={(x,y)2|x(0,20),ϵ(8sin(xϵ5)sin(xπ8ϵ5))<y<ϵ(8+sin(xϵ5)+sin(xπ8ϵ5))}.\begin{split}R^{\epsilon}=\Big{\{}(x,y)\in\mathbb{R}^{2}\;|\;x\in(0,20),\;&-\epsilon\Big{(}8-\sin\Big{(}\frac{x}{\sqrt[5]{\epsilon}}\Big{)}-\sin\Big{(}\frac{x\pi}{8\sqrt[5]{\epsilon}}\Big{)}\Big{)}<y\\ &<\epsilon\Big{(}8+\sin\Big{(}\frac{x}{\sqrt[5]{\epsilon}}\Big{)}+\sin\Big{(}\frac{x\pi}{8\sqrt[5]{\epsilon}}\Big{)}\Big{)}\Big{\}}.\end{split}

Moreover, the narrow strip θϵ\theta^{\epsilon} is defined by

θϵ={(x,y)2|x(0,20),\displaystyle\theta^{\epsilon}=\Big{\{}(x,y)\in\mathbb{R}^{2}\;|\;\;x\in(0,20),\; ϵ(8+sin(xϵ5)+sin(xπ8ϵ5)ϵ18(2+sin(xϵ3)))<y\displaystyle\epsilon\Big{(}8+\sin\Big{(}\frac{x}{\sqrt[5]{\epsilon}}\Big{)}+\sin\Big{(}\frac{x\pi}{8\sqrt[5]{\epsilon}}\Big{)}-\sqrt[18]{\epsilon}\Big{(}2+\sin\Big{(}\frac{x}{\sqrt[3]{\epsilon}}\Big{)}\Big{)}\Big{)}<y
<ϵ(8+sin(xϵ5)+sin(xπ8ϵ5))}.\displaystyle<\epsilon\Big{(}8+\sin\Big{(}\frac{x}{\sqrt[5]{\epsilon}}\Big{)}+\sin\Big{(}\frac{x\pi}{8\sqrt[5]{\epsilon}}\Big{)}\Big{)}\Big{\}}.
Refer to caption
Figure 4: Thin domain RϵR^{\epsilon} featuring the slender strip θϵ\theta^{\epsilon} for ϵ=1\epsilon=1.

The problem was discretized using a triangular mesh that is finer in the narrow strip, see Figure 5.

Refer to caption
Figure 5: Triangular mesh for the slender domain with finer density in the strip.

We analyze the behavior of the solutions as ϵ\epsilon tends to zero taking the forcing term ff as f(x)=1+sin(x)f(x)=1+sin(x).

Refer to caption
(a) ε=0.1\varepsilon=0.1
Refer to caption
(b) ε=0.04\varepsilon=0.04
Refer to caption
(c) ε=0.0008\varepsilon=0.0008
Figure 6: Contour levels of solutions for some values of ϵ\epsilon.

First we show a color map of the solutions with the corresponding contours levels for different values of ϵ\epsilon. For higher values of ϵ\epsilon, the dependence of the solutions on two dimensions can be seen, with significant variations in yy due to the applied force in the boundary strip.It can be clearly observed, Figure 6, that as ϵ\epsilon becomes small, the dependence of the vertical variable on the solutions disappears. This is consistent with the fact that the domain is thinning out, progressively shrinking in the vertical direction. This fact can also be observed in Figure 7 where the solution was represented for two values of ϵ\epsilon. While for ϵ=0.2\epsilon=0.2, it is clear that the solution undergoes dramatic changes from y=2y=-2 to y=2y=2, for ϵ=0.05\epsilon=0.05, it essentially maintains the same profile.

Refer to caption
(a) ε=0.2\varepsilon=0.2
Refer to caption
(b) ε=0.05\varepsilon=0.05
Figure 7: Solution for some values of ϵ\epsilon.

This phenomenon becomes more apparent in Figure 8, where slices of the solutions at y=0.2,0,0.2y=-0.2,0,0.2 for various ϵ\epsilon values are displayed. It is observed that as ϵ\epsilon decreases, the similarity among the slices increases significantly.

Refer to caption
(a) ε=0.2\varepsilon=0.2
Refer to caption
(b) ε=0.1\varepsilon=0.1
Refer to caption
(c) ε=0.07\varepsilon=0.07
Refer to caption
(d) ε=0.05\varepsilon=0.05
Figure 8: Solution for different values of yy.

This leads us to attempt to compare the solution for increasingly smaller values of ϵ\epsilon with the solution of the limit problem. In Figure 9, we represent the solutions for different values of ϵ\epsilon with the vertical variable fixed at y=0y=0, along with the solution to the limit equation. It appears to be inferred that, as expected, as ϵ\epsilon becomes smaller, the solution of the problem in the thin domain increasingly resembles the limit solution.

Refer to caption
Figure 9: uϵ(x,0)u^{\epsilon}(x,0) and the limit solution u(x)u(x).

Since Theorem 2.1 establishes strong convergence of the solutions to the limit problem’s solution, we have also examined the L2L^{2} norm of the difference for various values of ϵ\epsilon. In particular, for ϵ=0.1\epsilon=0.1 we get that the norm of the difference in the thin domain is 1.78531.7853, for ϵ=0.08\epsilon=0.08 is 1.0320521.032052 and if ϵ=0.04\epsilon=0.04 the norm of the difference is 0.4676110.467611. Therefore, the error norm appears to decrease linearly.

A natural question is whether such approximation results can be improved in order to describe the asymptotic behavior of the Dynamical System generated by the parabolic equation associated with (2) posed in more general thin regions of n\mathbb{R}^{n}. It is our goal to investigate this question in a forthcoming paper.

References

  • [1] G. Allaire “Homogenization and two-scale convergence” In SIAM J. Math. Anal. 23, 1992, pp. 1482–1518
  • [2] S. Armstrong, A. Gloria and T. Kuusi “Bounded Correctors in Almost Periodic Homogenization” In Arch Rational Mech Anal 222, 2016, pp. 393–426 DOI: 10.1007/s00205-016-1004-0
  • [3] J.. Arrieta and M. Villanueva-Pesqueira “Elliptic and parabolic problems in thin domains with doubly weak oscillatory boundary” In Communications on Pure and Applied Analysis 19.4, 2020, pp. 1891–1914 DOI: 10.3934/cpaa.2020083
  • [4] J.. Arrieta and M. Villanueva-Pesqueira “Thin domains with non-smooth periodic oscillatory boundaries” In Journal of Mathematical Analysis and Applications 446.1, 2017, pp. 30–164
  • [5] J.. Arrieta, A.. Carvalho, M.. Pereira and R.. Silva “Semilinear parabolic problems in thin domains with a highly oscillatory boundary” In Nonlinear Analysis: Theory, Methods and Applications 74.15, 2011, pp. 5111–5132
  • [6] J.M. Arrieta, A. Jiménez-Casas and A. Rodrígues-Bernal “Flux terms and Robin boundary conditions as limit of reactions and potentials concentrating at the boundary” In Rev. Mat. Iberoam. 24.1, 2008, pp. 183–211
  • [7] J.M. Arrieta, A. Nogueira and M.C. Pereira “Nonlinear elliptic equations with concentrating reaction terms at an oscillatory boundary” In Discrete Contin. Dyn. Syst. B 24.8, 2019, pp. 4217–4246
  • [8] J.M. Arrieta, A. Nogueira and M.C. Pereira “Semilinear elliptic equations in thin regions with terms concentrating on oscillatory boundaries” In Comput. Math. Appl. 77, 2019, pp. 536–554
  • [9] J.M. Arrieta and M.. Pereira “Homogenization in a thin domain with an oscillatory boundary” In Journal de Mathematiques Pures et Apliquees 96.1, 2011, pp. 29–57
  • [10] J.M. Arrieta and M.C. Pereira “The Neumann problem in thin domains with very highly oscillatory boundaries” In Journal of Mathematical Analysis and Applications 444.1, 2013, pp. 86–104
  • [11] J.M. Arrieta and M. Villanueva-Pesqueira “Thin domains with doubly oscillatory boundary” In Mathematical Methods in Applied Science 37.2, 2014, pp. 158–166
  • [12] J.M. Arrieta and M. Villanueva-Pesqueira “Unfolding Operator Method for Thin Domains with a Locally Periodic Highly Oscillatory Boundary” In SIAM Journal on Mathematical Analysis 48.3, 2016
  • [13] A. Bensoussan, J.. Lions and G. Papanicolaou “Asymptotic Analysis for Periodic Structures” North-Holland Publ. Company, 1978
  • [14] A.S. Besicovitch “Almost periodic functions” Dover, 1954
  • [15] X. Blanc and C. Bris “Homogenization Theory for Multiscale Problems: An introduction” Springer Nature Switzerland, 2023
  • [16] Andrea Braides “Almost periodic methods in the theory of homogenization” In Applicable Analysis 47.1-4, 1992, pp. 259–277 DOI: 10.1080/00036819208840144
  • [17] Andrea Braides, Valeria Chiadó Piat and Anneliese Defranceschi “Homogenization of almost periodic monotone operators” In Annales de l’I.H.P. Analyse non linéaire 9.4, 1992, pp. 399–432
  • [18] D. Cioranescu, A. Damlamian and G. Griso “The periodic unfolding method: theory and applications to partial differential problems” 3, Series in Contemporary Mathematics Springer, Singapore, 2018
  • [19] D. Cioranescu and J. Jean Paulin “Homogenization of Reticulated Structures” Springer Verlag, 1999
  • [20] J.. Hale and G. Raugel “Reaction-diffusion equation on thin domains” In J. Math. Pures and Appl. 71.9, 1992, pp. 33–95
  • [21] A. Jiménez-Casas and A. Rodríguez-Bernal “Asymptotic behavior of a parabolic problem with terms concentrated in the boundary” In Nonlinear Analysis: Theory, Methods and Applications 71.12, 2009, pp. e2377–e2383
  • [22] A. Jiménez-Casas and A. Rodríguez-Bernal “Singular limit for a nonlinear parabolic equation with terms concentrating on the boundary” In Journal of Mathematical Analysis and Applications 379.2, 2011, pp. 567–588 DOI: 10.1016/j.jmaa.2011.01.051
  • [23] S.M. Kozlov “Averaging Differential Operators with Almost-Periodic Rapidly Oscillating Coefficients” In Math. USSR-Sb. 35, 1979, pp. 481–498
  • [24] J.C. Nakasato and M.C. Pereira “An optimal control problem in a tubular thin domain with rough boundary” In Journal of Differential Equations 313, 2022, pp. 188–243 DOI: 10.1016/j.jde.2021.12.021
  • [25] O.. Oleinik and V.V. Zhikoz “On the homogenization of elliptic operators with almost-periodic coefficients” In Rend. Sem. Mat. Fis. Milano 52, 1982, pp. 149–166
  • [26] M. Prizzi, M. Rinaldi and K.. Rybakowski “Curved thin domains and parabolic equations” In Studia Mathematica 151, 2002, pp. 109–140
  • [27] G. Raugel “Dynamics of partial differential equations on thin domains” In Lecture Notes in Math. 1609 Springer, Berlin, 1995, pp. 208–315
  • [28] Jinping Zhuge “Uniform boundary regularity in almost-periodic homogenization” In Journal of Differential Equations 262.1, 2017, pp. 418–453 DOI: 10.1016/j.jde.2016.09.031